Overview - College of Engineering and Applied Science

advertisement
8 MICROPROPULSON
Andrew D. Ketsdever
Propulsion Directorate, United States Air Force Research Laboratory, Edwards AFB, CA, USA
KEYWORDS: Micropropulsion, propulsion, microspacecraft, microscale, system specific impulse.
8.1 Introduction
In its broadest definition, a micropropulsion system is one capable of producing a change in momentum
of a microspacecraft. (Ketsdever, 2000) The United States Air Force Research Laboratory (AFRL) defines
a microspacecraft as having a total mass of less than 100 kg. This broad definition of micropropulsion
allows the inclusion of a wide range of concepts ranging from microelectromechanical systems (MEMS)
fabricated thrusters with micrometer characteristic sizes to simply scaled-down versions of macro-scale
thrusters. Regardless of the characteristic size of a particular micropropulsion system, severe design
constraints will be imposed by resource limited microspacecraft, including:




Mass
Volume
Power
Maximum voltage
Micropropulsion concepts will not only have to address microspacecraft mission requirements, but will
also have to fit within the systems constraints imposed. Although this is true of all spacecraft
subsystems regardless of spacecraft mission or size, the constraints of some microspacecraft may pose
such severe design restraints that new paradigms will need to be developed. A complication to the
design process for micropropulsion systems will be the fact that, in many cases, microspacecraft
missions will require exceedingly low thrust levels, accurate and very low minimum impulse bits, and
low thrust noise. Also, many microspacecraft missions currently under investigation require relatively
large V, indicating that high specific impulse micropropulsion systems are necessary to perform these
missions under strict mass constraints.
A wide array of micropropulsion systems have been investigated, designed, built, and even flown. As
with larger spacecraft propulsion systems, these micropropulsion concepts incorporate both chemical
and electric means of producing high speed propellant species to create thrust. Both solid and liquid
chemical thrusters have been developed for use on microspacecraft missions where relatively large
thrust levels are required. Electric propulsion thrusters including micro-electrothermal, electrostatic,
and electromagnetic systems have also been investigated for missions where higher specific impulse is
desired. Electrothermal thrusters use electrical means to heat a propellant, which include arcjets [See
2.4.02 Arcjets] and resistojets. Electrostatic thrusters include a class of propulsion systems that use
electrostatic means to accelerate ions formed in a plasma discharge [See 2.4.05 Ion Thrusters].
Electromagnetic thrusters achieve acceleration of plasma propellants through a combined interaction
between electric and magnetic fields [See 2.4.03 Hall Thrusters and 2.4.04 Magnetoplasmadynamic
1
Thrusters]. A comprehensive survey of micropropulsion systems under development has been compiled
by Mueller (2000). Each system brings unique design challenges including:







Microscale combustion, heat transfer, and fluid dynamics
Microscale plasma formation
Material compatibility
Micro-valve leakage
Passage clogging
System reliability and lifetime
Manufacturing and integration complexities
Many of these design challenges will be addressed in the following sections.
Micropropulsion is considered an enabling technology for microspacecraft by making possible missions
which otherwise could not be performed. However, micropropulsion could also find application on
larger spacecraft to compensate for very small disturbing torques (e.g. solar pressure), providing
extremely fine attitude control. Several propulsive requirements for potential missions are given in the
following section. Reducing the physical size, operating power, and in some cases operating pressure
and temperature, may lead to significant reductions in operational performance and efficiency. For
example, a microchemical propulsion system will require high pressure operation for efficient
combustion at small physical scales. However, a particular microspacecraft mission may require low
thrust levels indicating that a micronozzle must be employed with a small throat diameter. High
temperatures along with small nozzle throat diameters will lead to a reduction in the operational
Reynolds number, leading to significant viscous losses. High temperature microchemical systems will
also suffer from high heat transfer rates which will further reduce the efficiency of these systems,
particularly in MEMS fabricated thrusters where the thermal conductivity of silicon-based materials is
quite high. On a positive note, there is significant potential for improvement in all manner of
micropropulsion systems, leaving future micropropulsion developers many fruitful areas of basic
research and applied engineering practice.
8.2 Sample Microspacecraft Missions and Requirements
Constellations or platoons of microsatellites working in collaboration have been envisioned to perform
the functions of larger spacecraft in a distributed way. Microspacecraft constellations will benefit from
increased survivability, flexibility in mission performance, distributed functionality, and more graceful
degradation of overall capability. Partitioning the functions of a single large spacecraft into a number of
smaller satellites operating cooperatively in close proximity or in widely dispersed orbits could lead to
significant improvement in mission performance. In some cases, this approach enables missions that
cannot be performed by a single satellite of any size. AFRL has proposed a standard definition of
spacecraft with masses below 100kg as detailed in Table 1.
By the AFRL definition, the first microspacecraft launched into orbit from Earth was also the first manmade object launched, Sputnik. Sputnik had a mass of approximately 84 kg. Figure 1 shows Sputnik I,
2
which was launched on October 4, 1957. Since the late 1950’s, trends in electronics, material sciences,
and general mechanics have led to advances in micro-scale components. This has led to a significant
increase in the performance of micro-scale systems which have been used on microspacecraft, making
today’s small spacecraft far more capable then their early predecessors. Today, universities are capable
of launching very sophisticated microsatellites. (Sweeting, 2002) Sputnik was a fairly simple radio
transmitter launched into orbit on a converted SS-6 nuclear missile, while today microsatellites with
sophisticated communication, attitude control, and propulsion systems are being designed, developed,
and launched.
Table 1: AFRL satellite classification
Total Spacecraft Mass
10-100 kg
1-10 kg
<1 kg
Description
Microspacecraft
Nanospacecraft
Picospacecraft
Figure 2 shows FalconSat 3 designed and developed by the United States Air Force Academy (USAFA).
FalconSat 3 was launched in March, 2007 as part of the secondary payload program for the Evolved
Expendable Launch Vehicles (EELVs). Fifty years after the launch of Sputnik, a highly capable
microsatellite with a mass of approximately 50 kg had payloads which demonstrated an advanced
micropropulsion attitude control system and an ability to characterize the ionosphere and local
spacecraft plasma environment.
Figure 1: Sputnik I launched October 4, 1957 by the Soviet Union (Photo courtesy of NASA).
3
Figure 2: FalconSat 3 launched March 8, 2007 (50 years after Sputnik I) by the United States Air Force
Academy (Photo courtesy of the Department of Astronautics, USAFA).
Several studies (Janson et al., 1999; Mueller, 2000; Schilling, Spores and Spanjers, 2000; Ziemer et al.,
2008) have provided propulsive requirements for many microspacecraft missions including Earth
observation, communications, space environmental sciences, satellite servicing, satellite inspection, and
technology demonstration. For the proposed TechSat21 mission, which has since been canceled by
AFRL but is still a good example of microspacecraft propulsive requirements, the proposed total V was
390 m/s of which 50 m/s was required for orbit raising, 20 m/s for drag makeup, 200 m/s for
stationkeeping over a 10 year mission lifetime, and 120 m/s for a deorbit maneuver at the spacecraft’s
end of life (EOL).(Schilling, Spores and Spanjers, 2000) A minimum impulse bit of 2mN-s was required for
close proximity maneuvers. An early version of the TechSat 21 experiment called for formation flying
maneuvers over a 30 day period with a V requirement of 30 m/s and a minimum thrust level of 2mN.
This would have allowed for a series of formations to be flown with separation distances between 5 m
and 5 km. Mueller (2000) also gives a range of propulsive requirements for microspacecraft slew
maneuvers with minimum impulse bits ranging from 10-8 to 10-3 N-s.
Janson et al. (1999) gives a good overview of potential micropropulsion maneuvers and the V required.
The V requirement for a microspacecraft to perform a minor altitude raising maneuver of 1 km at an
initial altitude of 700 km was calculated to be about 0.5 m/s for a 10 minute firing time. For North-South
stationkeeping in GEO for approximately 1 week duration the V requirement is over 300 m/s. A
deorbit burn at an initial altitude of 700 km requires a V of over 330 m/s for a low thrust spiral
maneuver over a 1 week period. On-orbit servicing or inspection missions may require up to 500 m/s of
V for orbit changing and attitude control. Many microspacecraft missions currently under
consideration will have relatively large V requirements, indicating that the specific impulse and
efficiency of micropropulsion systems will need to be correspondingly large. However, a large number
of microspacecraft missions will also require a rapid response to changing conditions to be effective.
4
These missions will require relatively large thrust levels, perhaps as large as 100’s of mN. As with their
larger counterparts, micropropulsion systems will continue to seek high thrust, high specific impulse
solutions.
The Laser Interferometer Space Antenna (LISA) mission will require precision formation flying to detect
gravitational waves produced by processes in the universe. To perform this mission and the precursor
LISA Pathfinder mission, a series of propulsive requirements have been derived. Thrust levels from 5 to
30 N with precision better than 0.1N have been demonstrated with colloid and field emission
thrusters. (Ziemer et al., 2008; Scharlemann et al., 2008) Thrust noise levels less than 0.1 N/Hz1/2 are
also required, making the LISA mission one of the most stringent in terms of micropropulsion
requirements. LISA Pathfinder will fly both the LISA Test Package (LTP) and the Space Technology 7
(ST7) payloads which contain FEEP and colloid propulsion respectively. Although this is not an
exhaustive list of propulsive requirements for microspacecraft missions, it is intended to give a broad
scope and range of requirements that will need to be met by future micropropulsion systems.
8.3 Micropropulsion System Design Considerations
As previously mentioned, there is a wide ranging definition for micropropulsion which includes concepts
from scaled down versions of larger-scale thrusters to specifically designed, MEMS fabricated propulsion
systems. This and following sections will investigate some of the design considerations required to
successfully scale down existing thrusters and some of the techniques for fabricating unique MEMS
designs.
Design issues for micropropulsion systems have arisen from scaling issues, the use of corrosive
propellants with new materials, contaminant handling in propellants, limitations in MEMS fabrication
techniques, thermal cycling, and environmental interactions. In general, micropropulsion systems will
experience decreased performance over their larger-scale counterparts due to losses associated with
small characteristic sizes, limitations on system mass and power, and the lagging development of
materials and micromachining techniques for propulsion specific applications. Engineers must address
these issues by making use of novel approaches that use small-scale properties to the overall system’s
benefit. With this in mind, careful attention should be paid to the characteristics that scale favorable
with reduced size. These characteristics hold the key to the design of efficient micropropulsion systems.
8.3.1 General Design Considerations
A complication for microsystems of all types is the fact that as a typical device is scaled down, the
volume goes with the characteristic dimension, L, cubed whereas the surface area only goes with L2.
Therefore, microsystems tend to have large surface area to volume ratios relative to their larger scale
counterparts. This has consequences for heat transfer and gradients of temperature, velocity and
species. In the heat transfer case, the heat loss in the system will increase as the characteristic length
scale is reduced. For example in a 1-D heat transfer case, the energy transferred per unit area through
conduction is given by Fourier’s Law as
𝑑𝑇
𝑄̇ = −𝑘
𝑑𝑥
(1)
5
where k is the thermal conductivity of the material. For microscale systems, the value of dx will be small
while the device area to volume ratio remains relatively large. This indicates from Eq. (1) that the
amount of heat transfer in microscale devices can be relatively large. Another way to look at this
situation, is that large thermal gradients in small scale devices will not be possible unless very low
thermal conductivity materials can be identified and used. For systems (such as propulsion) that rely on
temperature gradients for efficiency, heat transfer issues will become increasingly important as the
characteristic size of the device decreases.
Heat conduction through relatively thin material will generally produce uniform system temperatures
and the thermal response time of a typical microstructure will be low. Heat conduction is complicated
by MEMS fabrication in silicon substrates with thermal conductivities that are generally higher than
many aluminum alloys. Alexeenko et al. (2005) showed in a coupled thermal-fluid simulation that a
MEMS fabricated silicon thruster reached a uniform temperature, from the combustion chamber to the
nozzle exit plane, almost instantaneously. In most cases, active cooling of the silicon microthruster was
required for survivability. In micronozzle flows, the maximum heat transfer occurs at the nozzle throat
(minimum area). The heat transfer rate at the throat varies as Re0.8/D where Re is the flow Reynolds
number and D is the nozzle throat diameter. Therefore, in micronozzle flows, the amount of heat
transfer between the gas and nozzle walls is a competition between the characteristic size, D, and the
viscous losses in the nozzle governed by the Reynolds number.
8.3.2 MEMS Fabrication Techniques
MEMS encompasses a wide range of systems that include the integration of mechanical components,
sensors, actuators, and electronics onto a common silicon substrate.(Gad-el Hak, 2002) Generally,
MEMS devices have characteristic dimensions of less than 1mm and more than 1m. MEMS fabrication
techniques include silicon micromachining through chemical etching or electro-discharge, lithography,
molding, and thin-film deposition. Micromachining selectively etches parts of a silicon wafer whereas
thin-film deposition can add new layers all in an attempt to form mechanical and electronic devices.
Today, nanoelectromechanical systems (NEMS) processes are being developed which could lead to
further breakthroughs in propulsion system design. Although many of these fabrication techniques are
currently limited, the development of these areas for terrestrial applications will continue to have a
strong impact for propulsion system engineers. An excellent review of these technologies and their
applications can be found Gad-el Hak (2002).
A MEMS fabricated micro-resistojet (Lee et al., 2008) is shown in Fig. 4. The process for fabrication of
the device is shown in Fig. 5. First, a photoresistive material is spun on a silicon wafer. The wafer is
exposed to Ultraviolet light through an intervening mask. Light exposed sections of the photoresistive
material are removed in a developing process. Deep reactive ion etching (DRIE) (Gad-el-Hak, 2002) was
used to micromachine the 100 m wide gas expansion slots (pseudo-nozzles) in areas where the
photoresistive material was removed. The DRIE process does not affect the silicon substrate in areas
where the photoresistive material remains. Once the expansion slots are machined, a secondary
photoresist process is performed to mask areas for the thin film heater deposition. The thin film heater
is composed of a 0.03 m thick layer of titanium, a 0.06 m thick layer of platinum, and a 0.075 m thick
6
layer of gold. The layers act as a bonding agent with the silicon, a diffusion barrier, and the current
carrying layer respectively. Other examples of MEMS fabricated thrusters can be found in Mueller
(2000).
Figure 4: MEMS fabricated heater chip for a micro-resistojet concept.
Mask
Ink
Ultraviolet
Developer removes
exposed photoresist
light
Photoresist
exposed photoresist
Silicon wafer
Exposed
photoresist
Metallization
Unexposed
photoresist
Acetone removes
unexposed photoresist
photoresist
Expansion Slots Etched with DRIE
(Deep Reactive Ion Etcher)
Open Inlet
Vacuum Boundary
- 300 Å Ti (bottom), 600 Å Pt (middle), 750 Å Au (top layer)
Figure 5: MEMS fabrication techniques employed in the fabrication of the micro-resistojet.
8.3.3 System Derived Specific Impulse
In an attempt to quantify the fundamental viability of a micropropulsion concept for a particular
microspacecraft mission, the concept of an effective or system comparable specific impulse has been
developed. The effective (or system) specific impulse modifies the thruster’s intrinsic specific impulse
by taking into account system penalties such as:



High-pressure propellant storage and handling
Micro-valve leakage
Unusable propellant at EOL of a thruster
7

Massive power supplies
Although the idea of a system specific impulse is useful for all types of propulsion systems, it is
particularly useful for micropropulsion systems where systems-level constraints are paramount.
The effective specific impulse of a thruster system in terms of the excess mass associated with high
pressure propellant storage, micro-valve leakage, and left over propellant at EOL can be written as
(Ketsdever, Wadsworth and Muntz, 2000)
𝐼𝑠𝑝𝑒𝑓𝑓 =
𝑝
𝐼𝑠𝑝[1−( 𝑜,𝑑𝑒𝑠⁄𝑝𝑖,𝑠𝑡𝑜𝑟 )]
(1+𝛿)[1+(𝑀𝑡⁄𝑀 )]
𝑝,𝑠
(2)
where Isp is the intrinsic specific impulse, po,des is the design stagnation pressure, pi,stor is the initial
propellant tank storage pressure, is the fraction of propellant lost to valve leakage, Mt is the propellant
tank mass, and Mp,s is the mass of the stored propellant. Similar formulations that take into account
electrical subsystem (power processing units) mass, propellant feed system mass, and other propulsion
system dry mass can be envisioned. In this way, propulsion devices that may have relatively low intrinsic
performance can find niches where their combined benefits to the overall microspacecraft system
outweigh any single drawback.
8.4 Chemical Combustion at Microscales
Liquid and solid propellant thrusters are specifically targeted at microspacecraft missions where high
thrust is required or where minimal power might be available. Generally, missions requiring high thrust
also require spacecraft rapid response. As is the case for all space chemical propulsion systems,
propellants stored in either a liquid or solid state have several systems advantages over the use of
gaseous propellants stored at high pressure. Typical liquid propellants (hydrazines, nitrogen tetroxide,
methyl amine, chlorine trifluoride) are volatile and toxic which leads to material compatibility issues
with MEMS materials like silicon. Combustion temperatures are also quite high leading to heat transfer
issues discussed in previous sections.
Because of the large heat transfer rates expected in small scale devices, lower combustion chamber
temperatures are expected in micropropulsion systems utilizing combustion where the heat loss will be
proportional to L3/2. (Yetter et al., 2007) Lower combustion chamber temperatures will result in low
specific impulses. At smaller scales, velocity and temperature gradients near the combustion chamber
walls will tend to increase. This will produce larger mass, momentum and energy transport rates near
the wall making it difficult to maintain large differences between wall and bulk flow properties critical
for efficient operation.
The mixing length of propellants (fuel and oxidizer) is often used as a measure of the combustion
efficiency. Every chemical reaction process requires a finite time on the order of a few mean molecular
collision times of gas phase species in the combustion chamber. Therefore the combustion chamber
must have a minimum length of a few molecular mean free paths at the combustion chamber pressure
8
to be effective. In this scenario, length can be reduced at the cost of higher combustion chamber
pressure or shorter chemical timescales. The characteristic length of a combustion chamber is given by
𝑉
𝐿∗ = 𝐴𝑐
𝑡
(3)
where Vc is the combustion chamber volume, and At is the nozzle throat area. The characteristic length
of a chemical propulsion system’s combustion chamber depends on several factors including the
propellant type, injector type and configuration, chamber geometry, and whether a catalyst is used. As
one example, evaporation time for a droplet from a spray will need to be reduced, which leads to
equivalent reductions in droplet size in order to obtain efficient combustion. For micropropulsion
systems, the design criteria is for L* to be as small as possible while maintaining acceptable
performance.
Monopropellants appear attractive for micropropulsion applications since the mixing of a fuel and
oxidizer is not required. (Hitt, Zakrzwski and Thomas, 2001; Wu et al., 2006) However, long-term
survivability of catalyst material at elavated temperatures will be an issue for these systems. A variety
of new propellants are being considered to produce efficient combustion. (Yetter et al., 2007)
8.5 Micronozzle Flows
Many micropropulsion systems will require the expansion of propellant gases through micronozzle
geometries. As with large scale propulsion systems, nozzles convert thermal energy in a propellant into
kinetic energy. For micropropulsion systems, the expansion of a propellant gas through a small scale
nozzle poses several challenges. The low thrust levels needed for many microspacecraft missions will
require either low pressure operation of a thruster or a decrease in nozzle throat dimensions. Both of
these scenarios lead to low Reynolds number flows. Additionally, high temperature gas expansion from
micro-chemical or micro-electrothermal thrusters will also continue the trend of lowering the operating
Reynolds number of nozzle flows. The Reynolds number is a measure of nozzle efficiency in terms of the
viscous losses inherent in the subsonic boundary layers near the nozzle surfaces. At low Reynolds
number, the viscous losses in a micronozzle become large, degrading performance. At sufficiently low
Reynolds number, the viscous boundary layer (i.e. the subsonic flow region) can encompass a significant
fraction of the micronozzle geometry.
A typical two-dimensional, MEMS fabricated micronozzle is shown schematically in Fig. 5. These 2-D
nozzles have square or rectangular throat and exit plane sections which leads to inefficiencies due to
sidewall boundary layer formation; however, this micronozzle geometry is ideal for MEMS fabrication
techniques. Results from experiments with 34 micron throat diameter, 2-D nozzles indicated that thrust
efficiencies from 0.1 to approximately 0.8 could be expected in the Reynolds number range of 500-3500
for a nozzle expansion ratio of 7.1:1. A larger expansion ratio (16.9:1) micronozzle showed improved
performance for the same range of operating Reynolds number. A recent study by Louisos and Hitt
(2009) showed the numerical performance of a traditionally MEMS fabricated 2-D micronozzle geometry
computed in a 3-D domain. Figure 5 shows numerical results for a simulated 2-D MEMS fabricated
micronozzle.
9
Figure 5: An illustration of the subsonic layer as denoted by Mach contours less than unity for selected
cross sections in a 200μm deep nozzle operating at Re=60 (top) and Re =30 (bottom).
Figure 6 shows Mach number contours for a helium flow through a 3-D, conical micronozzle geometry at
two different Reynolds numbers. These results were obtained using the Direct Simulation Monte Carlo
(DSMC) numerical method. Note that the subsonic boundary layers have grown significantly at the low
Reynolds number in both Fig. 5 and Fig. 6 indicating degradation of micronozzle performance at the
lower pressure operation. Figure 7 shows the intrinsic specific impulse of the micronozzle flow at a
Reynolds number of 60 as a function of axial position through the micronozzle (X=0 corresponds to the
micronozzle throat). Figure 7 indicates that the intrinsic specific impulse decrease from a maximum just
past the throat to a minimum at the exit plane regardless of the micronozzle length. In the case of the
full length nozzle, the specific impulse at the micronozzle exit plane is actually lower than the specific
impulse at the micronozzle throat indicating that the nozzle at this Reynolds number is actually
degrading performance. In fact, Fig. 8 shows that at Reynolds numbers below approximately 100, a
thin-walled, sonic orifice tends to outperform a typical micronozzle geometry.
10
Figure 6: Predicted Mach number contours for helium flow inside a micronozzle geometry at two
different throat Reynolds numbers. Subsonic, viscous boundary layer is indicated by series of contours
with M<1.
Figure 7: Impact of micronozzle length on performance for a helium flow with a throat Reynolds number
of 60. The full length nozzle is 1.07x10-2 m in length.
11
1.1
Fn / Fo
1
0.9
0.8
0.7
0.6
0
50
100
150
Re*
Figure 8: Ratio of nozzle to orifice thrust for helium propellant as a function of throat Reynolds number.
8.6 Plasma Formation at Microscales
Because of the large surface area to volume ratios in micropropulsion systems, electron losses to
chamber walls will be the limiting factor in the efficiency of microscale plasma devices. In large scale ion
and Hall-effect thrusters, magnetic fields are used to contain the electron motion and prevent them
from being lost to the discharge chamber walls. The power lost to the walls can be estimated by the
ratio of the current to the wall to ionization potential of the propellant. In this case, the power lost at
the walls scales directly with the characteristic scale of the thruster, L. These magnetic fields also
increase the electron path length in the discharge chamber, increasing the probability of ionization. As
these types of systems scale down, proportionately larger magnetic fields are required to provide
efficient ionization. Essentially, the magnitude of the magnetic field required scales as
𝐵=
𝑚𝑒 𝑣𝑝
𝑞𝑅𝑔
(4)
where me is the electron mass, vp is the electron velocity perpendicular to the magnetic field, q is the
elemental charge, and Rg is the electron radius of gyration. To effectively contain the electrons, Rg has
to be a reasonably small fraction of the discharge chamber radius. As the scale size of the discharge
chamber decreases, the required magnetic field increases. For example, a 1mm diameter ion engine
would require on the order of a 10 Tesla magnetic field for electron containment. High power solenoids
or heavy permanent magnets would be required, clearly indicating that these types of thrusters will
encounter serious issues at small scales.
8.7 Micropropulsion Systems: Current State of the Art
A complete review of micropropulsion systems cannot be performed within the context of this section.
Instead, this section is intended to give an update on system development since the last comprehensive
review of Mueller (2000).
8.7.1 Microchemical thruster development
12
Louisos and Hitt (2009) discussed the general performance of a high purity, hydrogen peroxide (H2O2)
monopropellant microthruster. One benefit of hydrogen peroxide is that the products of the
combustion are non-toxic. Decomposition of the monopropellant occurs in a catalyst bed producing a
flame temperature of 886 K with 85% pure H2O2. Simulations for stagnation pressures from 5 kPa to 250
kPa yielded thrusts in the range of 1-20 mN after gas expansion through a micronozzle. Yetter et al.
(2007) described the development of a liquid monopropellant thruster that uses electrolytic ignition.
This thruster was fabricated from alumina-based ceramic tape with maximum operating temperatures
of 1200 K. Thrust levels of approximately 100 to 200 mN were measured using both XM46 and RK315A
propellants.
8.7.2 Micro-Pulsed Plasma Thrusters (PPT)
Pulsed plasma thrusters use an arc struck across a non-conducting material to ablate propellant and
generate thrust. [See 2.4.06 Pulsed Plasma Thrusters] The PPT uses a surface discharge across a solid
Teflon™ propellant to provide impulses in the range of 10 N-sec. (Spanjers et al., 2002) In recent
designs, a high voltage (~3 kV) capacitor is directly coupled with a coaxial propellant module. The
module consists of three electrodes each surrounded by an annulus of propellant as shown in Fig. 9.
The PPT is operated by initiating breakdown between the innermost electrodes which produces
enough plasma to initiate breakdown between the outermost electrodes. Pulses on the order of several
Joules of energy have been produced at repetition rates of approximately 1 Hz. Thrust-to-power ratios
of approximately 10 N/W have been demonstrated. Initial lifetime tests have been conducted that
have run the thruster to the point where all of the propellant has been consumed. A benefit of the PPT
scaling from its larger counterparts has been the ability to remove the initiating spark plug and igniter
circuitry. The PPT has been flown on FalconSat 3, a microspacecraft designed and built by USAFA. Two
of the PPT thruster modules are visible at the top of FalconSat 3 in Fig. 2 (large cubes with copper
rods).
Figure 9: Three-electrode PPT schematic (after Spanjers et al., 2002).
13
8.7.3 Vacuum Arc Thruster (VAT)
In the VAT design, a self-triggered vacuum arc is utilized to form a fully ionized metal plasma from a cold
cathode. (Schein et al., 2002) Plasma velocities upwards of 30 km/s have been observed for aluminum
cathodes with a corresponding specific impulse of about 3000 sec. The specific impulse varies with
cathode material. Discharges with arc currents between 10A and several kA have been produced with
pulse lengths on the order of several microseconds. One benefit of the VAT over a typical PPT is its
ability to operate at much lower voltages (<12 V). The VAT is operated in a pulsed mode up to 100 Hz
with typical average power levels from 1 to 100 W. Initial results have shown that the VAT can obtain
individual impulses on the order of 1 N-sec with thrust-to-power ratios of approximately 20 N/W.
The overall VAT mass, including the power processing unit (PPU), is about 0.1 kg with an overall
efficiency estimated to be about 13%.
8.7.4 Field Emission Electric Propulsion (FEEP)
The FEEP concept uses electrostatic acceleration of metal ions formed through field ionization of liquid
metal propellants, typically cesium or indium, as shown in Fig. 10. Both cesium and indium FEEP
thrusters are being developed for the LISA Pathfinder mission (as part of the LTP payload) which could
launch as early as 2010. Single FEEP modules can operate at continuous thrust levels from 0.1 to 15 N
with peaks near 35 N. They are capable of very small impulse bits (<5 nN-sec) with a very large specific
impulse (4000-8000 sec). The great benefit of FEEP thrusters for extremely accurate missions like LISA is
their low thrust noise values and high thrust resolution and repeatability. Thrust-to-power ratios of
nearly 15 N/W have been produced with relatively high electrical efficiencies. (Scharlemann et al.,
2008) Tajmar (2004) has investigated scaling FEEP thrusters using MEMS and other micro-fabrication
techniques to produce large arrays of emitters. Because indium does not wet silicon, traditional MEMS
techniques proved unsuitable. Laser machining was incorporated to produce an array of 441 emitters in
a 5x5 mm area. Although the thrusters were able to be fired, currents of only 5A were able to be
produced. Future design optimization could significantly improve the micro-FEEP performance.
(a)
(b)
14
Figure 10: FEEP thruster a) operating principle and b) operation of an 8cm-diameter module on indium
(Photo courtesy of ARC, Austria)
8.7.5 Colloid thruster development
Colloid propulsion utilizes the charging of small liquid drops produced in Taylor cones which are
accelerated through a high electrostatic potential [See 2.4.07 Colloid Thrusters] Colloid propulsion is
currently being developed for use on the ST7 payload for the LISA Pathfinder mission. Thrust in the
range of 5 to 30 N with specific impulse greater than 150 sec has been demonstrated. (Ziemer et al.,
2008) As with FEEP thrusters, high thrust precision and low thrust noise also makes colloid thrusters
attractive for high precision maneuvers on microspacecraft. The LISA mission architecture and flight
unit for the ST7 payload are shown in Fig. 11. Arrays of micro-fabricated colloid emitters have been
developed by Valasquez-Garcia, Martinez-Sanchez and Akinwande (2004). These colloid thrusters can
have between 4 and 1025 emitters in a 0.64 cm2 area. Recent advances have also been made in the
development of propellants for colloid thrusters where ionic liquids are being investigated.
(a)
(b)
Figure 11: a) LISA mission architecture, b) Colloid thruster module for the ST7 payload. (Photo courtesy
of the Jet Propulsion Laboratory)
8.7.6 Small-scale Hall thruster development
Low power Hall thrusters [See 2.4.03 Hall Thrusters] have been designed and tested. Recent
developments in small scale Hall thruster development has been significant. (Diamant et al., 2008;
Hargus and Charles, 2008; Ito et al., 2006; Biagioni et al., 2005) Diamant et al. (2008) investigated a low
power cylindrical Hall thruster over a range of input powers from 70 to 220 W. Over this range, they
observed anode efficiencies on the order of 15 to 35% indicating the issues that Hall thrusters will have
15
for low power microspacecraft mission applications. Hargus and Charles (2008) investigated a 200Wclass Hall thruster system operating with a 250 V anode potential. They observed maximum ion
velocities corresponding to 192 eV with xenon propellant which yields an approximate specific impulse
of 1700 sec. A system similar to the one tested by Hargus and Charles (2008) will fly on FalconSat 5, a
follow-on microspacecraft designed and built by USAFA. A small Hall effect thruster is described by
Bianioni et al. (2005). Stable operation was observed in the power range between 10 and 235 W with
anode efficiencies in the range of 15 to 32%. This system utilizes permanent magnets to contain the
electrons formed in the discharge. This study attempted to address issues in small scale Hall thruster
development in which limited operational efficiency was caused by electron losses to discharge chamber
walls and high heat transfer rates.
Ito et al. (2006) designed a much smaller Hall thruster system with a 0.5mm channel width and a 4mm
outer diameter. A permanent magnet with a field strength of 1 T in the ionization channel was used.
Average powers of 10-40 W were demonstrated with efficiencies in the range of 10-15%. The thrust
ranged from 0.6 to 1.6 mN with specific impulses in the range of 300-850 sec. All of these small scale,
low power Hall thrusters show the limitations on efficiency as the physical size and operational power
levels are reduced. These limitations are obviously not unique nor restricted to Hall thrusters.
Acknowledgements
The DSMC micronozzle calculations in Figs. 6 and 7 were performed by Dr. Sergey Gimelshein of the
University of Southern California. The micronozzle calculation in Fig. 5 was performed by Will Louisos
and Prof. Darren Hitt of the University of Vermont. The photo in Fig. 2 is courtesy of Sgt. Michael
Wickersheim of the United States Air Force Academy. The photo in Fig. 10(b) is courtesy of Dr. Martin
Tajmar of ARC, Austria. Figure 11(a) is courtesy of NASA/JPL. The photo in Fig. 11(b) is courtesy of Dr.
John Ziemer of the Jet Propulsion Laboratory. The author wishes to thank Dr. David Scharfe and Mr.
Shawn Laabs for their help in preparing this article.
References
Alexeenko AA, Levin DA, Fedosov DA, Gimelshein SF and Collins RJ. Performance analysis of
microthrusters based on coupled thermal-fluid modeling and simulation. J. Prop. and Power. 2005; 21:
95-101.
Biagioni L, Cesari U, Saverdi M and Andrenucci M. Development status of the HT-100 miniaturized Hall
effect thruster system. AIAA paper 2005-3875. Joint Prop. Conf., Tucson, AZ. July 2005.
Diamant KD, Pollard JE, Raitses Y and Fisch NJ. Low power cylindrical Hall thruster performance and
plume properties. AIAA paper 2008-4998. Joint Prop. Conf., Hartford, CT. July 2008.
Gad-el-Hak M. Introducion. In The MEMS Handbook, Gad-el-Hak M (ed). CRC Press, 2002; 1-5.
Hargus WA and Charles CS. Near exit plane velocity field of a 200W Hall thruster. J. Prop. and Power.
2008; 24: 127-133.
16
Ito T, Gascon N, Crawford W and Cappelli M. Further development of a micro Hall thruster. AIAA paper
2006-4495. Joint Prop. Conf., Sacramento, CA. July 2006.
Janson SW, Helvajian H, Hansen WW and Lodmell J. Microthrusters for nanosatellites. Second Intl. Conf.
on Integrated Nanotech. for Space Apps. (MNT99). Pasadena, CA. April 1999.
Ketsdever AD. System considerations and design options for microspacecraft propulsion systems. In
Micropropulsion for Small Spacecraft, Micci M. and Ketsdever A. (eds). AIAA, 2000; 187: 139-163.
Lee RH, Bauer AM, Killingsworth MD, Lilly TC, Duncan JA and Ketsdever AD. Free-moleculemicroresistojet performance using water propellant for nanosatellite applications. J. Spacecraft and
Rockets. 2008; 45: 264-269.
Louisos WF and Hitt DL. Numerical simulations of viscous flow in 3D supersonic bell micronozzles. AIAA
paper 2009-446. Aero. Sci. Meeting, Orlando, FL. January 2009.
Mueller J. Thruster options for microspacecraft: a review and evaluation of state-of-the-art and
emerging technologies. In Micropropulsion for Small Spacecraft, Micci M. and Ketsdever A. (eds). AIAA,
2000; 187: 45-137.
Scharlemann C, Genovese A, Buldrini N, Schnitzer R and Tajmar M. In-FEEP qualification test program for
LISA Pathfinder. AIAA paper 2008-4825. Joint Prop. Conf., Hartford, CT. July 2008.
Schein J, Qi N, Binder R, Krishnan M, Ziemer J, Polk, J and Anders A. Inductive energy storage driven
vacuum arc thruster. Rev. Sci. Instrum. 2002; 73: 925-927.
Schilling JH, Spores RA and Spanjers GG. Micropropulsion options for the TechSat21 space-based radar
flight. In Micropropulsion for Small Spacecraft, Micci M. and Ketsdever A. (eds). AIAA, 2000; 187: 3-23.
Spanjers G, Bromaghim D, Lake J, Dulligan M, White D, Schilling J, Bushman S, Antonsen E, Burton R,
Keidar M and Boyd I. AFRL MicroPPT development for small spacecraft propulsion. AIAA paper 20023974. Joint Prop. Conf., Indianapolis, IN. July 2002.
Sweeting MN. Micro/nanosatellites – the new world. In Smaller Satellites: Bigger Business?: Concepts,
Applications and Markets for Micro/Nanosatellites in a New Information World, Rycroft M and Crosby N
(eds). Springer, 2002; 1-20.
Tajmar M. MEMS indium FEEP thruster: manufacturing study and first prototype results. AIAA paper
2004-3619. Joint Prop. Conf., Fort Lauderdale, FL. July 2004.
Yetter RA, Yang V, Wu MH, Wang Y, Milius D, Aksay I and Dryer F. Combustion issues and approaches for
chemical microthrusters. In Advancements in Energetic Materials and Chemical Propulsion, Kuo K. and
Rivera J (eds). Begell House, 2007; 389-421.
17
Ziemer JK, Randolph TM, Franklin GW, Hruby V, Spence D, Demmons N, Roy T, Ehrbar E and Zwahlen J.
Delivery of colloid micro-Newton thrusters for the Space Technology 7 mission. AIAA paper 2008-4826.
Joint Prop. Conf., Hartford, CT. July 2008.
18
Download