You are on page 1of 10

42nd AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit AIAA 2006-4551

9 - 12 July 2006, Sacramento, California

AIAA-2006-4551

Test of a Turbo-Pump Fed Miniature Rocket Engine

C. Scharlemann1, M. Schiebl 2, K. Marhold 3, M. Tajmar4,


Space Propulsion, ARC Seibersdorf research, A-2444 Seibersdorf, Austria

P.Miotti 5,
Mechatronic GmbH, Tirolerstr.80, A-9500 Villach, Austria

C.Guraya 6, F.Seco 7
INASMET, E-20009 San Sebastián, Spain

C.Kappenstein 8, Y. Batonneau 9, R. Brahmi 10,


LACCO, University of Poitiers, F-86022 Poitiers, France

M. Lang 11
ESTEC-ESA, Keplerlaan 1, Noordwijk, The Netherlands

The increasing application of microsatellites (from 10 kg up to 100 kg) for a rising number of
various missions requires the development of suitable propulsion systems. Microsatellites have special
requirements for a propulsion system such as small mass, reduced volume, and very stringent
electrical power constraints. Existing propulsion systems often can not satisfy these requirements. The
present paper discusses the development and test of a bipropellant thruster complying with these
requirements.
The main development goal of this effort was the utilization of ethanol in combination with
hydrogen peroxide (H2O2) as a non-toxic propellant combination. The Turbo-Pump Fed Miniature
Rocket Engine (TPF-MRE) is a bipropellant thruster consisting of four subsystems: the propellant
pumps, a decomposition chamber with a monolithic catalyst, a turbine, and the thruster itself. The
turbine is driven by the decomposed hydrogen peroxide and magnetically coupled with a power
generator. The power produced is then used to generate a pressure head in order to deliver the
propellant into the combustion chamber. This system therefore constitutes a self-sustaining system and
does not rely on the limited power supply of a micro-satellite.

Previous test have shown that although the thruster can be operated with ethanol and oxygen, it
was not possible to ignite the thruster when utilizing hydrogen peroxide in a 70% concentration by
weight. A minor redesign of the thruster and the test facility was therefore initiated. This redesign
together with the use of hydrogen peroxide in higher concentration was speculated to improve this
behavior. However, even though the monolithic catalysts were able to decompose hydrogen peroxide in
a concentration of 87.5 % with nearly 100 % efficiency, it was not possible to ignite or operate the
thruster. Subsequently, a thorough investigation of the baseline design and operational conditions of
the thruster was conduced.
It was found that the failure of the thruster to ignite is due to a combination of reasons. The
combustion chamber length is too short to facilitate sufficient mixing of the propellants, making an
ignition impossible or very difficult at least. Additionally, the combustion chamber pressure which was
chosen such that it accommodates the performance of commercially available mircopumps is
considered too low. This further deteriorates the conditions for which an ignition is feasible.

I. INTRODUCTION
icrospacecrafts and –satellites with their distributed functionality offer a fail-safe degree unreachable by
M the former one-satellite mission strategy. By using multiple microsatellites for one mission, the failure of
one or even several satellites may still allow a full or at least partial successful completion of the mission
objective. However, microsatellites have in general more severe constrains in regard to the available electric

1
Senior Scientist, Space Propulsion, Email: carsten.scharlemann@arcs.ac.at, AIAA member
2
Research Assistant, Email: markus.schiebl@arcs.ac.at
3
Research Assistant, Email: klaus.marhold@arcs.ac.at
4
Head of Space Propulsion, Email: martin.tajmar@arcs.ac.at, AIAA member
5
Ph.D. student, Email: pierpaolo.miotti@mechatronic.at
6
Senior Scientist: Email: cristina.guraya@inasmet.es
7
Project director, Email: fernando.seco@inasmet.es
8
Professor, Email: charles.kappenstein@univ-poitiers.fr, AIAA member
9
Assistant Professor, Email: yann.batonneau@univ-poitiers.fr AIAA member
10
Assistant Professor, Email: rbrahmi@ext.univ-poitiers.fr
11
Senior Propulsion Engineer (ESTEC), Email: Martin.Lang@esa.int
1
American Institute of Aeronautics and Astronautics

Copyright © 2006 by C. Scharlemann. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
AIAA-2006-4551

power and the maximum allowable mass and volume of the propulsion system than their larger counterparts.
These constrains pose several challenging problems for the propulsion system design since in many cases a
simple downsizing of a larger system is not an option. In particular the small allowable total mass of the
propulsion system requires special care. Although this is subject of a detailed case-to-case study one can in
general exclude propellants which need cryogenic storage systems. If possible, one should also avoid the use of
pressurized propellant tanks in order to keep system masses low. The use of non-toxic, or green propellants,
offer increased safety and reduced costs.

This is even more important in the case of multiple microsatellites compared to one large satellite since the
possibility of a propellant leak increases linearly with the number of satellites. Using toxic or carcinogenic
propellants for microsatellites would therefore increase safety concern and the costs to deal with them would
rise disproportionately.

To satisfy the propulsion needs for such microsatellites the development of a bipropellant thruster system was
initiated. The bipropellant thruster presented here was developed in a multi-national effort. The development of
the thruster concept and design was performed by Mechatronic, Austria. The design process was supported by
numerical calculations performed by the University of Udine, Italy. The Spanish company INASMET
manufactured the thruster components according to the specification of Mechatronic and the catalysts were
provided by the University of Poitiers, France. Finally, the tests of the thruster and its subcomponents were
performed in the test facility of ARC-sr, Austria1,2.

The Turbo-Pump Fed Micro Rocket Engine (TPF-MRE) consists of four main components: the propellant
pumps, a decomposition chamber with a monolithic catalyst, a turbine, and the thruster itself. The pumps are
generating the pressure head to feed the fuel and oxidizer into the combustion chamber. While the fuel is fed
directly into the combustion chamber, the hydrogen peroxide is led over a catalyst bed (see fig. 1). In a
combination of catalytic and thermal decomposition, hydrogen peroxide is decomposed into water and oxygen.
Since this is a highly exothermic reaction, the mixture will leave the catalyst in gaseous form and is fed directly
into a microturbine. Using the energetic mixture, the microturbine generates the electrical power necessary to
drive the two pumps. After exiting the turbine, the water/oxygen mixture is fed via cooling channels along the
nozzle, throat, and combustion chamber walls into the combustion chamber. The envisioned performance of the
TPF-MRE is shown in table 1.

Fuel Pump Oxidizer


Pump

M M

G.G.

Battery G

Fig. 1: Schematic of the TPF-MRE system concept (M for micropumps, G for electric generator, GG for gas
generator (catalyst)).

This concept satisfies all the requirements for a propulsion system as outlined above. The use of hydrogen
peroxide satisfies the need for green propellants and allows keeping the system costs down. The use of pumps
instead of pressurized tanks will help to keep the total system mass low. The combination of the pumps and the
turbine constitutes a self sustaining system (with the exception of the start up phase) and therefore does not rely
on the reduced power availability on microsatellites.

The main development goal of this effort was the utilization of ethanol in combination with hydrogen
peroxide (H2O2) in a concentration of 70% by weight as a non-toxic, non carcinogenic propellant combination.

2
American Institute of Aeronautics and Astronautics
AIAA-2006-4551

Hydrogen peroxide is a well-known propellant for propulsion purposes. Various levels of hydrogen peroxide
concentrations are in use, ranging from 80 to 96% by weight3-6.

Table 1: Envisioned performance of the TPF-MRE


Thrust N 1
Spe cific Im pulse s > 230
De lta -V m .s -1 500
M a x im um m a ss gram s 400
M a x im um volum e litre 0.8
Sa te llite cla sse s kg 10 - 100
Ele ctrica l pow e r W 6

Most prominent fuel utilized in combination with hydrogen peroxide is kerosene7-9. However, ethanol was
chosen for this project due to its favorable thermochemistry, its “green propellant” characteristic, and since it
does not soot when exposed to a hot environment. The latter is particularly important with regard to the very
small dimensions of the injector.

In a first test series, an initial evaluation of the various components (thruster, turbine, and pumps) has been
conducted and the results have been reported in the past2. With respect to the thruster, the results were not
satisfactory. Although it was possible to ignite the thruster and achieve stable combustion with ethanol and
oxygen, the ignition did not occur when the oxygen was replaced with hydrogen peroxide in 70% concentration.
Several possible reasons for the failure to ignite and combust were identified and subsequently a re-design of
the thruster was initiated as discussed in the following.

II. REDESIGN OF THE TPF-MRE SYSTEM


A. THRUSTER
A-1. Propellants
The thruster was designed for ethanol and hydrogen peroxide in a 70 % concentration by weight. When initial
tests of the thruster with this propellant combination did not lead to a successful ignition and combustion, it was
necessary to consider the particular propellant choice as one of the reasons for the failure. Initiating again a
literature research, it was found that the combination of 70 % hydrogen peroxide and ethanol is rather exotic and
no other thruster is known working with this combination. In particular, the choice of 70 % hydrogen peroxide
results in decomposition temperatures of maximum 230°C which is more than 100°C lower than the auto-
ignition temperature of ethanol in dry air. Furthermore, the particular design of the hydrogen peroxide feeding
system facilitated relative large conductive heat losses leading to even lower temperature.
Beside of the necessary design changes with regard to the thruster (discussed in the following section), it was
therefore decided to test hydrogen peroxide in higher concentration than 70%. This should not only ease the
ignition difficulties but has also the potential to heat up the thruster structure enough to vaporize the ethanol
before injection.

Ethanol was originally chosen due to its favorable thermochemistry, its “green propellant” characteristic, and
since it does not soot when exposed to a hot environment. The latter is particular important with regard to the
very small dimensions of the injector. Furthermore, the combination of hydrogen peroxide and ethanol was
identified as one of the best choices especially in regard to the search for green propellants10.

In addition to higher H2O2 concentrations, a search for alternative fuels was initiated. Candidates included
DMAZ (Dimethylamine-2-Ethylazide), kerosene, and turpentine. Changing the propellant at this point of the
project is a dramatic change since it changes the combustion chemistry and with it the flame temperature, heat
of reaction etc. This again has significant impact on the thruster design such as the cooling channels, nozzle
design etc. However, from former tests with ethanol and oxygen and without any cooling of the thruster2 it was
known that the thruster has a certain tolerance to off-design operation and would therefore allow at least short
duration tests even if the combustion temperature exceeds the design level.

The understanding was that it is most important to lower the self-ignition temperature of the H2O2/fuel
mixture. With regard to this, turpentine offers the best conditions according to findings of Musker et al.11.
Turpentine is a mixture of bicyclic monoturpenic hydrocarbons with a high beta-pinene content. It is produced
from pine trees and is non-toxic and non-carcinogenic. Turpentine is a colorless, transparent, oily liquid with a
penetrating odor. Its properties such as its viscosity etc. are close enough to ethanol to allow the use of the same
test facility as utilized for ethanol.

3
American Institute of Aeronautics and Astronautics
AIAA-2006-4551

In conclusion, it was decided to test two propellant combinations: (i) ethanol and H2O2 in 87.5 %
concentration and (ii) turpentine and H2O2 in 87.5 % concentration.

A-2. Thruster design


It was important to accomplish the redesign without requiring changes of too many parts due to the limited
time and funding still available in this project. It was therefore not possible to include major design changes.
The realized design changes included an improved sealing of the thruster and the replacement of the glow plug.
The original thruster design used a glow plug to vaporize the ethanol prior to injection. Injection of vaporized
ethanol would benefit the mixing and combustion process. However, the necessary input power to vaporize a
mass flow of about 0.05 g/s of ethanol is roughly 100 W. The maximum input power of the glow plug turned out
to be about 8 W. Exceeding this power level lead to the destruction of the glow plug. Better suitable alternative
glow plugs, however, would be larger than the thruster itself and the power input, necessary for the thruster
initiation would possibly exceed the battery performance. It was therefore decided to abandon the use of a glow
plug at this point. It was clear that this would mean the injection of fluid ethanol into the combustion chamber
and subsequently a more difficult ignition process. Figure 2 shows the cross section of the old (left side) and the
new design (right side). The change in the sealing concept at the nozzle can be seen. This change allowed the
sealing of the thruster without the need of welding which has been proven to cause leakage (mainly due to the
difficulties to weld Nimonic 105). As discussed above, the new design on the right of fig. 2 does not incorporate
a glow plug anymore.

Fig. 2: cross section of the old (left) and the new (right) thruster designs. The red arrow depicts the fuel injection path

The relative long hydrogen peroxide feeding line from the decomposition chamber to the thruster allowed the
decomposed gases to cool down excessively. Passing through the feeding lines, the hot gases are heating up not
only the feeding lines but also a very large fixing plate (stainless steel) which accommodates the pressure
gauges (see left side of fig. 3). This constitutes a rather large thermal drain for the decomposed hydrogen
peroxide, resulting in gas temperatures even lower then the 230°C which are possible with hydrogen peroxide in
a concentration of 70%.
Subsequently, the feeding connections were changed with shorter ones and the fixing plate was removed (see
right side of fig. 3).

B. CATALYST AND DECOMPOSTION CHAMBER


The use of hydrogen peroxide in a concentration of 87.5 % by weight increases the theoretically possible
decomposition temperature from 230°C up to 670°C. The originally utilized decomposition chamber was
manufactured from aluminum and needed to be replaced in order to endure the increased temperature.
Furthermore, in the interest of the highest possible decomposition temperatures, the re-design focused on a
reduction of the thermal losses and subsequently a decrease of the transition times.
The new decomposition chamber is manufactured from stainless steel (X8CrNiS 18-9). General focus of the
re-design was a reduction of the thermal masses of the decomposition chamber and the generation of a thermal
path between inlet and outlet of the decomposition chamber. The latter allows a certain heat exchange between
the decomposition chamber structure and the liquid hydrogen peroxide in order to increase its temperature prior

4
American Institute of Aeronautics and Astronautics
AIAA-2006-4551

to injection into the catalyst. This was found to increase the decomposition efficiency and decreases the
transition time7. The propellant manifold was designed such that it ensures a homogeneous mass flow through
all the injector holes. The volume of the propellant manifold generates a certain residence time of the propellant
to support the increase in temperature through a heat exchange with the structure of the decomposition chamber.

Decomposition
chamber

Decomposition
Fixing chamber
plate
Thruster

Thruster

Fig. 3: Old thruster test assembly (left) with the stainless steel fixing plate and the old version of the decomposition
chamber in comparison with the new test assembly

The new design has a smaller volume and even though stainless steel has a higher material density than
aluminum, the weight of the new decomposition chamber is only 60 g compared to the former aluminum one
with 120 g.

Figure 4 shows a cross-sectional view of the new decomposition chamber. The position of the thermocouple
which measures the decomposition temperature is indicated in fig. 4 by the red dot. The color coding in fig. 4 is
for better visualization of the various parts and does not refer to any material selection (all parts are made of
stainless steel). Figure 3 (right side) shows the new decomposition chamber assembled together with the
thruster.
Gas filled, metalic
o-rings

Propellant manifold
Injector plate with 49
injector holes

Fig. 4: Cross sectional view of the decomposition chamber (catalyst and thermocouple not shown)

The geometry of the utilized monolithic catalyst is identical with the one reported recently2 and the
manufacturing procedures are similar to the ones described in past publications12. However, the washcoating
procedure and the catalyst itself had to be adapted for the higher H2O2 concentration and temperature
respectively. The monoliths are wash-coated and impregnated with commercial sodium permanganate solution
or permanganic acid HMnO4, by LACCO, France. In general, the monoliths are immersed in the washcoat or
impregnation solution for several hours, under mechanical stirring. The solvent is released in a sand bath heated
to 50 °C until no water is left. Then, the samples are thermally treated under atmospheric conditions. To
eliminate the sodium cation from the surface, the samples were washed in water under mechanical stirring and
5
American Institute of Aeronautics and Astronautics
AIAA-2006-4551

afterwards dried for several hours. A second impregnation under identical conditions was applied to improve
the catalytic behavior. Finally, the samples are heated under air at 750 °C. To evaluate the quality of each step,
the weight of the samples was measured before and after each step.

III. TEST RESULTS


The test facility consists of three major subsystems: the pressurization system, the control system, and the
diagnostic system (see fig. 5). The hydrogen peroxide and the fuel reservoirs are pressurized with nitrogen. The
level of pressurization is regulated with a standard manual pressure regulator ball valves. In general, the test
facility is identical to the one recently presented in detail2. The cut-off valve and the drain valve are added to
flush the system in order to remove gas bubbles during start-up and to clean out the hydrogen peroxide lines if
necessary. The control system consists of two mass flow controllers. The hydrogen peroxide mass flow
controller is a so called CORI-FLOW meter. This flow meter allows direct mass measurement (independent of
fluid properties) and does not require a lengthy calibration process by the manufacturer each time the hydrogen
peroxide concentration is changed. The details of the chemical propulsion test facility utilized for the present
test series have been described in the past2.

The control of the mass flow regulators and the data acquisition were performed by an in-house developed
program running in LabVIEW.

Fig. 5: Schematic of the bipropellant test facility

A. Decomposition
The decomposition temperatures obtained in this tests series ranged between 550°C and 600°C (see fig. 6).
This indicates that at least 95-97 % of the hydrogen peroxide has been decomposed at the location of the
thermocouple. The transition time (time until thermal equilibrium is reached) is still rather large with up to 40 s.
This has to be optimized in the future since prolonged transition times reduce the performance of the thruster
and make pulsed operation difficult. Furthermore, during this transition time, large oscillations of the
decomposition temperature have occasionally been observed.

6
American Institute of Aeronautics and Astronautics
AIAA-2006-4551

700

600

Decomposition Temperature [°C]


500

400

300

200

100

0
0 50 100 150 200 250
Time [s]

Fig. 6: Decomposition temperatures obtained for several runs

The utilized catalyst consumed a total of 550 g of hydrogen peroxide. In fig. 6 the ratio of the maximum
temperature reached during each test to the theoretical possible is depicted as a function of the total consumed
mass. The 550 g of decomposed hydrogen peroxide is only about 25 % of the total mass the catalyst are
supposed to be able to consume. However, up to this point the catalyst has not shown any signs of performance
decrease. From the low mass loss of the catalyst over time (see fig. 7) one can further predict that the catalyst
will be able to operate for the total design time.

1.5 1

0.9

Max. Decomp. Temp/theor. temp


1.25
0.8

0.7
1
0.6
Weight [g]

[°C]
0.75 0.5

0.4
0.5
0.3
Weight G1 0.2
0.25
Weight G3
0.1
Max temp/ Theoretical temp
0 0
0 100 200 300 400 500 600
Total mass load [g]

Fig. 7: The constant decomposition temperature and the low loss of catalyst material indicates a sufficient lifetime of
the catalysts

B. Thruster
As discussed in section I, it was assumed that the failure to ignite the thruster was due to the very low
temperatures of the injected decomposed hydrogen peroxide. The main objective of the recent test was therefore
to increase those temperatures. Furthermore, an alternative green propellant, turpentine was tested.
The increase in temperature was obtained by utilizing hydrogen peroxide with higher concentrations and a
change of the thruster assembly (removing thermal mass such as the fixing plate). Furthermore, the fuel was
injected in liquid or in gaseous form. The latter was obtained with a commercially available vaporizer unit.
Since the power input into this vaporizer is rather large, this is not intended to be a part of the thruster system
but instead was intended at this point only as a help such that the ignition and combustion of the thruster can be
investigated.

In spite of the promising results with regard to the hydrogen peroxide decomposition and the high
temperatures obtained, it was not possible to achieve auto-ignition or combustion neither with ethanol nor with
turpentine. Several experiments with varying injection pressure, mass flow rates and fuel to oxidizer ratios have
been tried. None of those variations facilitated ignition followed by a stable combustion. The only exception of
this was when an external trigger electrode2 was injected via the throat area. With the help of this ignition
electrode, it was possible to ignite the thruster. However, as soon as the ignition electrode was pulled back, the
combustion stopped immediately. The reason for this behavior is discussed in the following section.
7
American Institute of Aeronautics and Astronautics
AIAA-2006-4551

IV. DISCUSSION
With regard to the hydrogen peroxide decomposition, it was shown that monolithic catalyst can indeed
decompose sufficient large mass flow rates with a relative high efficiency even when hydrogen peroxide in a
87.5 % concentration is used.
However, even with such high hydrogen peroxide concentration and consequently high decomposition
temperatures, it was not possible to generate conditions favorable to ignite the thruster. This was unexpected
considering the fact that several systems exist which are operating with such and even lower concentration.

However, comparison with existing hydrogen peroxide based thruster systems documented in the literature
reveals several major differences to the present system:
(i) Most hydrogen peroxide based thrusters use kerosene as propellant
(ii) All of the existing comparable systems operate with combustion chamber
pressures about one magnitude higher than in the present system (3 bar)
(iii) The present system has a lower propellant residence time than comparable systems

The fact that the present system uses ethanol instead of kerosene makes a comparison rather difficult.
However, it is not believed that the failure to ignite the thruster is due only to the particular choice of the fuel.

It is well-known that increasing the pressure in the combustion chamber increases the probability of ignition
due to an improved heat transfer. The driving effect is the improvement of heat transfer from the hot
decomposed hydrogen peroxide to the fuel. Unfortunately, most of the auto-ignition studies were again
performed for a kerosene/hydrogen peroxide combination. However, considering the similarities of those fuels,
it can be assumed that those tests are to a certain degree representative also for ethanol.

One of the most extensive auto-ignition studies was performed by Walder13. He developed the following
correlation for the auto-ignition of kerosene in hydrogen peroxide

( )
log L* ⋅ p1.15 =
3720
T
⋅ (− 2.62 )
1

with the characteristic length, L*, the combustion pressure p, and the combustion temperature T. The exponent
of the pressure term and the two values, have been found experimentally and they can vary with varying
experimental conditions. However, from this equation it is clear that with increasing combustion chamber
pressure, p, the probability of auto-ignition increases. This influence was again recently reported by Sadov for
hydrogen peroxide together with kerosene14. He defines pressure-temperature domain where self-ignition of
kerosene with H2O2 (92 – 98 % concentration) is possible as shown in fig. 8. Unfortunately, no similar data are
available for ethanol and hydrogen peroxide.

Fig. 8: Necessary temperature and pressure range for self-ignition for H2O2 and kerosene [Sadov]

It is obvious that the TPF-MRE with its design combustion chamber pressure of 3 bar (0.3 MPa) operates far
outside of the regime indicated in fig. 8 for kerosene and hydrogen peroxide. The main reason for choosing such
a low pressure was the pressure limitation of the pumps. The studies performed at the beginning of this project
showed that commercially available pumps can provide pressures only up to 6 bar. Including pressure drops in
8
American Institute of Aeronautics and Astronautics
AIAA-2006-4551

the turbine and the feeding lines, this resulted in a recommended combustion chamber pressure of roughly 3 bar.
Although from system level a justified choice, it did not include considerations with regard to the combustion
process such as recently published by Sadov (see fig. 8).
The necessity of higher combustion chamber pressure would also explain the ability to ignite the thruster
when an external trigger electrode is inserted into the combustion chamber via the throat. This very crude
method to ignite the thruster has, as discussed in the previous section, succeeded to ignite the thruster but the
combustion immediately stopped when the electrode was pulled out of the combustion chamber. Inserting the
electrode through the throat area has effectively decreased the throat area, or in other words, has increased the
combustion chamber pressure to a level where ignition is possible. Upon removing the electrode, the
combustion chamber pressure decreases again and the combustion cannot be sustained.

The combustion chamber design, in terms of characteristic parameter such as the characteristic length, L*,
was chosen based on comparisons with historical data. The characteristic length is defined as followed:
Vchamber
L* = = Rc ⋅ Lchamber (2)
Athroat

with the combustion chamber volume, Vchamber, the throat area, Athroat, the contraction ratio, Rc, and the length of
the chamber, Lchamber. Typical values15 of L* for a general bipropellant system range between 0.8 and 3.0 m and
between 0.6 and 1.35 m for systems based on hydrogen peroxide9. Although the characteristic length is a very
helpful parameter for the initial design of a combustion chamber, one needs to be careful to base the design of a
new system on historical data. The characteristic length is defined only by the geometry of the combustion
chamber and it is advisable to compare only systems with the same propellant combination and a narrow range
of mixture ratios and combustion chamber pressures.

Nevertheless, in the design of the TPF-MRE, the characteristic length was chosen to be equal to 1 m based on
comparison with other systems using hydrocarbon fuels in general but not based on systems using hydrocarbon
and hydrogen peroxide. The chamber diameter was chosen in an optimization process to reduce the wall surface
and therefore the heat flux from the combustion to the chamber walls. This optimization is extremely important
for small chambers since the thermal load on such small chambers is well-known to be the primary problem.
With regard to the minimization of the thermal load, a diameter of 15 mm was chosen. The chamber volume
was then evaluated based on the above mentioned characteristic length, the chamber diameter and the earlier
determined throat area.

This resulted in a combustion chamber length of 10 mm (not including the convergent section) and a
contraction ratio of 100. In comparison with other systems, the former is very short and the latter extremely high
(Rc is in general equal or smaller than 6). Altogether, the combustion chamber design results in a residence time
of roughly 150 – 200 µs.

Studies by Sisco et al.9 investigated the correlation between contraction ratio and auto-ignition characteristics
for a kerosene/hydrogen peroxide combination. They have not observed auto-ignition for a residence time below
75 µs. Increasing the residence time to 130 µs, they obtained auto-ignition. Although 130 µs is a value which
compares favorable with the residence time of 150 to 200 µs of the present system, one has to consider the
difference between ethanol and kerosene. Studies by Sadov14 have shown that ignition delays for an
ethanol/hydrogen peroxide system are about 4 - 5 times larger than for a comparable kerosene/hydrogen
peroxide system increasing therefore the necessary residence time beyond 130 µs for an ethanol system.

Another indication of the ignition problems with ethanol is the development of the MIT micro-rocket16,
which has initially envisioned the use of ethanol with oxygen. The use of pure oxygen instead of decomposed
hydrogen peroxide decreases the necessary residence time. Due to the small scale of the thruster, they
determined a residence time of 100µs. To assure proper propellant mixing and combustion they calculated a
necessary chamber pressure of roughly 10 to 20 bar. Still, even with this high chamber pressure and using
oxygen instead of hydrogen peroxide, they had to discard the idea of using ethanol due to ignition problems and
ended up using methanol.

V. CONCLUSIONS AND OUTLOOK


In summary, comparison between the present system and kerosene based systems indicates that the
combustion chamber design does not allow enough time for a successful mixing of the propellants which is
obviously a requirement for successful ignition. In particular, the combustion chamber length was designed too
short. The low combustion chamber pressure of only 3 bar is further worsening the conditions for a successful
mixing and ignition.
9
American Institute of Aeronautics and Astronautics
AIAA-2006-4551

Consequently, a thorough design change of the thruster has to be initiated. To support this effort, a simple
bread-board model is in preparation to identify the best combustion chamber design in terms of characteristic
length and contraction ratio and to obtain data for the ethanol/hydrogen peroxide system similar to those of
Sadov (see fig. 8).

In spite of the present problems with the TPF-MRE system, this effort is considered to be worthwhile due to
the obvious advantages of the presented system outlined in the various sections above.

ACKNOWLEDGEMENTS
This project was funded by the European Space Agency (ESA). We would like to thank the technical officer
at ESA, Martin Lang, for his help and cooperation during this project. Very much appreciated is the help of the
ARC-sr technicians Florin Plesescu, Michael Baca, and Karl Andres.

REFERENCES
[1] Miotti, P. et al., “Bi-propellant Micro-Rocket Engine”, AIAA-2004-3690, 40th Joint Propulsion
Conference, 2004
[2] Scharlemann, C., Marhold, K., Tajmar, M., Miotti, P., Guraya, C., Seco, F., Soldati, A.,
Campolo, M., Perennes, F., Marmiroli, B., Brahmi, R., Kappenstein, C., Lang, M., “Turbo-
Pump Fed Miniature Rocket Engine”, AIAA-05-3654, Joint Propulsion Conference, Tucson,
Arizona, 2005
[3] Tao Zhang, Tao Li, Jiwen Sun, Xiaodong Wang, Lei Ma, Dongbai Liang and Liwu Lin,
“Propulsive Performance of Hypergolic H2O2/Kerosene Bipropellant”, The First International
Conference on Green Propellants, The Netherlands, June 2001,
[4] Whitehead, J., “Hydrogen Peroxide Propulsion for Smaller Satellites”, SSC98-VIII-1, Preprint
for Utah State University Conference on Small Satellites, 1998
[5] Ponzo, J., “Small Envelope, High Flux 90% Hydrogen Peroxide Catalyst Beds”, AIAA-03-
4622, Joint Propulsion Conference, Huntsville, AL, 2003
[6] Bloom, R., Davis, N., Levine, S., “Hydrogen Peroxide as a Propellant”, J. of the American
Rocket Society, No. 80, March, 1950]
[7] M.Ventura, M., Wernimont, E., “History of the reaction Motors super Performance 90%
H2O2/Kerosene LR-40 rocket engine”, AIAA-01-3838, 37th Joint Propulsion Conference, Salt
Lake City, 2001
[8] Johnson, C., Anderson, W., Ross, R., “Catalyst Bed Instability within the USFE H2O2/JP-8
Rocket Engine”, AIAA-00-3301, 36th Joint Propulsion Conference, Huntsville AL, 2000
[9] Sisco, J., et al., “Autoignition of Kerosene by Decomposed Hydrogen Peroxide in a Dump
Combustor Configuration”, AIAA-03-4921, 39th Joint Propulsion Conference, Huntsville AL,
2003
[10] German, B. et al., ”An evaluation of Green Propellants for an ICBM Post-Boost Propulsion
System”, Missile Sciences Conference, Monterey, CA, 2000
[11] Musker, A., Roberts, G., Ford, S., Reakers, E., Westbury, T., “Auto-Ignition of Fuels Using
Highly Stabilized Hydrogen Peroxide”, AIAA-2005-4454, 41st Joint Propulsion Conference,
Tucson, Arizona, 2005
[12] Brahami, R. Batonneau, Y., Kappenstein, C., Miotti, P., Tajmar, M., Scharlemann, C., Lang,
M., "Cermaic Catalyst for the decomposition of H2O2. Influence of the wash-coat procedure
and the active phase", 8th International Hydrogen Peroxide Propulsion Conference, Purdue
University, Indiana, USA, September 2005
[13] Walder, H., “An investigation into the thermal ignition of hydrogen peroxide and kerosene”,
Report number RPD 7, Royal Aircraft Esablishment, May, 1950
[14] Sadov, V., “HP-Based Green Rocket Propellants”, International conference on Green
Propellants for Space Propulsion, Noordwijk, Netherlands, 2001
[15] Sutton, G., Biblaraz, O., Rocket Propulsion Elements, 7th edition,John Wiley &Sons, Inc, 2001
[16] London, et al., ”High-Pressure Bipropellant Microrocket Engine”, J. Propulsion and Power,
Vol. 17, No. 4, July-August, 2001

10
American Institute of Aeronautics and Astronautics

You might also like