Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Combustion and Flame 162 (2015) 2686–2692

Contents lists available at ScienceDirect

Combustion and Flame


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / c o m b u s t fl a m e

New HAN-based mixtures for reaction control system and low toxic
spacecraft propulsion subsystem: Thermal decomposition and possible
thruster applications
Rachid Amrousse a,b,⇑, Toshiyuki Katsumi c, Noboru Itouyama a, Nobuyuki Azuma b, Hideshi Kagawa b,
Keigo Hatai b, Hirohide Ikeda b, Keiichi Hori a
a
ISAS/JAXA, Japan Aerospace Exploration Agency, 3-1-1 Yoshinodai, Chuo-ku, Sagamihara, 252-5210 Kanagawa, Japan
b
TKSC/JAXA, Japan Aerospace Exploration Agency, 2-1-1 Sengen, Tsukuba, 305-8505 Ibaraki, Japan
c
Nagaoka University of Technology, 1603-1, Kamitomioka, Nagaoka, 940-2188 Niigata, Japan

a r t i c l e i n f o a b s t r a c t

Article history: HAN-based liquid monopropellant has been studied for its possible substitution of hydrazine toxic com-
Received 27 February 2015 ponent. The onset temperature of decomposition was provided by DTA-TG thermal analysis. Moreover,
Received in revised form 31 March 2015 the effect of 5 M HNO3 and 5 M NH2OH addition on catalytic decomposition was also examined. The
Accepted 31 March 2015
major products measured by mass spectrometer were N2, NO, N2O, NO2, H2O and NH3. Otherwise, the
Available online 17 April 2015
burning rates of HAN-based monopropellants were measured from the strand burner videos taken by
high speed camera. The effect of methanol addition; as fuel; on the burning rates was demonstrated.
Keywords:
On the other hand, HAN-based liquid monopropellant was tested and decomposed in 20 N thruster
HAN-monopropellant
High speed camera
and different catalysts were compared. Honeycomb catalysts were tested instead Shell 405 catalysts,
Combustion flame the obtained data demonstrate the improvement of burning reactions and pressure slopes after
Decomposition HAN-based decomposition.
Ó 2015 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

1. Introduction conducted to characterize the ignition and combustion behavior


of HAN. Pressure effects on the burning rates as well as flame
Hydroxyl ammonium nitrate (HAN, NH3OHNO3) is a promising structures of HAN and HAN-based liquid propellants were investi-
substitute for hydrazine because of its low toxicity, high density, gated [4–10]. The behavior of the propellant droplets was observed
high specific impulse, and low freezing point. As known, the con- in terms of micro-explosion and gasification characteristics
ventional monopropellant is hydrazine (N2H4), but the high vapor [11,12], the phenomenon that occurred between the liquid and
pressure of hydrazine and its high toxicity cause high storage and gas phase under high pressure and temperature environments
handling costs. Thus, the substitution of hydrazine by a less toxic [13,14], and droplet combustion [15]. The decomposition products
monopropellant is desired [1,2] in order to investigate low toxic of HAN were detected by employing various diagnostic techniques
spacecraft propulsion subsystem with green propellant. In artillery [15–19]. Chemical kinetic studies on HAN have been performed in
guns, one of the significant advantages of using liquid propellants order to determine reaction mechanisms and global kinetic rates.
rather those solid propellants is the capacity to control the Reactions for dilute HAN in the presence of a large excess of
combustion by metering the propellants [3]. Few years ago, among HNO3 were investigated [20–28]. N2 and NO formation reactions
possible hydrazine substitutes, aqueous energetic ionic liquids in HAN decomposition were also studied [23,29,30]. However,
comprising an ionic oxidizer, an ionic or molecular fuel, and the experiments were conducted with very low concentrations of
water as solvent, are often proposed and investigated as new HAN at low temperatures to maintain isothermal conditions.
monopropellants [1,2]. Moreover, extensive studies have been Based on the previous reaction studies [20–30] and experimental
observations, two reaction mechanisms were proposed for fused
HAN [31] and concentrated HAN mixtures [32].
⇑ Corresponding author at: JAXA, Japan Aerospace Exploration Agency, 3-1-1
Yoshinodai, Chuo-ku, Sagamihara, 252-5210 Kanagawa, Japan. Fax: +81 42 759
However, the adiabatic temperature of several liquid propel-
8284. lants are very high and in order to lower these temperatures, a
E-mail addresses: rachid.amrousse@yahoo.fr, rachid.amrousse@jaxa.jp lot of mixtures were tested such as HAN with methanol and
(R. Amrousse).

http://dx.doi.org/10.1016/j.combustflame.2015.03.026
0010-2180/Ó 2015 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
R. Amrousse et al. / Combustion and Flame 162 (2015) 2686–2692 2687

ethanol [33], HAN in the presence of TEAN (triethanol ammonium the methanol was used to control the burning rate of the
nitrate) [34], HAN with hydrazinium nitrate (HN) [35] for reaction monopropellant.
control systems and green propellant for low toxic spacecraft
propulsion subsystem. Moreover, the army has developed XM45 2.2. Thermal analysis
(LGP1845) and XM46 (LGP1846) for gun applications, which are
homogeneous mixtures of HAN, TEAN and water [36]. HAN decomposition was followed by differential thermal analy-
In an attempt to develop a new rocket propellant, many sis-thermal gravimetric (DTA-TG) apparatus (Riken) coupled to
formulations have been considered. The most recent application mass spectrometer (MS) at 1 atm (±1–2 °C as uncertainty analysis).
of green propellant was developed by Ball Aerospace and Known masses of catalyst (±0.2 mg as uncertainty mass measure-
Technology corporation which it used AF-M315E, it demonstrates ment) were put into aluminum cell holder followed by spray
stable performance of a flightweight attitude control thruster for injection of known volumes of liquid monopropellant. The mass
at least 10,000 total pulses [37–39]. The most recent formulation and temperature were recorded versus time, under helium flow
in Japan consists of HAN 73.6%, water, methanol as fuel and ammo- (50 ml min1) and heating rates were fixed at 5 °C min1 for ther-
nium nitrate (AN) as additive to stabilize the mixture for long-term mal decomposition and 1 °C min1 for catalytic decomposition.
storage [40]. Our previous research data showed that Shell 405 The analyses by DTA-TG enable to obtain the following informa-
catalyst was very active of HAN decomposition. However, this cata- tion: (i) evaporation of different solvents (endothermic peak); (ii)
lyst deactivate after several firing tests due to high exhaust gas onset temperature of decomposition given by the inflexion point
temperatures and oxidation medium [41]. Then, the use of struc- of the temperature curve; (iii) concentration of the HAN mixtures
tured catalysts as those supported on honeycomb monoliths has at decomposition; (iv) efficiency of the catalyst (exothermic peak)
been long considered in chemical industry and it has increased and (v) detection of emitted products by MS.
with the significance of environmental catalysis in pollution abate-
ment applications [42]. For reaction control systems (RCS), cata- 2.3. Strand burner
lysts with high attrition resistance and a low pressure drop are
required. In addition, the thin catalytic coating allows high effi- The combustion data of HAN-based liquid monopropellant at
ciency and selectivity [42–44]. The honeycomb monoliths can be elevated pressures were obtained from strand burner experiments
made of ceramic or metallic materials. More extensive studies on in a pressure vessel. The open-button strand burner was located in
ceramic monoliths were performed due to their wide use in the center of pressure vessel, into which the glass tube equipped
three-way automobile catalysts, mainly cordierite, commonly with wire and prefilled with HAN propellant as illustrated in Fig. 1.
used, among others, owing to its large-scale production based on The gas products vented out and the vessel was equipped with
extrusion which is quite cheap. Cordierite offers high mechanical an electrical feedthrough as resource to provide the required
strength, high resistance to elevated temperatures and tempera- energy for ignition of the propellant. A pressure regulator was used
ture shocks due to its low thermal expansion coefficient. for monitoring the static pressure into the vessel and windows for
Moreover, it presents great adhesion stability, which is a very light access. High-speed graph motion pictures were recorded with
important property for monolithic catalysts [43,45]. The most a high-speed camera (Roper Scientific, CR Imager, Model 2000);
usual way to coat structured supports such as honeycomb mono- with 250 fps as framing rate; equipped with Nikon lens at maxi-
liths is the well known washcoating procedure. The monolith is mum of 2200 frames per second. In a typical experiment, the
dipped in the slurry prepared with the catalytic material and after strand burner was equipped with prefilled glass tube with HAN
a certain time withdrawn carefully. The characteristic of the final propellant and placed into pressure vessel. The vessel was then
coating are a complex function of monolith properties, slurry prop- pressurized with nitrogen. Ignition of the propellant was accom-
erties and preparation conditions [43]. To ensure good perfor- plished by an electrical discharge. After test, the vessel was safely
mances of propellant-catalyst systems, highly active catalysts depressurized, opened, cleaned and moved to another test. In order
were prepared and tested for HAN decomposition at low tempera- to measure the burning rate, short intervals of images were
tures [40,41,46–48]. The main purposes of this paper are: (i) to obtained from video files. Then, Image J program was used to
determine the temperature onset and gas phase temperatures of measure the surface of liquid varied versus time.
HAN-based monopropellant by DTA-TG analysis; (ii) the analysis
of outlet gas products was performed by online mass spectrometer
(MS); (iii) the burning rates and combustion temperatures were
measured at constant pressures, the effect of methanol addition Wire
on the burning rates was carefully examined; because it is quasi-
impossible to apply HAN–H2O binary solution into reaction control
systems (RCS) due to its high adiabatic temperature and instable
combustion characteristics. On the other hand, HAN-based mono-
propellant solution contains methanol as fuel and AN as stabilizer
Glass tube
was decomposed by injection of known mass flows controlled by
electrovalve into preheated catalytic bed in vacuum chamber.
Solution
2. Experimental

2.1. Monopropellant solution

HAN-based liquid monopropellant was prepared by Hosoya


Company (Tokyo, Japan). The solution composition is HAN
73.6 wt.% – AN 3.9 wt.% – H2O 6.2 wt.% – MeOH 16.3 wt.% (HAN/
AN/H2O/MeOH = 95/5/8/21; stoichiometric conditions); which AN
is ammonium nitrate and MeOH is methanol. AN was added to
HAN solution in order to decrease the freezing point. However, Fig. 1. Glass tube contains HAN solution and equipped with wire igniter.
2688 R. Amrousse et al. / Combustion and Flame 162 (2015) 2686–2692

2.4. Thruster tests 100 0


191 °C
The HAN thruster 20 N was already described in previous
papers [41,49], in which the pressure slope was placed in the com-

Diff. temperature (ºC mg-1)


75
bustion chamber (after catalytic bed). It was developed in JAXA
several years ago as part of research project. The upper part of 201 °C -5

Weight (%)
HAN-based propellant tank was filled with N2 gas from a gas cylin-
der. The N2 pressure in the tank was fixed for HAN catalytic 50
decomposition reaction in order to give an almost constant feeding
rate of HAN. The monolithic catalyst was charged into the thruster.
- 10
Then, the catalyst bed was preheated before test at 210–300 °C.
The preheating of catalyst was performed by flexible sheath heater 25 154 °C
manufactured by SAKAGUCHI E.H VOC CORP with 88 W as power
and 35 VDC as power-supply voltage.
When the electromagnetic valve was opened, HAN-based 0 - 15
monopropellant was pressed by the N2 gas into the thruster. 0 5 10 15 20 25 30 35 40
Ignition delay, catalyst bed temperatures, chamber pressures and Time (min)
feeding rate of HAN were recorded by Lab-view software (NI
USB-6229) with the sampling rate of 1000 Hz. The accuracy of Fig. 2. Thermal decomposition of 95/5/8/21 HAN-based mixture.

pressure sensor is ±0.5% FSO (Full Scale Output) which is equal to


0.076 MPa.
0 0
The opening time of the electromagnetic valve was varied from
1 to 5 s. The average mass flow for each thruster firing test is
around 12.0 g s1 (±0.05 ml as electrovalve accuracy) and -20
1.2 MPa as initial pressure. After tests, the catalytic bed was purged -5
by N2 flow, which to remove the produced gases and to decrease

Weight (wt.%)
ΔΤ (°C mg )

temperatures.
-1

-40

-10
3. Results and discussion
-60

The tree different experiments carried out in this study can be


-15
distinguished by their heating method and initial pressure condi- -80 65 °C
tions. Firstly, during the thermal analysis, the propellant was trea-
ted under gradual increase of temperature at 1 atm. Secondly; the
combustion flame structure was performed under external electric -100 -20
power at different pressures. Finally, HAN monopropellant was 0 10 20 30 40 50 60 70
injected into preheated catalytic bed. The presented results were Time (min)
showed the possibility to apply HAN-based as green monopropel-
Fig. 3. Catalytic decomposition of 95/5/8/21 HAN-based mixture over Ir-based
lant into space applications; such as reaction control systems (RSC) catalyst.
and low toxic spacecraft propulsion subsystem.

3.1. Thermal analysis has been observed in the case of HAN–water [33], however, this
phenomenon was absent by methanol addition that can accelerate
Thermal and catalytic decomposition processes of HAN-based the evaporation process. Based on experimental results, methanol
liquid monopropellant with additives; followed by online appears to play a very major role in the decomposition of HAN-
exhausted gas analysis; were carried out by DTA-TG/MS. The evo- based.
lutions of temperature and weight loss as a function of time during Therefore, in both cases, the decomposition starts only when
the thermal decomposition process are represented in Fig. 2. The the solvents have been quantitatively removed leading to almost
thermal decomposition of HAN-based has been started around pure liquid HAN (melting point 48 °C). From these results, we
154 °C after a large endothermic reaction which demonstrated by can conclude that solvents exhibit, as expected, a stabilizing effect
total evaporation of methanol and water as additives. The thermal on HAN, which is the consequence of dispersed ions (NH3OH+ and
decomposition occurs in two short steps: (i) highly exothermic NO3 ) in solutions. The ionic reagents must therefore be in contact
reaction with a maximum at 191 °C, which correspond to the high with each other for decomposition to engender.
loss of HAN weight, while (ii) the second step was less exothermic The decomposition rates of HAN–H2O solution were signifi-
and reach maximum at 201 °C. cantly influenced by the amount of methanol. Indeed, the metha-
During the thermal decomposition process, an endothermic nol was added to HAN/H2O solution in order to decrease its
peak corresponding to the evaporation of methanol and water burning rate, obtain higher Isp than hydrazine even in no-
has been observed and resulted in concentration of solution. detonation region. Moreover, the methanol addition makes HAN
Then a second, rapid weight loss happens, correspond to exother- solution in category 4 (harmful and swallowed) of GHS (Globally
mic decomposition of HAN/AN. Harmonized System) of classification and labeling of chemicals;
Figure 3 shows obtained temperature and weight loss profile which is defined by LD50rat (Lethal Dose). LD50 of HAN solution
versus time in the presence of catalyst. The temperature profile estimated to 300–2000 mg kg1, which is much lower in the case
was changed in the case of catalytic decomposition process to a of hydrazine (60 mg kg1).
violent exothermic reaction of HAN-based in one step at low tem- It should be noted that the catalytic performance varied from
perature onset. Moreover, the preliminary low exothermic process catalyst to other due to nature of noble metal, preparation, metal
R. Amrousse et al. / Combustion and Flame 162 (2015) 2686–2692 2689

2,5 10
4 Table 2
m/z = 17 Synthesized HAN-based mixtures with 5 M NH2OH addition.
m/z = 18
m/z = 28 Mixtures Final Ratio HAN- NH2OH HAN-based
2 10
4 m/z = 30 volume based:NH2OH volume volume (ml)
m/z = 44 (ml) (ml)
m/z = 46
Absolute intensity (a.u.)

m/z = 63 HAN-based 5.0 0 0.0 5.0


4 HAN-based + 1:10 5.5 1:10 0.5 5.0
1,5 10
HAN-based + 2:10 6.0 2:10 1.0 5.0
HAN-based + 3:10 6.5 3:10 1.5 5.0
HAN-based + 4:10 7.0 4:10 2.0 5.0
4
1 10 HAN-based + 5:10 7.5 5:10 2.5 5.0

5000

140
HNO3 (5M)
0 HNO3 addition
NH2OH (5M)
0 5 10 15 20 25 120
Time (min)

Onset temperature (°C)


100 NH OH addition
Fig. 4. MS-online analysis of thermal decomposition of 95/5/8/21 HAN-based 2

mixture.
80

content etc. The present catalyst based on iridium was highly 60


active and represents similar temperature onset as well as com-
mercial Shell 405 catalyst (70 °C). The chemical reaction products 40
were detected by online MS analyzer. Figure 4 shows the most
relevant mass fragmentations (m/z) after catalytic decomposition
20
of HAN-based mixture as a function of time and absolute intensity.
The error bars represent a 95% confidence interval. It has been
0
reported that N2 can be formed by following reactions [17,25,50]:

HAN-based + 1:10

HAN-based + 2:10

HAN-based + 3:10

HAN-based + 4:10

HAN-based + 5:10
HAN-based

NH2 OHðgÞ þ HNO ! N2 ðgÞ þ 2H2 OðgÞ

HNO þ HAN ! N2 ðgÞ þ 2H2 OðgÞ þ HNO3 ðgÞðrecombinationÞ


Otherwise, N2O and NO gaseous phases were reported to be
formed by the possible reactions of NH2OH or NH3OH+ with insta-
Fig. 5. Catalytic decomposition of 95/5/8/21 HAN-based mixture in the presence of
ble intermediates HONO and HNO. NO2 probably formed from HNO3 and NH2OH additional volumes.
HNO3 decomposition.
In addition, it was not possible to explain all the measured
decomposition products in this study, which are shown in Fig. 4, showed the effect of additional 5 M HNO3 on onset temperature
by these mechanisms. Therefore, some reactions obtained from delay, however, additional 5 M NH2OH accelerates the onset tem-
the literature survey were added to the previous reaction mecha- peratures of decomposition. According to these data, the chemical
nisms. It is agreed upon that the first step of HAN decomposition reactions of HAN-based monopropellant without addition and
proton (H+) transfer from NH3OH+ to NO 3 producing NH2OH and HAN-based with additional species were reacted differently into
HNO3. Since NH2OH is a much stronger base than water [51], any the catalytic surface. The analysis of online gas products showed
protons that HNO3 releases will associate with NH2OH, not water. similar product species; it means that the catalytic decomposition
Consequently, HNO3 dissociation releasing a proton will simply even with additional species does not have effect on the thermal
reverse, leading to no net reaction. decomposition process of HAN-based. Nevertheless, from these
In addition, Tables 1 and 2 lists detailed information of two pre- results, may present Ir catalyst was selective to decompose totally
pared series of HAN-based. The purpose was to understand the NH2OH. However, it was not active to decompose HNO3 totally in
effect of addition of 5 M HNO3 and 5 M NH2OH on the catalytic the gas phase.
decomposition behavior of HAN-based in the presence of Ir-based In Fig. 5, representative results of onset temperatures of
catalyst. The obtained results were presented in Fig. 5, which decomposition are illustrated for HAN-based after addition of dif-
ferent amounts of HNO3 and NH2OH.
Table 1 In order to verify the selectivity of Ir-based catalyst, two differ-
Synthesized HAN-based mixtures with 5 M HNO3 addition. ent tests: (i) the decomposition of 60% HNO3 and (ii) 50% NH2OH
were performed. NH2OH totally decomposed at very low tempera-
Mixtures Final Ratio HAN- HNO3 HAN-based
volume based:HNO3 volume volume (ml) ture (45–50 °C). Though, HNO3 was only evaporated due to large
(ml) (ml) endothermic reaction. Therefore, the catalytic decomposition of
HAN-based 5.0 0 0.0 5.0
HAN-based monopropellant was started by total decomposition
HAN-based + 1:10 5.5 1:10 0.5 5.0 of NH2OH to form intermediate species such as HNO and HNO2
HAN-based + 2:10 6.0 2:10 1.0 5.0 acids. These instable acids reacted with non reacted HAN and
HAN-based + 3:10 6.5 3:10 1.5 5.0 non reacted HNO3 to form more stable species as well as gas outlet:
HAN-based + 4:10 7.0 4:10 2.0 5.0
NO, N2, N2O, NO2, H2O etc. [52]. Moreover, HNO3 can be a good sta-
HAN-based + 5:10 7.5 5:10 2.5 5.0
bilizer of HAN-based mixtures.
2690 R. Amrousse et al. / Combustion and Flame 162 (2015) 2686–2692

3.2. Combustion flame structure 1.0e+02

The ignition of HAN-based monopropellant was carried out on

Burning rate (mm s-1)


strand burner by external electric power at constant pressures. 1.0e+01
Figure 6 shows different pictures of HAN-based combustion tests
taken by high speed video camera. According to the obtained 1.0e+00
results, the burning rates have been influenced by initial pressures.
In fact, small and white generated bubbles were observed in the
case of P = 1 MPa may due to the evaporation of water. However; 1.0e-01
from P = 3 MPa; the bubble volumes increased and the color of
transparent solution became brown, this phenomenon may caused
by the formation of NO2 gas, evaporation of water, and partial and 1 2 3 4
slow decomposition of HAN-based. At P = 4 and 6 MPa, the bubble Pressure (MPa)
diameter became smaller when we increased initial pressure. It
means that the bubbles of decomposition were generated but the Fig. 7. Effect of methanol addition on the calculated burning rates of HAN/AN-
high pressure assumes their decrepitating (more clear in case of based mixture in the strand burner.

P = 8 MPa).
On the other hand, Fig. 7 shows the measured burning rates of
HAN solutions with different mass percents of methanol. 1.0e+02
: 95/5/8/21
According to those data, the burning rates of 95/5/8/0, 95/5/8/29
: 95/5/8/21 + Ir
and 95/5/8/50 mixtures jumped from the moderate rates to extre-

Burning rate (mm s-1)


mely high rates. The burning jump was absent in the case of 95/5/ 1.0e+01
8/21 mixture which present the best linear burning rate equation.
In fact, the initial mass of methanol has a strong effect on the
stability of burning rates. 1.0e+00
For other mixtures, they can be applied as propellant mixtures
according to their high burning rates. However, the jump of the
burning rates to high regions cannot be demonstrated by combus- 1.0e-01
tion mechanism, but rather by hydrodynamic stability [53]; which
was explained by Landau–Levich instability theory for liquid pro-
pellants by adding reactive/diffusive instability and applied the
1 10
extended theory to an HAN-based monopropellant. Otherwise,
Pressure (MPa)
supported Ir-particles were mixed with HAN-based liquid mono-
propellant at room temperature. Figure 8 presents the measured Fig. 8. Effect of iridium particles on the calculated burning rates of HAN-based
burning rates of HAN-based solution (95/5/8/21) mixed with sup- mixture in the strand burner.
ported iridium particles on alumina (as catalyst). The results
showed that the measured burning rates are higher than those
bed temperature above 210 °C, but in order to obtain a nominal
obtained after combustion of HAN-based mixture alone and these
start 350 °C is preferred for honeycomb catalysts.
data can be explained through the following reason: before igni-
Figure 9a and b (zoom) illustrate the evolution of pressure as a
tion, the electric power activates the Ir-particles at the surface by
function of time during the burning reaction processes of HAN-
temperature increasing; and the hot Ir-particles decrepitate the
based. The burning reactions of HAN-based liquid monopropellant
generated bubbles.
were conducted over honeycomb catalysts using 20 N thruster
showed that the pressure slopes in the case of Ir honeycomb cata-
3.3. Thruster tests lysts 400 and 600 cpsi were higher than commercial Shell 405 cata-
lyst. In fact, the pressure slope in the case of commercial Shell 405
It should be mentioned that before firing the thruster, the reac- catalyst was around 1.5 MPa s1, while this value was doubled
tor is pre-heated to a sufficient temperature, typically 210–350 °C to 3.5 MPa s1 in case of 30 wt.% Ir honeycomb catalysts.
(depending on bed load). The reaction will start already with a heat Moreover, the three curves showed second pressure slopes due

P = 1 MPa P = 3 MPa P = 4 MPa P = 6 MPa P = 8 MPa


Fig. 6. Pictures taken by high speed camera during combustion of 95/5/8/21 HAN-based mixture in the strand burner.
R. Amrousse et al. / Combustion and Flame 162 (2015) 2686–2692 2691

0.5 0.4

0.4
0.3

Pressure (MPa)
Pressure (MPa)

0.3
0.2
- S405 (Shell catalyst)
- Ir-CuO/honeycomb 600 cpsi
0.2 - Ir/honeycomb 400 cpsi
- S405 (Shell catalyst)
- Ir/honeycomb 600 cpsi 0.1
- Ir/honeycomb 400 cpsi
0.1

0 0.5 1.0 1.5 2.0


0 0.5 1.0 1.5 2.0 Time (s)
Time (s) (a)
(a)
0.4
0.5

0.3

Pressure (MPa)
0.4
Pressure (MPa)

0.3 0.2

0.2 0.1

0.1
0 0.05 0.1 0.15 0.2
Time (s)
0 0.1 0.2 0.3 0.4 (b)
Time (s)
Fig. 10. (a) Burning reaction tests of 95/5/8/21 HAN-based mixture over 3 different
(b) catalysts with different active phases on 20 N thruster and (b) zoom of reaction
timing.
Fig. 9. (a) Burning reaction tests of 95/5/8/21 HAN-based mixture over 3 different
Ir-based catalysts on 20 N thruster and (b) zoom of reaction timing.

to second reactions of gas phase products over catalysts. The sec- slope was accelerated and it was faster. Moreover, it is well known
ond pressure slope has been started earlier (after 1 s of monopro- that the decomposition of HNO3 is very slow and requires high
pellant injection) in the case of Ir honeycomb catalyst 600 cpsi. activation energy; for this, the temperature of preheated catalytic
Contrariwise, slopes were delayed in the case of commercial cata- bed reactor was varied between 210 and 350 °C. On the other hand,
lyst and honeycomb 400 cpsi. The difference could be demon- Kajiyama et al. studied the effect of CuO on the thermal
strated by the size of channels of catalyst surfaces; honeycomb decomposition of AN [54]. In fact, during the thermal process, they
600 cpsi has small size channels compared to honeycomb 400 cpsi. assumed that CuO allows the formation of copper complexes such
Indeed, the generated gas products have more contact or longer as [Cu(NH3)2](NO3)2, which increases the reaction rate of AN pre-
residence time within 600 cpsi catalyst surface; that mean gaseous sent in HAN-based monopropellant as stabilizer with the intention
phase undergoes fast reaction with solid catalyst surface in this to decease its freezing point.
case.
We can conclude that HAN-based monopropellant mixture can 4. Conclusion
be applied as candidate of hydrazine replacement due to high and
sudden pressure slopes. In fact, the obtained results can be According to the synthesized HAN-based solution properties
improved by modifying some parameters such length of catalytic that present several advantages such as high density and high
bed, injection system and heater performance. specific pulse, this monopropellant can be applied for reaction con-
In order to improve the delay of second pressure slopes, new trol system instead toxic hydrazine:
honeycomb catalyst 600 cpsi based on Ir and CuO active phases.
Obviously, the honeycomb catalyst was successively impregnated – The present (20%) Ir catalyst showed high performance on the
by Ir- and then Cu-precursors in order to file 20 wt.% and 10 wt.% HAN-based decomposition and it can substitute Shell 405 cata-
of both active phases; respectively. From the firing test results pre- lyst. However, Ir active phase was selective to decompose
sented in Fig. 10a and b (zoom) in which the pressure was pre- totally NH2OH and same catalyst suffered against HNO3 gas.
dicted as a function of time, Ir–CuO biphasic catalyst has two – The burning rates measured from strand burner tests were
advantages compared to Ir monophasic catalysts: (i) same value involved by addition of methanol on HAN/AN mixtures. 95/5/
of pressure slope than other Ir honeycomb catalysts but economi- 8/21 mixture presents most acceptable burning rates.
cally more cheaper, and (ii) accelerates the second pressure slope However, the burning rates of other mixtures have been jumped
for about 1–1.5 s. Thus, the maximum pressure value (0.4 MPa) after attend 4 MPa as pressure. The burning rates have been
has been reached in a short time. This catalyst may be active for increased after mixing HAN-based propellant with supported
HNO3 decomposition in the gas phase state that is why the second Ir-particles.
2692 R. Amrousse et al. / Combustion and Flame 162 (2015) 2686–2692

– The new synthesized Ir–CuO monolithic catalyst showed high [18] H.S. Lee, S.T. Thynell, in: Proceedings of the 33rd AIAA/ASME/SAE/ASEE Joint
Propulsion Conference, July 6–9, 1997, Seattle, WA, pp. 3232–3241.
performance for HAN-based decomposition. In fact, the burning
[19] Y.J. Lee, T.A. Litzinger, Combust. Sci. Technol. 141 (1999) 19–27.
reactions and the pressure slopes were improved comparing [20] A. Bothner-By, L. Friedman, J. Phys. Chem. 20 (1952) 459–462.
data to those obtain by Shell 405 commercial catalyst. Also, [21] M.N. Hughes, G. Stedman, J. Chem. Soc. B (1963) 2824–2830.
the cost of honeycomb catalysts was much lower than commer- [22] T.D.B. Morgan, G. Stedman, M.N. Hughes, J. Chem. Soc. B (1968) 344–349.
[23] M.R. Bennett, G.M. Brown, L. Maya, F.A. Posey, Inorg. Chem. 21 (1982) 2461–
cial catalyst. 2468.
[24] G.C.M. Bourke, G. Stedman, J. Chem. Soc. Perkin Trans. 2 (1992) 161–162.
[25] J.R. Pembridge, G. Stedman, J. Chem. Soc. Dalton Trans. 11 (1979) 1657–1663.
[26] R.J. Gowland, G. Stedman, J. Inorg. Nucl. Chem. 43 (1981) 2859–2862.
Acknowledgments [27] R.J. Gowland, G. Stedman, J. Chem. Soc. Chem. Commun. (1983) 1038–1039.
[28] E. Jones, G. Pota, G. Stedman, Catal. Lett. 24 (1994) 211–214.
Authors would like to thank Japan Aerospace Exploration [29] F.T. Bonner, L.S. Dzelzkalns, J.A. Bonucci, Inorg. Chem. 17 (1978) 2487–2494.
[30] F.T. Bonner, N. Wang, Inorg. Chem. 25 (1986) 1858–1962.
Agency for the project support. Dr. Rachid Amrousse presents his [31] J.C. Oxley, K.R. Brower, SPIE 872 (1988) 63–69.
thank to JAXA (Sagamihara branch and Tsukuba branch) for the [32] N. Klein, Ignition and combustion of the HAN-based liquid propellants, in:
grant of research project, Yokohama and Tokyo universities for Proceedings of the 27th JANNAF combustion subcommittee meeting, vol. 1.
CPIA Publication, Sunnyvale, 1990, pp. 443–450.
different experimental devices. [33] R. Amrousse, T. Katsumi, T. Sulaiman, B.R. Das, H. Kumagai, K. Maeda, K. Hori,
Inter. J. Energy Mater. Chem. Propul. 11 (2012) 241–257.
References [34] Y.J. Lee, T.A. Litzinger, Combust. Sci. Technol. 141 (1999) 19–36.
[35] F.B. Apollo, N. Sakae, M. Haruke, A. Muneo, IHI Eng. Rev. 43 (2010) 22–28.
[36] N.A. Messina, L.S. Ingram, M. Tarczynski, J.D. Knapton, Prop. Explos. Pyrotech.
[1] A. Cybulski, J.A. Moulijn, Structured Catalysts and Reactors, Marcel Dekker,
16 (1991) 88–93.
New York, 1993.
[37] M.C. Deans, B.D. Reed, L.A. Arrington, in: 50th AIAA/ASME/SAE/ASEE Joint
[2] J.A. Moulijn, V. Leeuwen, V. Santen, Stud. Surf. Sci. Catal. 79 (1993) 5–8.
Propulsion Conference, July 28–30, 2014, Cleveland, OH, pp. 3480–3490.
[3] W.F. Morisson, J.D. Knapton, M.J. Bulman, in: L. Stiefel, M. Summerfield (Eds.),
[38] R.A. Spores, R. Masse, S. Kimbrel, in: 50th AIAA/ASME/SAE/ASEE Joint
Gun Propul. Technol., Progress in Astronautics and Aeronautics, vol. 109, AIAA,
Propulsion Conference, July 28–30, 2014, Cleveland, OH, pp. 3482–3493.
Washington, DC, 1988, p. 143.
[39] C.H. McLean, W.D. Deininger, J. Joniatis, in: 50th AIAA/ASME/SAE/ASEE Joint
[4] R.H. Comer, Proceedings of the 16th International Symposium on Combustion,
Propulsion Conference, July 28–30, 2014, Cleveland, OH, pp. 3481–3501.
The Combustion Institute, Pittsburgh, 1976.
[40] R. Amrousse, W. Fetimi, K. Farhat, K. Hori, Appl. Catal. B 127 (2012) 121–128.
[5] S.R. Vosen, 22nd International Symposium on Combustion, The Combustion
[41] R. Amrousse, T. Katsumi, A. Bachar, R. Brahmi, M. Bensitel, K. Hori, Reac. Kinet.
Institute, Pittsburgh, 1988.
Mech. Catal. 111 (2014) 71–88.
[6] S.R. Vosen, Combust. Sci. Technol. 68 (1989) 85–99.
[42] A. Cybulski, J.A. Moulijn, Catal. Rev.: Sci. Eng. 36 (1994) 179–270.
[7] S.R. Vosen, Combust. Flame 82 (1990) 376–388.
[43] P. Avila, M. Montes, E.E. Miro, Chem. Eng. J. 109 (2005) 11–36.
[8] W.F. Oberle, G.P. Wren, Closed chamber combustion rates of liquid
[44] V. Tomasic, F. Jovic, Appl. Catal. A 311 (2006) 112–121.
propellant 1846 conditioned ambient, hot and cold vulnerability testing of
[45] J.L. Williams, Catal. Today 69 (2001) 3–9.
liquid propellant LGP 1846, Proceedings of 27th JANNAF Combustion
[46] X. Ren, M. Li, A. Wang, L. Li, X. Wang, T. Zhang, Chin. J. Catal. 28 (2007) 1–2.
Subcommittee Meeting, vol. 557, CPIA Publication, Huntsville, 1990, pp.
[47] C.H. Hwang, S.W. Baek, S.J. Cho, Combust. Flame 161 (2014) 1109–1116.
377–385.
[48] E.W. Schmidt, D.F. Gavin, Catalytic decomposition of hydroxyl ammonium
[9] W.F. Oberle, G.P. Wren, Burn rates of LGP 1846 conditioned ambient, hot, and
nitrate-based monopropellants, US005485722A Patent 5,485,722, 1996.
cold, ballistic research laboratory technical report no. BRL-TR-3287, 1991.
[49] R. Amrousse, T. Katsumi, Y. Niboshi, N. Azuma, A. Bachar, K. Hori, Appl. Catal. A
[10] S.T. Jennings, Y. Chang, D. Koch, K.K. Kuo, Peculiar combustion characteristics
452 (2013) 64–68.
of XM46 liquid propellant, in: Proceedings of 34th JANNAF combustion
[50] W.A. Rosser, S.H. Inami, H. Wise, J. Chem. Phys. 67 (1963) 1753–1757.
subcommittee meeting, vol. 662, 1997, pp 321–331.
[51] D.D. Perrin, IUPAC Chemical Data Series: No. 29 Ionization Constants of
[11] D.L. Zhu, C.K. Law, Combust. Flame 70 (1987) 333–342.
Inorganic Acids and Bases in Aqueous Solution, Pergamon Press, NY, 1982.
[12] C. Call, D.L. Zhu, C.K. Law, S.C. Deevi, J. Propul. Power 13 (1997) 448–450.
[52] H. lee, T.A. Litzinger, Combust. Flame 135 (2003) 151–169.
[13] M.E. Kounalakis, G.M. Faeth, Combust. Flame 74 (1988) 179–192.
[53] T. Katsumi, T. Inoue, J. Nakatsuta, K. Hasegawa, K. Kobayashi, S. Sawai, K. Hori,
[14] T.W. Lee, L.K. Tseng, G.M. Faeth, J. Propul. Power 6 (1990) 382–391.
Combust. Explo. Shock 48 (2012) 536–543.
[15] C.A. Van Dijk, R.G. Priest, Combust. Flame 57 (1984) 15–24.
[54] K. Kajiyama, Y. Izato, A. Miyake, J. Therm. Anal. Calorim. 113 (2013) 1475–
[16] J.T. Cronin, T.B. Brill, J. Phys. Chem. 90 (1986) 178–181.
1480.
[17] J.T. Cronin, T.B. Brill, Combust. Flame 74 (1988) 81–89.

You might also like