You are on page 1of 10

Further Development of an Aluminum and Water Solid Rocket Propellant

David E. Kittell1, Timothe L. Pourpoint2, Lori J. Groven3, and Steven F. Son4 Purdue University, West Lafayette, Indiana, 47907

Nanoscale aluminum and water has been used as a stepping stone towards in-situ rocket propellants and as a testbed for nanoenergetic composite propellants. A baseline formulation of nanoscale aluminum and water was developed and demonstrated with a sounding rocket flight in 2009. Performance of the propellant was not optimized, hence a reformulation was sought with an emphasis on improved safety and more efficient combustion. The chosen reformulation is a bimodal powder distribution of 70 wt.% Novacentrix 80 nm Al and 30 wt.% Valimet 2 m Al at an equivalence ratio of 0.813 (optimized for sea level Isp). The mixture also includes 3 wt.% ammonium dihydrogen phosphate, to inhibit the slow reaction of nanoaluminum with water, and 1 wt.% polyacrylamide to improve material suspension. Ammonium dihydrogen phosphate can protect nanoaluminum in solution for several hours, but degredation can occur while mixing, and pH increases from slightly acidic to basic with increased mixing time and temperature. The stress of mixing might be removing the coating and exposing nanoaluminum to water. It is also shown that nanoaluminum reacts faster in basic aqueous solutions than in solutions with neutral pH. Static motor tests reveal that propellant formulations with neutral pH provide better performance. Implementations of shorter mixing times and reduced temperatures are used to control the pH of the propellant, resulting in increased Isp values of as much as 30%.

Nomenclature
%Ac ADP ALICE ALICE II At CEA cf c* Isp LB m MW o/f poly-A UB C* = = = = = = = = = = = = = = = = = active content of aluminum in fuel ammonium dihydrogen phosphate nanoaluminum and water propellant reformulated nanoaluminum and water propellant throat area chemical equilibrium analysis thrust coefficient propellant characteristic velocity specific impulse lower bound mass molecular weight oxidizer to fuel ratio polyacrylamide upper bound combustion efficiency equivalence ratio

Graduate Student, Aerospace Engineering, 500 Allison Rd. West Lafayette, IN 47907, Student Member. Research Assistant Professor, Aerospace Engineering, 500 Allison Rd. West Lafayette, IN 47907, Senior AIAA Member 3 Senior Research Scientist, Mechanical Engineering, 500 Allison Rd. West Lafayette, IN 47907, AIAA Member 4 Professor, Mechanical Engineering, 500 Allison Rd. West Lafayette, IN 47907, Senior AIAA Member 1 American Institute of Aeronautics and Astronautics
2

I. Introduction
N-SITU propellants could revolutionize the logistics of any weight constrained problem. For rocket applications, calculations show that effective specific impulse (Isp) is doubled if half of the propellant is obtained en route [1]. With renewed evidence of water or ice on the moon and Mars [2,3] a water and aluminum propellant could be one refueling option for manned space missions. Historically, water and aluminum has been proposed as an underwater propellant and as a means to store and release hydrogen, but none of these roles have been possible until the development of commercially available nanoscale aluminum. The high specific surface area of nanoaluminum lowers the activation energy of the reaction to useful levels. Research on the feasibility of an aluminum and water solid rocket motor led to the ALICE propellant (deionized water and Novacentrix 80nm at = 0.75) and to a flight demonstration [4]. This baseline ALICE formulation was not optimized but served as a stepping stone for on-going research. Performance of ALICE is tailorable by using different mixtures of aluminum powders and through choice of stoichiometry. The high regression rate of the baseline propellant can be decreased with increasing micron aluminum content, however a larger improvement is sought in combustion efficiency. As the propellant burns it leaves behind varying amounts of molten aluminum as slag. A new approach investigates correlations between propellant performance and pH. This research selects an optimized ALICE reformulation and further shows how pH affects the solid rocket motor performance. The paper is divided into two major sections, with the first half focusing on propellant reformulation: design, manufacture, and the resulting qualities. The second half concerns rocket motor performance and covers the results of small scale rocket motor firings at two different pH levels.

II. Propellant Reformulation


A. Theoretical Performance The amount of unoxidized aluminum in a powder is described by a percent active content (%Ac). This is an important parameter for theoretical calculations because it determines how much alumina is present on the fuel particles. For nanoscale aluminum, oxide content can be significant (on the order of 20 to 40 wt.%) whereas for micron-sized particles, oxide content is typically lower (below 1 wt.%). NASA chemical equilibrium code [5] is used to plot sea level Isp and adiabatic flame temperature as a function of equivalence ratio and active content. Initial reactant temperatures are set to 255 K to reflect the frozen nature of the propellant, and the results are shown below in Figs. 1 and 2.
235 230
100% 94% 88% 82% 76% 70% Active Aluminum Content

Sea Level Isp, sec

225 220 215 210 205 200 195 190 0.4 0.6 0.8

Equivalence Ratio,

1.2

1.4

1.6

Figure 1. Theoretical specific impulse of water and aluminum mixtures as a function of equivalence ratio, , and active aluminum content. NASA CEA calculations assume a perfectly expanded nozzle at sea level with a chamber pressure of 6.89 MPa and initial reactant temperature of 255 K.

2 American Institute of Aeronautics and Astronautics

Adiabatic Flame Temperature, K

3200 3000 2800


f

2600
a b c

2400
Active Aluminum Content:

2200 2000 1800 0.4 0.6 0.8 1 1.2

a = 70% b = 76% c = 82% d = 88% e = 94% f = 100%

Equivalence Ratio,

1.4

1.6

Figure 2. Adiabatic flame temperature of water and aluminum mixtures as a function of equivalence ratio, , and active aluminum content. NASA CEA calculations assume a perfectly expanded nozzle at sea level with a chamber pressure of 6.89 MPa and initial reactant temperature of 255 K. From these calculations, sea level Isp peaks at 230 s and adiabatic flame temperature is limited to about 3000 K both of these results agree with previous work [6]. From this data, optimum Isp is located off stoichiometric conditions tending to be fuel lean with decreasing active aluminum content. However, maximum flame temperatures occur at stoichiometric conditions and fall off steeply for either fuel lean or fuel rich conditions. The discontinuity in flame temperatures occurs near the melting point of alumina at 2327 K [7] because thermal energy is required for the heat of fusion of alumina. The reformulated propellant is mixed for maximum sea level Isp in contrast to previous research that assumed maximum performance coincides with stoichiometric conditions [8,9]. Wood [8] recommends an equivalence ratio of 1 to take advantage of the higher flame temperatures and Connell et al. [9] demonstrate in small-scale rocket motor tests that higher equivalence ratios perform better. Unfortunately, these formulations are much harder to mix than a fuel lean formulation. Also, with increased aluminum content in the propellant slag accumulation can be greater in rocket motors. The nanoaluminum considered for this study is Novacentrix 80 nm and has an active aluminum content of 77%. Using only this material in a water mixture limits theoretical Isp to 212 s and the adiabatic flame temperature to about 2800 K. The bimodal powder distribution in the reformulation has an active content of 83.9% increasing theoretical Isp by seven seconds. The proportions of aluminum and water required for the reformulation are derived from balancing the chemical equation for the combustion of water with aluminum,
%!" !"!"

!" +

!!%!" !"!"! !!

!"! !! +

! %!" ! !"!"

!! !

! %!" ! !"!"

!!%!" !"!"! !!

!"! !! +

! %!" ! !"!"

!! .

(1)

Using the molar coefficients in Eq. (1) and defining the propellant equivalence ratio in terms of oxidizer to fuel ratio, a generalized formula for any water and aluminum propellant system is obtained, !
! !

= %!"

! !"!! ! ! !"!"

(2)

B. Reaction Safety Starting from the baseline formulation of ALICE (deionized water and Novacentrix 80nm at = 0.75) chemical additives for propellant safety were considered. Puszynski et al. [10] investigated the reaction kinetics of water and nanoaluminum mixtures with dissolved organic inhibitors. The results show that reactions can be suppressed over several hours. The best inhibitors were ascorbic acid, succinic acid, and oleic acid although the authors also indicated success with ammonium dihydrogen phosphate (ADP). 3 American Institute of Aeronautics and Astronautics

Bulian [11] describes the ageing effects of nanoaluminum in water and uses ADP for water processing nanothermites (nanoaluminum-nanooxide mixtures). He reports that nanoparticles have a large surface charge per unit area, and that hydrophobic material is needed to thinly coat nanoaluminum in water to prevent degredation. This type of nanoparticle surface coating has been shown to be successful in inhibiting nanoaluminum degredation in dilute nanothermite formulations. To investigate the applicability of these findings for ALICE formulations, 3 wt.% of solids ADP was added to baseline ALICE, and the modified propellant was placed in a reaction flask connected to a simple gas analyzing burette. Data collected includes volume of gas evolved over time (assumed to be hydrogen) and propellant temperature. Results are displayed in Fig. 3 and show that the nanoaluminum-water reaction is suppressed after mixing by about 10 hours at a temperature of 30C. Some initial reaction could have occurred during mixing and this is the subject of current research.

H2 Production, mL

40 30 20 10 0 0 2 4 6 8 10

Time, hours

500

Temperature, K

450 400 350 300 0 2 4 6 8 10

Time, hours

Figure 3. Stability of 5 g of ALICE with 3 wt.% ADP. Sample has been mixed at room temperature. C. Propellant Selection Connell et al. [9] investigated bimodal distributions of 80 nm Novacentrix Al and 2 m Valimet Al at = 0.943 and showed that micron addition up to 30-40 wt.% would have little effect on regression rate. This range maintains the more vigorous combustion associated with nanoaluminum but improves theoretical Isp by a few seconds. Following this work, a bimodal aluminum powder distribution was selected consisting of 70 wt.% nano and 30 wt.% micron. Because of the lower specific surface area of the micron particles, there was concern that the propellant might settle and stratify. A chemical additive to increase suspension was sought from the literature. Ivanov et al. [12] as well as Shafirovich et al. [7] reported success using the cross-linking polymer polyacrylamide (poly-A) to gel water with ultrafine aluminum powders. They also claim that ultrafine powders only combust well when the water is gelled. Here, poly-A is included at 1 wt.% of total solids and prepared ahead of time with a portion of the water. The equivalence ratio is selected in order to maximize sea level Isp. Table 1 summarizes the reformulation, which is deemed ALICE II. Table 1. ALICE II reformulation with ! = 0.813. Constituent Novacentrix 80 nm Al Valimet 2 m Al Deionized Water ADP poly-A Amount 70% of Al powder 30% of Al powder o/f = 1.034 3 wt.% 1 wt.%

4 American Institute of Aeronautics and Astronautics

D. Propellant Mixing Any quantity of propellant greater than five grams is mixed in a polypropylene mixing jar on a Resodyn resonating mixer (Resodyn Acoustic Mixer, Inc.). Depending on the investigation, the propellant is either mixed at room temperature or chilled an insulated container is sometimes used and/or short breaks in the mixing are made to allow the propellant to cool. In both cases, deionized water is pre-chilled in a refrigerator and the nanoaluminum is stored under an inert argon environment. One of many challenges with ALICE II is getting repeatable consistency. Larger amounts of material take more effort to mix completely and it is difficult to scale mixing routines. Mixing proceeds in two phases pre collapse and post collapse. At first, water adheres to the nanoaluminum and forms small balls (see Fig. 4b) that eventually bounce and coalesce together into larger balls. With enough time these balls collapse together. The collapse is visible on the LabRAM software as a sudden dip in acceleration, and post collapse mixing simply drives the viscosity of the propellant down and no change in acceleration is observed. An Omega PHH-37 probe shown in Fig. 4a is used to measure the pH of the propellant after it has coalesced together. Prior to addition, the deionized water with ADP has a pH of ~5, and after collapse ALICE II has a pH ~7. The pH can be driven up to ~10 however the time required, to go from a pH of 7 to ~10, depends on the amount of material, mixing routine (frequent breaks or not) and sample temperature. Interestingly, the lowest pH measurements correspond to the highest viscosity versions of ALICE II. Viscosity measurements will likely have a strong correlation with pH and will be thouroghly investigated in the near future. The reason viscosity decreases could be that the nanoaluminum has begun slowly reacting and is no longer gelling the mixture sufficiently. Another consideration is the effect of pH on particle dispersion. Regardless, the change in pH is a sign of premature nanoaluminum reaction and ADP is unable to inhibit this initial reaction during mixing. Possible reasons include mixing stresses removing some of the thin ADP coating and the lack of ADP mobility due to the high viscosity of the mixture. Figure 4 highlights propellant appearance and pH versus mixing.

(a)

(b)

(c)

(d)

Figure 4. ALICE II appearance with pH: a) Omega PHH-37 probe used to measure pH, b) ALICE II near the collapse point, c) ALICE II with pH = 6.78, d) ALICE II with pH 9-10. E. Nanoaluminum Reaction Kinetics According to the Arrhenius equation, used to describe the rate of a chemical reaction, a decrease in temperature will exponentially decrease the reaction rate. This helps to explain why a small (1020C) decrease in temperature can have such a significant effect on propellant mixing. Another way to control the reaction rate is to change the activation energy. Novacentrix 80 nm Al was reacted in varying basic and acidic solutions of ammonium hydroxide and hydrochloric acid at room temperature, and a simple manometer recorded hydrogen production. The results are shown below in Fig. 5 and show that basic solutions tend to accelerate nanoaluminum reactions. A highly acidic solution can also accelerate reactions, but in comparison to the basic solutions this is inconsequential. More importantly, reactions are slowest at a pH near neutral (6.47) and this is where the propellant should be mixed to minimize nanoaluminum degredation with respect to pH. On-going research will investigate ways to control pH, and although ADP provides a level of safety from the fast reaction of the nanoaluminum with water, the best way to inhibit degradation while mixing is still via temperature control and shorter mixing times.

5 American Institute of Aeronautics and Astronautics

2
Moles of Hydrogen Produced

1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 6 12 18 Time, minutes

pH = 12.91 pH = 12.12 pH = 11.53 pH = 6.47 pH = 1.13

24

30

Figure 5. Reaction of Novacentrix 80 nm aluminum in basic and acidic solutions. The same mass of nanoaluminum was used for each test.

III. Motor Tests


A. Igniter The sounding rocket flight of ALICE in 2009 utilized a complete igniter motor [13] in which igniter propellant was separated from the main combustion chamber by a convergent-divergent nozzle. Traditional igniters designed for ammonium perchlorate composite propellants (APCP) do not work because the total energy delivered is too low. Also, the goal of the igniter motor is to elevate chamber pressures in a timely manner since ALICE burns more efficiently at higher pressures [14]. The first attempt to ignite ALICE II in a rocket motor was conducted with a new design. A dual layered center perforated grain was used with composite propellant lining the inside port of the ALICE II grain. This concept was first tested to ignite an aluminized APCP grain and proved very successful. Unfortunately, the dual layered grain could not ignite ALICE II most likely because too much energy from the igniter was directed out the motor. The flaw of this design was the lack of an impinging flame, and a center perforated APCP grain was picked to sit at the head end of the motor. The disadvantages of this option are that the igniter burn out time is dependent on operating pressures and that both propellants need to share the same nozzle. On the other hand, a hot subsonic flame impinging on ALICE II can potentially improve combustion efficiency. The igniter grain is initiated by a Pyrodex pellet used for a muzzle-loading rifle. B. Performance Calculations Because of the unusual configuration of the rocket motor, with the igniter playing a nontrivial role in combustion, performance metrics are selected with care. First, because of the shapes of the pressure histories all integrations are taken over the action time of the motor as defined by Sutton [15] to be the points at 10% peak pressure. The equation used for combustion efficiency defined by Heister [16] is !! =
!! !! !! ! !" !!"#! ! !

(3)

where c* is the theoretical value from thermochemical equilibrium calculations. This presents a problem because although values are known individually for ALICE II and the igniter propellant, during the motor firing it is an unknown variable. To remedy this situation upper and lower bounds were calculated by observing the inequality ! !"#!$%& > ! !"#!$%%&'( and also by modifying the mass to include the igniter grain to obtain 6 American Institute of Aeronautics and Astronautics

!" =

!! !! !! ! !" !! !!"!#$ ! ! !"#!$%&

!! ! !! !! !" = !!"!#$ ! ! !"#!$%%&'(

!" .

(4)

The actual combustion efficiency does not have to lie in the center of this interval. If the propellant dominates motor operations then the upper bound for combustion efficiency is approached, and conversely if the igniter is supporting the motor then the lower bound is approached. Specific impulse follows a more standard definition that includes the weight of the igniter as part of the propellant !"# =
!! ! ! !"

!!"!#$ !!

(5)

Specific impulse is a performance measure of both the propellant and the rocket nozzle combined. Accordingly, theoretical Isp values can be attained only if the nozzle is perfectly contoured and expanded. For all tests the nozzle has an inlet angle of 30.8 and an exit cone angle of 14.0. These angles are picked to fit the available space and create an expansion ratio of 7.8, which is slightly under expanded. Finally, the thrust and pressure of the rocket motor are related together by a thrust coefficient, which can be defined with respect to integral calculations !! =
!! ! ! !" ! !! ! ! !! !"

(6)

Beyond these calculations, percent mass remaining is also computed by collecting as much of the motor slag as possible and attributing it entirely to the ALICE II grain. From all of these measurements a complete picture of the motor operational performance is obtained. C. Uncertainty Calculations The pressure transducer and load cell used for testing the rocket motors are known sources of error in these calculations. The corresponding measurement uncertainties are 52 kPa for pressure measurements (0.25% of full scale of Druck PMP-1200-1260 pressure transducer) and 6.7 N for force measurements (assumed 0.3% of full scale of Interface SSM-AJ-500 load cell). Uncertainties are added in quadrature according to the formula ! =
! !!!

!!

!/! !" ! !!!

(7)

where R is a function of the variables !! . The uncertainty in R (denoted !) is a function of the individual variable uncertainties, !! . Uncertainties for total impulse, specific impulse, and thrust coefficient are calculated from Eq. (7) and are summarized below in Eq. (8) through (11):
!! ! !" !

= !! ! ,

(8)

!! !! !" !

= !! !! ,
!! !

(9)

!"# =

!!"!#$ !!

(10)

!! = !!

!! ! !! ! ! !"

!! !! !! ! !! !"

(11)

7 American Institute of Aeronautics and Astronautics

C. Results Four motor tests were conducted with varying pH. The propellants used for tests A and B had a pH between 9-10 and those for tests C and D had a pH between 7-8. Pressure and thrust histories are shown in Fig. 6 and 7, and test details including grain geometries and propellant characteristics are organized below in Tables 2 through 4. These motors had maximum pressure levels between 8.9 and 19.5 MPa and thrust levels between 414 and 921 N depending on pH. There is a small amount of variation between tests because propellant grains are hand made; the igniter ports are not perfectly centered. Total mass of propellant and igniter is within 1% (6g) for all tests. Another source of variation is the age of the ALICE II propellant grain. All grains were aged 1 day except for test B that was allowed to age 18 days. Tests A and B showed similar delivered impulse but test B had more remaining mass. Ongoing research is aimed at investigating the effects of grain aging. Propellant with pH values near neutral showed an approximate 30% increase in total impulse and specific impulse. Regarding combustion efficiency, the lowest values are near 50% (for the LB of the basic tests) and the highest are near 80% (for the UB of the neutral tests). A combustion efficiency of 80% is higher than the average combustion efficiency of baseline ALICE at 50% [13], but still leaves room for improvement. Also, because of the wide variation between the basic and neutral curves, it is likely that the logarithmic burn rate parameters are different depending on pH. The shorter burn times of the neutral grains reiterate this point. Error propagation was the greatest through load cell measurements. Uncertainty in pressure impulse was less than 1% and uncertainty in specific impulse was less than 2.6% of calculated values. Observed offsets were on this same orders of magnitude. The load cell is connected between a mount and a low friction rail sled, and the sled can seize up and offset measurements if not adjusted carefully. Despite these errors, tests A through D are consistent with themselves and the fact that thrust coefficients are similar attests to the repeatability of the tests.

20 18

Chamber Pressure, MPa

16 14 12 10 8 6 4 2 0 0.4 0.2 0 0.2 0.4 0.6

A B C D

pH 9 10 pH 9 10 pH 7.72 pH 7.03

Time, sec

0.8

Figure 6. Pressure histories for ALICE II motor tests A through D. A 0.635 cm diameter conical graphite nozzle throat was used for all tests. Curves are offset so that time 0 corresponds to ignition of ALICE II.

8 American Institute of Aeronautics and Astronautics

1000 900 800 700

A B C D

pH 9 10 pH 9 10 pH 7.72 pH 7.03

Thrust, N

600 500 400 300 200 100 0 0.4 0.2 0 0.2 0.4 0.6 0.8 1

Time, sec

Figure 7. Thrust curves for ALICE II motor tests A through D. A 0.635 cm diameter conical graphite nozzle throat was used for all tests. Curves are offset so that time 0 corresponds to ignition of ALICE II.

Table 2. Results from static motor tests. Test A B C D pH 9-10 9-10 7.72 7.03 Pmax ( MPa ) 9.86 8.89 16.77 19.51 Fmax (N) 489 414 778 921 tb (s) 0.75 0.73 0.55 0.54 Impulse ( MPa-s ) 4.136 .039 4.431 .038 5.223 .029 5.350 .027 Impulse ( N-s ) 191 5 198 5 246 4 251 4 Isp (s) 120.2 3 124.1 3 149.7 2 153.5 2 cf 1.41 .04 1.46 .04 1.49 .02 1.49 .02 C* LB 52.53% 55.94% 63.96% 65.59% C* UB 65.13% 69.36% 79.30% 81.33%

Table 3. ALICE II grain details. Test A B C D Mass (g) 117.07 117.54 118.65 118.02 Percent Mass Remaining 21.40% 32.20% 18.68% 16.89% Grain Age ( days ) 1 18 1 1 Theoretical c* ( m/s ) 1242.8 1242.8 1242.8 1242.8 OD ( cm ) 3.18 3.18 3.18 3.18 ID ( cm ) 1.27 1.27 1.27 1.27 L ( cm ) 12.55 12.83 12.68 12.60

Table 3. Igniter grain details. Test A B C D Mass (g) 44.75 45.27 49.55 48.92 Theoretical c* ( m/s ) 1541 1541 1541 1541 OD ( cm ) 2.54 2.54 2.54 2.54 ID ( cm ) 1.31 1.27 1.27 1.27 L ( cm ) 6.94 7.06 7.92 7.94 Port Offset ( mm ) 0.8 0.3 0.6 1.5

9 American Institute of Aeronautics and Astronautics

IV. Conclusions
A reformulated aluminum and frozen water solid rocket propellant was selected based on theoretical optimization, literature, and previous studies. The reformulation was tailored with a bimodal Al particle distribution of 70 wt.% Novacentrix nAl and 30 wt.% Valimet 2 m at an equivalence ratio optimized for sea level specific impulse. While mixing the propellant, pH increases because of the slow reaction between nanoaluminum and water. Ammonium dihydrogen phosphate is unable to inhibit this reaction but does add a level of safety to the propellant otherwise. Its reduced effectiveness during mixing could be due to the stress of mixing removing some of the coating and exposing nanoaluminum to water. This problem might be compounded by the lack of ammonium dihydrogen phosphate mobility in the high-viscosity propellant. The slowest nanoaluminum reactions occur near a neutral pH and temperature remains the best way to prevent propellant degredation while mixing. Regarding rocket motor performance, propellant with pH values near neutral showed an approximate 30% increase in total impulse and specific impulse over more basic propellant. Future research will investigate ways to manufacture and test propellants of different pH and also focus on eliminating any premature aluminum reactions in order to achieve greater combustion efficiencies. It is apparent that a tailored version of nanoaluminum would be beneficial for inhibiting reactions due to propellant manufacturing and further increasing performance.

Acknowledgments
The authors would like to thank Kevin Zaseck and Robert Reeves for conducting the study on nanoaluminum reaction rates in basic and acidic soultions. The authors would also like to thank Dr. Mitat Birkan of the Air Force Office of Scientific Research and NASA under contract numbers FA9550-09-1-0073 and FA9550-07-1-0582, and finally Mr. Travis Sippel, Mr. Steven Shark, and Mr. Chris Zaseck for assistance in propellant mixing and rocket motor hardware development and set up.

References
Linne, D. L., and Meyer, M. L., A Compilation of Lunar and Mars Exploration Strategies Utilizing Indigenous Propellants, NASA TM 105262, 1992. 2 Schultz, P. H., Hermalyn, B., Colaprete, A., Ennico, K., Shirley, M., and Marshall, W. S., The LCROSS Cratering Experiment, Science Magazine, Vol. 330, No. 6003, 2010, pp. 468-472. 3 Byrne, S., Dundas, C. M., Kennedy, M. R., Mellon, M. T., McEwen, A. S., Cull, S. C., Daubar, I. J., Shean, D. E., Seelos, K. D., Murchie, S. L., Cantor, B. A., Arvidson, R. E., Edgett, K. S., Reufer, A., Thomas, N., Harrison, T. N., Posiolova, L. V., and Seelos, F. P., Distribution of Mid-Latitude Ground Ice on Mars from New Impact Craters, Science Magazine, Vol. 325, No. 5948, 2009, pp. 1674-1676. 4 Wood, T. D., Pfeil, M. A., Pourpoint, T. D. , Tsohas, J., Son, S. F., Connell, T. L., Risha, G. A., and Yetter, R. A., Feasibility Study and Demonstration of an Aluminum and Ice Solid Propellant, 45th Annual AIAA JPC Conference, 2009, AIAA 2009-4890. 5 Gordon, S., and McBride, B. J., Computer Program for Calculation of Complex Chemical Equilibrium Compositions and Applications, Part II. Users Manual and Program Description, NASA Ref. Pub. 1311, June 1996. 6 Sippel, T. R., Characterization of Aluminum and Ice Solid Propellants, Masters Thesis, Mechanical Engineering Dept., Purdue University, West Lafayette, IN, 2010. 7 Shafirovich, E., Diakov, V., and Varma, A., Combustion of novel chemical mixtures for hydrogen generation, Journal of Combustion and Flame, Vol. 144, 2006, pp. 415-418. 8 Wood, T. D., Feasibility Study and Demonstration of an Aluminum and Ice Solid Propellant, Masters Thesis, Mechanical Engineering Dept., Purdue University, West Lafayette, IN, 2010. 9 Connell, T. L., Risha, G. A., Yetter, R. A., Yang, V., and Son, S. F., Combustion of Bimodal Al and Ice Mixtures, 8th International Symposium on Special Topics in Chemical Propulsion, Cape Town, South Africa, November 2009. 10 Puszynski, J. A., and Groven, L. J., Formation of Nanosized Aluminum and Its Applications in Condensed Phase Reactions, Inorganic Nanoparticles; Synthesis, Applications, and Perspectives, CRC Press, 2010, Chap. 6. 11 Bulian, C., Ignition and Combustion Characteristics of Nanoenergetic Materials, Ph.D. Thesis, Nanoscience and Nanoengineering Dept., South Dakota School of Mines and Technology, Rapid City, SD, 2010. 12 Ivanov, V. G., Leonov, S. N., Savinov, G. L., Gavrilyuk, O. V., & Glazkov, O. V., Combustion of mixtures of ultradisperse aluminum and gel-like water, Combustion, Explosion, and Shock Waves, Vol. 30, No. 4, 1994. 13 Pourpoint, T. D., Sippel, T. R., Zaseck, C., Wood, T. D., Risha, G. A., and Yetter, R. A., Detailed Characterization of Al/Ice Propellants, 46th Annual AIAA JPC Conference, 2010, AIAA 2010-6905. 14 Ivanov, V. G., Gavrilyuk, O. V., Glazkov, O. V., and Safronov, M. N., Specific Features of the Reaction between Ultrafine Aluminum and Water in a Combustion Regime, Combustion, Explosion, and Shock Waves, Vol. 36, No. 2, 2000, pp. 213-219. 15 Sutton, G. P., and Biblarz, O., Rocket Propulsion Elements, 8th ed., John Wiley & Sons, Inc., New Jersey, 2010, Chap. 12. 16 Heister, S. D., Solid Rocket Motors, Space Propulsion Analysis and Design, edited by R. W. Humble, G. N. Henry, and W. J. Larson, 1st ed., McGraw-Hill, New York, 1995, Chap. 6.
1

10 American Institute of Aeronautics and Astronautics

You might also like