You are on page 1of 10

JOURNAL OF PROPULSION AND POWER

Vol. 33, No. 5, September–October 2017

Viscosity and Regression Rate of Liquefying Hybrid


Rocket Fuels

M. Kobald,∗ C. Schmierer,† H. K. Ciezki,‡ and S. Schlechtriem§


DLR, German Aerospace Center, 74239 Hardthausen, Germany
and
E. Toson¶ and L. T. De Luca**
Polytechnic University of Milan, 20156 Milan, Italy
DOI: 10.2514/1.B36207
The combustion behavior of paraffin-based hybrid rocket fuels with gaseous oxygen as an oxidizer has been
analyzed in detail. Regression rate tests have been done in a two-dimensional radial microburner at the DLR, German
Aerospace Center and at the Space Propulsion Laboratory. Fuel samples have been characterized by viscosity
measurements, tensile tests, and a differential scanning calorimeter. Tensile tests showed significant improvement in
maximum stress and elongation when polymers in low concentration were added to the paraffin samples. The values
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

of the liquid fuel viscosities differed significantly between the selected fuels. This affected the droplet entrainment
process during combustion and the regression rates of the fuels. The entrainment and regression rate increased for the
decreasing fuel liquid layer viscosity. An exponential relation has been found between the liquid fuel layer viscosity
and the regression rate, which can be used to predict the regression rate of new liquefying fuels by measuring their
viscosity.

Nomenclature capability are additional advantages. Less piping and valves due to
b, m = curve fit parameters only one liquid component introduce less complexity and reduced
c = characteristic velocity, m∕s costs compared to liquid rocket engines. Applications, especially in
Gox = oxidizer mass flux, kg∕m2 ⋅ s space tourism (like SpaceShipOne some years ago), clearly show the
h = melt layer thickness, m potential of hybrid rocket engines. One of the disadvantages might be
m_ ox = oxidizer mass flow rate, kg∕s the use of polymeric fuels. These fuels, like hydroxyl-terminated
p = pressure, bar polybutadiene (HTPB) or high-density polyethylene (HDPE) show
r_ = regression rate, mm∕s relatively low regression rates that result in the necessity for multiport
μ = dynamic viscosity, Pa ⋅ s fuel grains for high-thrust applications. This multiport design
σ = surface tension, N∕m increases the residual mass of unburned fuel, and thereby decreases
the delivered specific impulse. Instabilities are also more likely with
Subscripts this type of design.
Carrick and Larson evaluated cryogenic solid hybrid rocket fuels
ent = entrainment [2]. They used cryogenic solid n-pentane and measured regression
l = liquid rates 5–10 times higher than those of polymeric hybrid fuels.
Following these studies, tests have been done with hydrocarbons with
longer chains that are solid at ambient temperature [3,4]. These fuels
are paraffin-based and show a three to five times higher regression
I. Introduction
rate at similar mass fluxes compared to polymers. This is achieved by
H YBRID rocket engines do have advantages compared to
classical solid- or liquid-propellant rocket engines. A high
number of publications during the last years shows the interest in
a different combustion mechanism. Paraffin fuels form a liquid layer
on the fuel surface during the combustion [3]. It is expected that the
low viscosity and low surface tension of the liquid fuel enable an
hybrid rocket engines [1]. Considering storage and handling, they additional mass transfer by entrainment of liquid droplets. The gas
have zero Trinitrotoluene equivalent compared to solid propellants. flow over the surface induces liquid layer instabilities that produce
This is an important aspect in determining the safety and total cost of the droplet entrainment [5,6]; see also Figs. 1 and 2. Optical results of
an engine. Controllable thrust, shutoff capability, and restart the entrainment process of such low-viscosity liquefying fuels are
shown in [7,8].
Presented as Paper 2014-3646 at the 50th AIAA/ASME/SAE/ASEE Joint To account for the increased regression rate by entrainment, the
Propulsion Conference, Cleveland, OH, 28–30 July 2014; received 11 classical hybrid combustion theory needs to be modified to consider
February 2016; revision received 25 November 2016; accepted for publication the reduced heating of the entrained fuel, the reduced blocking effect
14 December 2016; published online Open Access 16 March 2017. Copyright due to two-phase flow, and the increased heat transfer due to the
© 2016 by the Authors. Published by the American Institute of Aeronautics
and Astronautics, Inc., with permission. All requests for copying and
increased surface roughness. Much more fuel is transported into the
permission to reprint should be submitted to CCC at www.copyright.com; flame zone by the entrainment before being totally vaporized. Scale-
employ the ISSN 0748-4658 (print) or 1533-3876 (online) to initiate your up tests have confirmed that the theory is applicable also for large
request. See also AIAA Rights and Permissions www.aiaa.org/randp. engines [9]. Karabeyoglu et al. [3] assumed an empirical formula
*Group Leader, Institute of Space Propulsion, Propellants Department. from the literature to quantify the value of entrainment mass transfer
Member AIAA. as shown in Fig. 1. Here, the upper part of the equation contains the

Ph.D. Candidate, Institute of Space Propulsion, Propellants Department. operational parameters of the combustion: dynamic pressure pdyn and

Head, Institute of Space Propulsion, Propellants Department. Senior (indirectly) the oxidizer mass flux. The lower part contains the fuel
Member AIAA.
§
Director, Institute of Space Propulsion. Member AIAA.
properties: the surface tension σ and the melt layer viscosity μl of the

Ph.D. Candidate, Aerospace Engineering Department, Via Giuseppe La fuel. Literature values for the exponents are given in [3,10,11]. The
Masa 34. exponents α and β are believed to be between one and two, whereas γ
**Professor, Aerospace Engineering Department, Via Giuseppe La Masa and π should be almost one. Karabeyoglu et al. [3] also stated that the
34. Associate Fellow AIAA. melt layer viscosity should have a greater influence on the regression
1245
1246 KOBALD ET AL.

Fig. 1 Liquefying fuel combustion theory, taken from the work of Karabeyoglu et al. [6].

been chosen due to their different viscosity data. Detailed laboratory


experiments have been performed before to measure the viscosity and
surface tension of the different fuels [14]. These are the two material
parameters that are expected to have the biggest influence on the
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

entrainment mass flow, according to Fig. 1. All samples for the


ballistic tests have been blackened by additives during fabrication to
Fig. 2 Liquefying fuel combustion image, taken from the work of limit radiation effects to the fuel surface during combustion.
Kobald et al. [8] (oxidizer mass flow from left to right). Generally, the amount of blackening additive was less than 2%, and
therefore has negligible impact on the performance. Three different
additives have been chosen in order to investigate and improve the
rate than the surface tension, meaning γ > π. Not many results are mechanical properties of the paraffin samples. Stearic acid (SA) was
published on the influence of additives on the regression rate, used in a combination of paraffin 6003 and 6805. But, the change in
viscosity, and mechanical properties of paraffin-based fuels. viscosity and regression rate is pretty small compared to the other
Nakagawa and Hikone did an investigation on the dependence of additives. Then, a nanoclay material from manufacturer Byk is used
the regression rate on the fuel viscosity [12]. They investigated with an average particle size of 10 μm. Research by Wang et al. [15]
paraffin and gaseous oxygen (GOX) as propellants in a two- has shown that even small amounts of 1–2% nanoclay added to
dimensional slab burner with windows in the side for optical access paraffin increase the maximum tensile stress up to 100%. The strain at
[13]. Pure paraffin and paraffin blends with different viscosities were rupture could be improved up to 400%. It is seen that the ambient
used. The tests were run at atmospheric pressure where they could temperature has significant influence on the results. This is because a
show that droplets were generated during combustion and entrained
paraffin with a lower melting temperature than our samples was used.
in the flow. They assumed the heat transfer coefficient to be
The last additive investigated was a commonly available polymer
proportional to μ−1∕6 . The mass flux of the tests was very low. The
with a melting temperature range similar to our paraffin samples.
viscosities, which were compared with the tests, were measured only
at one temperature of 120°C. The average melt layer temperature was
expected to be much higher [4]. B. Testbench
The experimental tests have been performed at the Institute of
Space Propulsion at DLR Lampoldshausen at testbench M11.3 [8].
II. Test Setup Ignition has been done with a small pyrotechnical charge inside the
A. Paraffin Fuels and Additives chamber. A test sequence is programmed before the test and is run
Combined experimental activities have been done at the DLR, automatically by the testbench control system. All tests have been
German Aerospace Center’s (DLR’s) Institute of Space Propulsion in done with the same timing settings shown in Table 2. Each test
Lampoldshausen and at the Space Propulsion Laboratory (SPLab) of duration for this test campaign was 3 s. For redundancy, a separate
Politecnico di Milano. Four different baseline paraffin waxes are measurement system has been used for data acquisition during the
being investigated as fuels in this work. They are used in pure form as tests. Several pressure and temperature measurement positions in the
well as with different additives to modify mechanical and rheological test facility have been used to control the test conditions.
properties. Their properties are reported from the manufacturer in The oxidizer mass flow rate is set by a pressure regulator and a flow
Table 1. Type 6003 is a pure paraffin wax, and type 0907 is a control valve. It is measured with a Coriolis flow meter with an
microcrystalline wax that is used (for example) in hot glues. Type accuracy better than 0.35% and a repeatability better than 0.2% of the
6805 has the same application, but it is a paraffin wax. The last, type flow rate. A typical mass flow rate measurement of the oxidizer is
1276, is used for coatings, gloss, and sealing. Coatings manufactured shown in Fig. 3a. The measured oxidizer mass flow is shown as
with these waxes exhibit higher strength, and hence abrasion m_ ox;cor . The mass flow seems to increase very slowly and is unsteady.
resistance as well as an improved gloss impression. Its special
formulation is based on waxes and a variety of different additives
according to the manufacturer, Sasol Wax. The paraffin waxes have Table 2 Automatic test sequence
Time, s Action
Table 1 Wax properties given by manufacturer T0 − 30 Start of sequence
T0 − 29 Commanding control valve
Kinematic T0 − 15 Set dome regulator pressures
Sasol Congealing Penetration at viscosity at T0 − 03 Start of data acquisition
wax point, °C Oil content, % 25°C, 1∕10 mm 100°C, mm2 ∕s T0 − 0.1 Start electrical igniter
6003 60–62 0–0.5 17–20 — — T0 Open oxidizer main valve
6805 66–70 0–1 16–20 6–8 T0  2 Close oxidizer main valve
0907 83–94 0–1 4–10 14–18 T0  12 Start nitrogen purge
1276 64–68 — — 8–13 880–920 T0  22 End of sequence
KOBALD ET AL. 1247

Fig. 3 Typical measurement data: example of test 61.

This is due to the position of the Coriolis sensor connected to the 6805/ 2% Nanoclay 6805/ 5% Polymer/ 2% Nanoclay
combustion chamber over some distance, upstream in the testbench. 6805/ 5% Polymer
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

6805 /10%Polymer

Dynamic viscosity [Pa·s]


Therefore, the mass flow is not measured synchronously with the 0 0907 1276
10
measured combustion chamber pressure, due to transient effects in
the feedlines. But, the total oxidizer mass, which is important for the
regression rate calculations, remains valid. The oxidizer mass flow
has also been calculated with the measured pressure difference across −1
10
the injector, shown as m _ ox;calc in Fig. 3a. It is seen that the mass flow
increases rapidly after ignition and is much more steady than
indicated by the Coriolis measurement. The combustion chamber
−2
pressure data show smooth and stable combustion with almost no 10
oscillations; see Fig. 3b.
100 10 2 10 3
Shear rate [1/s]
III. Experiments Fig. 5 Viscosity of paraffin fuels at a constant temperature of 120°C.
A. Viscosity Measurements
The viscosity data of the pure paraffin samples investigated here
have already been presented in detail in previous publications paraffin waxes with carbon numbers of about 30, a critical pressure of
[14,16]. There are also different publications available that give about 6.5 bar and a critical temperature of approximately 840 K is
analytical formulas for material properties like surface tension or predicted [19]. In these conditions, no clear distinction can be made
viscosity of paraffin fuels: for example, from Marano and Holder between the liquid and the gas phases of the fuel. The surface tension
[17,18]. But, these are often only valid for straight paraffin waxes and the heat of vaporization are approaching zero. If we insert into the
with distinct carbon numbers. These cannot be applied for some equation in Fig. 1a value close to zero for the surface tension, this will
of our pure samples, as well as the mixtures with additives. _ ent. Supercritical conditions are different
result in a value of infinity for m
Measurements of the viscosity of the samples with additives at because a small pyrolysis layer above the liquid layer is assumed.
different temperatures are shown in Fig. 4. The viscosity increases Karabeyoglu et al. [6] later extended their entrainment relation in Fig. 1.
significantly with increasing polymer addition. The effect on the They predicted the entrainment formula for sub- and supercritical
viscosity when adding nanoclay to paraffin and polymer samples is operations based on a relation by Gater and L'Ecuyer [10], without
negligible due to the generally high viscosity of the polymer. Figure 5 accounting for the surface tension component in the denominator.
shows the viscosity at constant temperature. The pure 6805 sample
with nanoclay particles shows non-Newtonian shear thinning B. Tensile Tests
behavior. The other samples behave as Newtonian fluids. Tensile tests have been performed in order to examine the
One of the difficulties to find a general entrainment formula for
mechanical properties of the fuel mixtures. Standardized fuel
liquefying hybrids is that most experiments are performed well above
samples have been produced to perform the tests; 10 samples of each
the supercritical temperature and pressure of the paraffin waxes. For
type have been tested. The results are shown in Fig. 6. Only the
0
10 6805 / 2% Nanoclay 3
6805 / 5% Polymer
6805 / 5% Polymer / 2% Nanoclay
Dynamic viscosity [Pa·s]

6805 / 10% Polymer 2.5


−1
10
tensile stress [−]

2
Normalized

1.5
−2
10
1 6805 / 10% SA
6805 / 2% Nanoclay
0.5 6805 / 5% Polymer
−3 6805 / 10% Polymer
10
80 100 120 140 160 180 200 6805 / 5% Polymer 2% Nanoclay
0
Temperature [°C] 0 0.5 1 1.5 2
Fig. 4 Viscosity of paraffin 6805 with different additives at a constant Tensile strain [%]
shear rate. Fig. 6 Results of tensile tests.
1248 KOBALD ET AL.

samples that broke uniformly in the foreseen middle part of the fuel 0
have been used for data analysis. The shown data are averaged from at

Specific heat flow [mW/g]


least three test samples. An extensometer was used to measure the
elongation of the sample. The maximum tensile strength values were −0.5
normalized versus the baseline paraffin sample 6805  10% SA. It
was seen that the addition of 2% nanoclay doubled the value of the
−1
elongation. The maximum tensile strength was not improved
significantly. A 5% polymer addition yielded an almost two-time
increase in strength and three-time increase in elongation. The
−1.5
biggest increase was achieved with a 10% polymer addition. The 89.35% Micro + 10.15% SA + 0.5% Dye (19 bar)
strength was increased to three times when compared to the pure 88% Micro + 10% SA + 2% G (16 bar)
88% Micro + 10% SA + 2% CB (19 bar)
paraffin sample, and maximum elongation achieved a five times −2
higher value. Good results were also achieved for 5% polymer and 40 0
60 20
80 100
2% nanoclay. As shown previously, the addition of additives changed Temperature [°C]
the viscosity of the propellant mixture significantly. This then Fig. 8 DSC tests: melting behavior of microparaffin-based (Micro)
influenced not only the mechanical properties but also the regression mixtures (Sasol 0907).
rate. Results by Maruyama et al. [13] with a 20% ethylene vinyl
acetate addition to paraffin showed an increase of 1.6 in maximum
2. Regression Rate Comparison
strength and 2.2 in maximum strain.
Experimental activities for regression rate measurements have
been performed at the SPLab of Politecnico di Milano using a two-
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

C. Black Pigment Analysis


dimensional (2-D) radial microburner [20]. All paraffin-based
Further tests at the SPLab have been performed on the effects of mixtures tests were performed at the reference conditions of a
different methods of coloring the fuel samples. To avoid possible 210 Nl∕ min oxidizer volume flow rate. Carbon black and dye data
sloughing of the fuel, most of the radiation from the flame has to be tests performed at a 19 bar combustion chamber pressure were also
absorbed at the fuel surface. Thus, to increase the fuel absorptivity of compared with previous data tests performed at 16 bar on mixtures
radiation, a percentage of graphite is commonly added to paraffin. containing the same paraffin and graphite [21]. GOX was used as
Due to its high density at solid state (reference density is oxidizer. The oxidizer flow rate and pressure were maintained
2250 kg∕m3 ), graphite tends to settle down, thus making difficult the constant during the combustion, thus granting quasi-steady operating
achievement of a homogeneous mixture during the casting process. conditions. The results of the ballistic characterization are shown in
Two other pigments suitable for paraffin-based fuels have been Figs. 9 and 10. Ensemble regression rate curves were obtained by
tested. Two percent of graphite synthetic powder (less than 20 μm) is interpolation of three to four tests for each tested formulation. For
substituted with 2% of ultrafine carbon black (CB) powder PBk7 and macroparaffin-based mixtures, considering an oxidizer mass flux
with 0.5% of a concentrated liquid black candle dye, which contains range between 60 and 200 kg∕m2 s, the highest mean regression rate
pigments with oil-soluble solvents and without mineral oil. The lower value was achieved by the mixture containing dye, followed by
dye content in the fuel mixture is due to achievement of a
homogeneous black color with 0.5% of the dye. Good homogeneity
3.5
is obtained for the mixture with pigment, whereas the graphite and
PBk7 sedimented during solidification of the paraffin. 3
Regression rate [mm/s]

1. Differential Scanning Calorimeter Comparison 2.5


A differential scanning calorimeter (DSC) was used in order to
identify the thermal behavior of the fuel mixtures. Two cycles were 2
performed at a 10°C∕ min heating velocity with a specimen mass
lower than 5 mg. The test temperature range varied from 0 up to 1.5
200°C. Figures 7 and 8 show, respectively, the melting behavior of
macroparaffin-based (Sasol 6003) mixtures and microparaffin-based 1 88% Macro + 10% SA + 2% CB (19 bar)
89.35% Macro + 10.15% SA + 0.5% Dye (19 bar)
(Sasol 0907) ones, focusing on the temperature range of 0–100°C. No 88% Macro + 10% SA + 2% G (16 bar)
significant differences were appreciated for macroparaffin-based 0.5
50 100 150 200 250
fuels, whereas within the microparaffin ones, the graphite baseline
Oxidizer mass flux [kg/(m2s)]
showed a sensible lower specimen absorbed heat in the considered
Fig. 9 Regression rate ensemble curves for macroparaffin-based
interval.
mixtures (Sasol 6003).

3.5
88% Micro + 10% SA + 2% CB (19 bar)
0 3 89.35% Micro + 10.15% SA + 0.5% Dye (19 bar)
88% Micro + 10% SA + 2% G (16 bar)
Regression rate [mm/s]
Specific heat flow [mW/g]

2.5
−1
2

1.5
−2
88% Macro + 10% SA + 2% CB (19 bar)
1
89.35% Macro + 10.15% SA + 0.5% Dye (19 bar)
88% Macro + 10% SA + 2% G (16 bar)
−3 0.5
0 20 40 60 80 100 50 100 150 200 250 300
Temperature [°C] Oxidizer mass flux [kg/(m2s)]
Fig. 7 DSC tests: melting behavior of macroparaffin-based (Macro) Fig. 10 Regression rate ensemble curves for microparaffin-based
mixtures (Sasol 6003). mixtures (Sasol 0907).
KOBALD ET AL. 1249

Table 3 Fuel viscosities and ballistic test results


Fuel combination Normalized viscosity; μ∕μref r_∕_rHTPB at Gox  50, kg∕m2 s r_∕_rHTPB at Gox  100, kg∕m2 s Curve fit R2 No. of tests
6003  10% SA 1 6.86 7.20 0.920 9
6805  10% SA 1.11 6.96 7.19 0.916 12
0907 1.125 6.44 6.53 0.650 10
6805  5% polymer 5.69 5.10 4.96 0.973 10
6805  10% polymer 30.52 3.52 3.28 0.877 6
6805  2% nanoclay 1 8.04 6.71 0.735 6
1276 60.72 3.32 2.80 0.605 14
HDPE 3125 1.18 1.09 0.965 3
SP1-a (literature) — — 3.07 3.04 — — — —
HTPB (literature) — — 1.0 1.0 — — — —

graphite and carbon black; see Fig. 9. This was associated with a values. But, it has to be mentioned that the number of tests for this fuel
marked increase on GOX sensitivity. For microparaffin-based formulation is not enough for a reliable curve fit due to a higher
mixtures in Fig. 10, high data scattering was observed for ensemble scatter when compared to the other samples. The mixture with the
regression rate curves on dye and carbon black mixtures; see polymer shows a low variance from the curve fit. The resulting curve
associated error bars. But, as a general trend, higher GOX sensitivity of the 10% polymer mixture lies close to the literature data curve of
for the mixture containing dye was also remarkable in this case. SP-1a.
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

The results of Table 3 are visualized in Figs. 13 and 14. The


D. Ballistic Tests at the DLR, German Aerospace Center viscosity is normalized with a reference value at 120°C, and the
Currently, 75 tests have been done at DLR testbench M11.3 in regression rate is normalized with the literature value of HTPB.
order to characterize the performance of the different paraffin-based Figure 13 shows the values of the different paraffin fuel samples in
fuel combinations described herein. A 2-D radial microburner was detail, and Fig. 14 includes the results of HDPE fuel. The HDPE
used. Cylindrical fuel samples have been tested with a length of viscosity is taken from [4] at an average between the melting and the
120 mm, a 32 mm outer diameter, and a variable inner diameter to boiling temperatures. It is also expected from the literature that a
realize different mass flux values. The oxidizer mass flow was set liquid melt layer is formed during the combustion of HDPE. But, the
between 15 and 30 g∕s. Table 3 shows the tested fuel combinations, viscosity of HDPE is too high to produce unstable waves in the liquid
the number of tests, the normalized viscosity, the normalized film layer, and no droplet entrainment is generated. Therefore, there
regression rates, and the coefficients of the curve fits. Each test time is some similarity in the combustion mechanism, but the actual
was programmed for 2 s, whereas the actual time of combustion was regression rate through droplet entrainment is missing for HDPE.
However, it is interesting to note that HDPE can still be included in
about 3 s due to residual oxygen in the pipe after the main shutoff
the following relation. And, the deviation is quite small to the data for
valve. This effect has been included in the regression rate calculation.
liquefying fuels with much lower viscosities; see Table 4 and Figs. 13
Tests have been done at representative settings of chamber pressure,
and 14. The regression rate data points of the fuels in Figs. 13 and 14
mass flux, and mixture ratio because they would also be expected in
are shown at an oxidizer mass flux of 50 and 100 kg∕m2 s. The
operational systems; see also Fig. 3b.

1. Regression Rate Measurements O2/SP−1a, Literature

An overview about the averaged regression rates of all tests with 4 O2/HTPB, Literature
Regression rate [mm/s]

pure fuel samples is shown in Fig. 11. The data have been diameter O2/6805−10%SA

averaged [22]. Here, it is seen that paraffins 6003 and 6805, each with 3 O2/6805−05% Polymer

10% SA, show the highest regression rates. They also have the lowest O2/6805−10% Polymer
O2/6805−2% Nanoclay
viscosity. Furthermore, the regression rates are decreasing as the 2
6805/10%SA
viscosity values of the paraffin samples are increasing. Sample 1276
6805/05% Polymer
has a significantly lower regression rate than the other samples, as 1
6805/10% Polymer
expected from the higher viscosity. But, it still burns about three times
6805/2% Nanoclay
faster compared to HDPE measurements or literature values 0
for HTPB. 0 50 100
Figure 12 shows the results for the additives in combination with Oxidizer mass flux [kg/(m2s)]
6805. The mixture with nanoclay shows the highest regression rate Fig. 12 Regression rate results of paraffin fuel with additives.

O2/SP−1a, Literature 1
10 6003/10% SA
O2/HTPB, Literature
O2/0907 fit 6805/10% SA
4 0907
Regression rate [mm/s]

O2/6003−10%SA fit
regression rate [−]

O2/6805−10%SA fit 6805/5% Polymer


6805/10% Polymer
Normalized

O2/1276 fit
3 O2/HDPE fit
1276
6003/10% SA
0907
6805/10% SA
6003/10%SA
0907
2 6805/10%SA 6805/5% Polymer
1276 6805/10% Polymer
HDPE 1276
1 Curve fit for G=50
0
10 Curve fit for G=100

100 101 102


0 Normalized viscosity [−]
0 50 100 150 200 250
Fig. 13 Normalized regression rate data versus viscosity, without
Oxidizer mass flux [kg/(m2s)] HDPE fuel (red symbols indicate an oxidizer mass flux of 50 kg∕m2 s, and
Fig. 11 Regression rate results of paraffin fuel samples. blue symbols indicate 100 kg∕m2 s).
1250 KOBALD ET AL.

1
10 6003/10% SA

Regression rate [mm/s]


6805/10% SA 4
0907
6805/5% Polymer
regression rate [−]

6805/10% Polymer 3
Normalized

1276
HDPE
6003/10% SA 2
6805/10% SA
0907 6805/ 10% SA
6805/5% Polymer 1 6805/ 5% Polymer
6805/10% Polymer 6805/ 10% Polymer
1276 6805/ 2% Nanoclay
HDPE 0
0
10 Curve fit for G=50 0 20 40 60 80 100 120 140
0 1 2 3 Curve fit for G=100
10 10 10 10 Oxidizer mass flux [kg/(m2s)]
Normalized viscosity [−]
Fig. 15 Regression rate results with error bars of paraffin 6805 with
Fig. 14 Normalized regression rate data versus viscosity, including additives.
HDPE fuel (red symbols indicate an oxidizer mass flux of 50 kg∕m2 s, and
blue symbols indicate 100 kg∕m2 s).
100
corresponding points are approximated by an exponential curve fit.
The following equation [Eq. (1)] was used for the curve fits:

characteristic velocity [%]


95
 m
μ
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

r_
b

Efficiency of
(1) 90
r_ref μref
0907
The corresponding parameters b and m for the different curve fits 85 6003/ 10% SA
are shown in Table 4. Equation (1) is important for extrapolating and 6805/ 10% SA
6805/ 5% Polymer
predicting the regression rates of new fuels with different viscosities. 80 6805/ 10% Polymer
It is expected that more tests at a broader mass flux range increase the 6805/ 2% Nanoclay
reliability of the results and give more information about the effect of 1276
75
the oxidizer mass flux. 0.5 1 1.5 2
Error bars are not shown in the regression rate data for better Oxidizer-to-fuel ratio [−]
visibility. Table 5 shows the definitions for the error calculations. Fig. 16 Overview of ballistic test efficiencies.
Figure 15 shows the associated errors in the regression rate data.

2. Combustion Efficiencies 100%


1800
Figure 16 shows the calculated combustion efficiencies in percent 95%
Characteristic velocity [m/s]

versus the average mixture ratio for all the tests. Calculations are 90%
based on the measured chamber pressure, the measured nozzle throat 1600
85%
area, the mixture ratio, and the amount of burned propellants. The
pure paraffin samples achieve efficiencies between 85 and 95% at low CEA c*exp/c*theo [%]
mixture ratios smaller than one. The highest efficiencies are achieved 1400 0907
6003/ 10% SA
for 6805 with 5% polymer, with all above 95%. The lower burning 6805/ 10% SA
rate results of 1276 and 6805 with 10% polymer result in efficiencies 1200 6805/ 5% Polymer
around 90% on average. Figure 17 shows the actual values of the 6805/ 10% Polymer
characteristic velocities. The theoretical efficiency data from 6805/ 2% Nanoclay
1276
Chemical Equilibrium with Applications (CEA) [23] for different 1000
percentages is also included. Here, it is seen that the actual 0.5 1 1.5 2 2.5
Oxidizer-to-fuel ratio [−]
characteristic velocities of 1276 and 6805 with 10% polymer are the
highest for all tests, although the efficiency in percentage is lower for Fig. 17 Overview of ballistic test efficiencies.
these tests in Fig. 16. This is because they are also closer to the
optimum mixture ratio of about 2.2, where the maximum
characteristic velocity is achieved. The fuel samples of 1276 and
6805 with 10% polymer show a rougher surface after the tests when
Table 4 Curve fit parameters compared with the pure samples. It has to be noted that no
Curve fit for : : : b m Curve fit R2 postcombustion chamber is included in this setup, so the combustion
products cannot react completely in the tests. It is seen in Fig. 17 that
Figure 13 Gox  50 kg∕m2 s 6.852 −0.183 0.99
Figure 13 Gox  100 kg∕m2 s 7.12 −0.225 0.993 combustion tests with 5% polymer at a lower mixture ratio achieve
Figure 14 Gox  50 kg∕m2 s 7.097 −0.213 0.988 the highest efficiencies.
Figure 14 Gox  100 kg∕m2 s 7.175 −0.232 0.998
3. Combustion Instability
Instabilities appear for all kinds of rocket engines: solid, liquid,
Table 5 Estimation of individual and hybrid [24]. They result in a failure of the engine in a small
measurement uncertainties amount of time by a sudden pressure increase or increased heat
transfer to the walls. For hybrids, instabilities are often limited to
Value Variable Quantity certain amplitudes of oscillations [25]. Although the pressure
Initial inner fuel diameter Δdi 0.2 mm amplitude might not be critical, it still has side effects, and therefore
Initial outer fuel diameter Δdo 0.2 mm should be avoided. During our tests, the pressure data were recorded
Initial fuel length Δl 0.5 mm with 20 kHz in order to also evaluate the possible high-frequency
Burn time Δtb 0.3 s
Fuel mass Δmf 0.1 g
acoustic instability. Most of the tests resulted in stable operation with
Oxidizer mass Δmox 1% low-pressure oscillations, so the regression rate data results were not
affected by oscillations.
KOBALD ET AL. 1251

IV. Conclusions Journal of Heat and Mass Transfer, Vol. 13, No. 12, 1970,
pp. 1925–1939.
The combustion behavior of different paraffin samples burning doi:10.1016/0017-9310(70)90093-1
with gaseous oxygen has been investigated in a 2-D radial [11] Nigmatulin, R., Nigmatulin, B., Khodzaev, Y. A., and Kroshilin, V.,
microburner. The regression rate data of the liquefying paraffin-based “Entrainment and Deposition Rates in a Dispersed-Film Flow,”
fuels were related directly to the viscosity of the corresponding liquid International Journal of Multiphase Flow, Vol. 22, No. 1, 1996,
fuel samples for the range of values investigated. An increase in the pp. 19–30.
liquid layer viscosity resulted in a decreased regression rate. An doi:10.1016/0301-9322(95)00044-5
exponential relation was found, which correlated the regression rate [12] Nakagawa, I., and Hikone, S., “Study on the Regression Rate of
of liquefying fuels as a function of their liquid viscosity, as shown in Paraffin-Based Hybrid Rocket Fuels,” Journal of Propulsion and
Power, Vol. 27, No. 6, 2011, pp. 1276–1279.
Figs. 13 and 14. The relation also depended on the averaged oxidizer doi:10.2514/1.B34206
mass flux during the tests. This relation covered liquefying paraffin- [13] Maruyama, S., Ishiguro, T., Shinohara, K., and Nakagawa, I., “Study on
based fuels over a wide range of viscosity values. In detail, two Mechanical Characteristic of Paraffin-Based Fuel,” 47th AIAA/ASME/
exponential curve fits were generated at two different oxidizer mass SAE/ASEE Joint Propulsion Conference and Exhibit, AIAA Paper
flux values, which differed slightly. Thereby, more tests at a greater 2011-5678, 2011.
range of oxidizer mass flux values should be done in order to evaluate [14] Kobald, M., Toson, E., Ciezki, H., Schlechtriem, S., Di Betta, S.,
the effect of this parameter. The experimental data also included high- Coppola, M., and De Luca, L. T., “Rheological, Optical and Ballistic
density polyethylene fuel, which was expected to not show droplet Investigations of Paraffin-Based Fuels for Hybrid Rocket Propulsion
Using a 2-D Slab-Burner,” EUCASS Book Series Advances in
entrainment. Thereby, in between these data must exist an upper limit
Aerospace Sciences — Progress in Propulsion Physics, Vol. 8, 2016,
for the entrainment process of liquefying fuels that is related to the pp. 263–282.
viscosity of the liquid layer. Combustion efficiencies showed good doi:10.1051/eucass/201608263
results for all paraffin fuels. Research on the effect of the blackening Wang, J., Severtson, S. J., and Stein, A., “Significant and Concurrent
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

[15]
agent showed advantages for dye due to a smaller amount that needed Enhancement of Stiffness, Strength, and Toughness for Paraffin Wax
to be used and no settlement during solidification. Mechanical Through Organoclay Addition,” Advanced Materials, Vol. 18, No. 12,
strength tests of the fuels showed good improvement of maximum 2006, pp. 1585–1588.
tensile strain and strength with polymer addition. doi:10.1002/(ISSN)1521-4095
[16] Kobald, M., “Combustion Phenomena of Advanced Hybrid Rocket
Fuels,” Ph.D. Dissertation, Univ. Stuttgart, Inst. für Raumfahrtsysteme,
Stuttgart, Germany, 2015.
Acknowledgments [17] Marano, J. J., and Holder, G. D., “General Equation for Correlating the
The cooperation and support of the Space Propulsion Laboratory Thermophysical Properties of n-Paraffins, n-Olefins, and Other
and the M11 team of the DLR, German Aerospace Center are greatly Homologous Series. 1. Formalism for Developing Asymptotic Behavior
acknowledged. Correlations,” Industrial and Engineering Chemistry Research, Vol. 36,
No. 5, 1997, pp. 1887–1894.
doi:10.1021/ie960511n
[18] Marano, J. J., and Holder, G. D., “General Equation for Correlating the
References Thermophysical Properties of n-Paraffins, n-Olefins, and Other
[1] Shimada, T., “Status Summary of FY 2011 Hybrid Rocket Research Homologous Series. 2. Asymptotic Behavior Correlations for PVT
Working Group,” 9th International Conference on Flow Dynamics, Properties,” Industrial and Engineering Chemistry Research, Vol. 36,
Tohoku Univ. Paper OS3-24, Sendai, Japan, 2012. No. 5, 1997, pp. 1895–1907.
[2] Carrick, P. G., and Larson, C. W., “Lab Scale Test and Evaluation of doi:10.1021/ie960512f
Cryogenic Solid Hybrid Rocket Fuels,” 31st AIAA/ASME/SAE/ASEE [19] Kontogeorgis, G. M., and Tassios, D. P., “Critical Constants and
Joint Propulsion Conference and Exhibit, AIAA Paper 1995-2948, Acentric Factors for Long-Chain Alkanes Suitable for Corresponding
1995. States Applications. A Critical Review,” Chemical Engineering
[3] Karabeyoglu, M. A., Altman, D., and Cantwell, B. J., “Combustion of Journal, Vol. 66, No. 1, 1997, pp. 35–49.
Liquefying Hybrid Propellants: Part 1, General Theory,” Journal of doi:10.1016/S1385-8947(96)03146-4
Propulsion and Power, Vol. 18, No. 3, 2002, pp. 610–620. [20] Paravan, C., “Ballistics of Innovative Solid Fuel Formulation for Hybrid
doi:10.2514/2.5975 Rocket Engines,” Ph.D. Dissertation, Politecnico di Milano,
[4] Karabeyoglu, M. A., “Transient Combustion in Hybrid Rockets,” Ph.D. Dipartimento di Ingegneria Aerospaziale, Milan, 2012.
Dissertation, Stanford Univ., Stanford, CA, 1998. [21] Toson, E., Kobald, M., Di Betta, S., De Luca, L. T., Ciezki, H., and
[5] Karabeyoglu, M. A., and Cantwell, B. J., “Combustion of Liquefying Schlechtriem, S., “Rheological and Ballistic Investigations of Paraffin-
Hybrid Propellants: Part 2, Stability of Liquid Films,” Journal of Based Fuels for Hybrid Rocket Propulsion Using a 2d Radial Micro-
Propulsion and Power, Vol. 18, No. 3, 2002, pp. 621–630. Burner,” 5th European Conference for Aeronautics and Space Sciences,
doi:10.2514/2.5976 Technical Univ. and Astrium, Munich, 2013, Paper 459.
[6] Karabeyoglu, M. A., Cantwell, B. J., and Stevens, J., “Evaluation of [22] Karabeyoglu, M. A., Cantwell, B. J., and Zilliac, G., “Development of
Homologous Series of Normal-Alkanes as Hybrid Rocket Fuels,” 41st Scalable Space-Time Averaged Regression Rate Expressions for Hybrid
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Rockets,” Journal of Propulsion and Power, Vol. 23, No. 4, 2007,
AIAA Paper 2005-3908, 2005. pp. 737–747.
[7] Kobald, M., Verri, I., and Schlechtriem, S., “Theoretical and doi:10.2514/1.19226
Experimental Analysis of Liquid Layer Instabilities in Hybrid Rocket [23] Gordon, S., and McBride, B. J., “Computer Program for Calculation of
Engines,” CEAS Space Journal, Vol. 7, No. 1, March 2015, pp. 11–22. Complex Chemical Equilibrium Compositions and Applications,”
doi:10.1007/s12567-015-0076-2 NASA RP-1311, Cleveland, OH, 1994.
[8] Kobald, M., Ciezki, H., and Schlechtriem, S., “Optical Investigation of [24] Yang, V., and Anderson, W., Liquid Rocket Engine Combustion
the Combustion Process in Paraffin-Based Hybrid Rocket Fuels,” 49th Instability, 1st ed., Vol. 169, Progress in Astronautics and Aeronautics,
AIAA/ASME/SAE/ASEE Joint Propulsion Conference, AIAA Paper AIAA, Washington, D.C., 1995, pp. 3–14.
2013-3894, 2013. doi:10.2514/4.866371
[9] Karabeyoglu, M. A., Zilliac, G., Cantwell, B. J., DeZilwa, S., and [25] Karabeyoglu, M. A., De Zilwa, S., Cantwell, B., and Zilliac, G.,
Castellucci, P., “Scale-Up Tests of High Regression Rate Paraffin-Based “Modeling of Hybrid Rocket Low Frequency Instabilities,” Journal of
Hybrid Rocket Fuels,” Journal of Propulsion and Power, Vol. 20, No. 6, Propulsion and Power, Vol. 21, No. 6, 2005, pp. 1107–1116.
2004, pp. 1037–1045. doi:10.2514/1.7792
doi:10.2514/1.3340
[10] Gater, R. A., and L’Ecuyer, M. R. L., “A Fundamental Investigation of C. Segal
the Phenomena that Characterize Liquid Film Cooling,” International Associate Editor
This article has been cited by:

1. Xue-li Liu, Song-qi Hu, Yin Wang, Wei-meng Zhang, Lin-lin Liu. 2023. Droplet entrainment characteristics of HTPB/
Paraffin blended fuels for hybrid rocket motors. Acta Astronautica 180. . [Crossref]
2. Mario Tindaro Migliorino, Marco Fabiani, Christian Paravan, Daniele Bianchi, Francesco Nasuti, Luciano Galfetti, Rocco
Carmine Pellegrini, Enrico Cavallini. 2023. Numerical and experimental analysis of fuel regression rate in a lab-scale hybrid
rocket engine with swirl injection. Aerospace Science and Technology 140, 108467. [Crossref]
3. Tsuyoshi Oishi, Mitsuru Tamari, Takashi Sakurai. 2023. Experimental Investigation of a Swirling-Oxidizer-Flow-Type
Hybrid Rocket Engine Using Low-Melting-Point Thermoplastic Fuel and Oxygen. Aerospace 10:8, 713. [Crossref]
4. Christian Paravan, Anwer Hashish, Valerio Santolini. 2023. Test Activities on Hybrid Rocket Engines: Combustion
Analyses and Green Storable Oxidizers—A Short Review. Aerospace 10:7, 572. [Crossref]
5. Jiaxiao Luo, Zelin Zhang, Xin Lin, Zezhong Wang, Wu Kun, Gongxi Zhou, Senhao Zhang, Fei Li, Xilong Yu, Jie Wu.
2023. Flame Dynamics in the Combustion Chamber of Hybrid Rocket Using Multiangle Chemiluminescence. Journal of
Propulsion and Power 39:4, 482-491. [Abstract] [Full Text] [PDF] [PDF Plus]
6. Yash Pal, Sasi Kiran Palateerdham, Sri Nithya Mahottamananda, Subha Sivakumar, Antonella Ingenito. 2023. Combustion
performance of hybrid rocket fuels loaded with MgB2 and carbon black additives. Propulsion and Power Research 12:2,
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

212-226. [Crossref]
7. Riccardo Bisin, Christian Paravan. 2023. A new strategy for the reinforcement of paraffin-based fuels based on cellular
structures: The armored grain — Ballistic characterization. Acta Astronautica 206, 284-298. [Crossref]
8. Yin Wang, Song-qi Hu, Xue-li Liu, Yan Zhang, Lin-lin Liu. 2023. Droplet entrainment and its role in the combustion
of HTPB/paraffin fuels. Acta Astronautica 204, 107-115. [Crossref]
9. James Meier, John Reynolds, Sean Whalen, Jaimin Patel, Michael J. Bortner, Gregory Young. 2023. Improved Hybrid
Rocket Performance by Additively Manufactured Gel-Infused Solid Fuels. Journal of Propulsion and Power 39:1, 97-105.
[Abstract] [Full Text] [PDF] [PDF Plus]
10. Xue-li Liu, Song-qi Hu, Yin Wang, Lin-lin Liu, Yan Zhang. 2023. Numerical investigation on the droplet entrainment
from gas sheared liquid film of hydroxyl-terminated polybutadiene/paraffin fuels. Physics of Fluids 35:1. . [Crossref]
11. Colin D. Hill, Will Nelson, Craig T. Johansen. 2022. Evaluation of a Paraffin/Nitrous Oxide Hybrid Rocket Motor with
a Passive Mixing Device. Journal of Propulsion and Power 38:6, 884-892. [Abstract] [Full Text] [PDF] [PDF Plus]
12. Sri Mahottamananda, Yash Pal, Mengu Dinesh, Antonella Ingenito. 2022. Beeswax–EVA/Activated-Charcoal-Based Fuels
for Hybrid Rockets: Thermal and Ballistic Evaluation. Energies 15:20, 7578. [Crossref]
13. Xin Lin, Dandan Qu, Xuedong Chen, Zezhong Wang, Jiaxiao Luo, Dongdong Meng, Guoliang Liu, Kun Zhang, Fei Li,
Xilong Yu. 2022. Three-dimensional printed metal-nested composite fuel grains with superior mechanical and combustion
properties. Virtual and Physical Prototyping 17:3, 437-450. [Crossref]
14. Faris M. AL-Oqla, Mohammed T. Hayajneh. 2022. Hybrid material performance assessment for rocket propulsion. Journal
of the Mechanical Behavior of Materials 31:1, 160-169. [Crossref]
15. Emmanuel Péres de Araújo, Leandro José Maschio, Leonardo Henrique Gouvêa, Luís Gustavo Ferroni Pereira, Ricardo
Vieira. 2022. Thermal, Viscosimetric and Thermomechanical Combined Assessment of Mixture Modelled Composite
Fuels for Hybrid Propulsion. Propellants, Explosives, Pyrotechnics 47:4. . [Crossref]
16. Yin Wang, Song-qi Hu, Xue-li Liu, Lin-lin Liu. 2022. Regression rate modeling of HTPB/paraffin fuels in hybrid rocket
motor. Aerospace Science and Technology 121, 107324. [Crossref]
17. Colin Hill. Evaluation of a Paraffin/Nitrous Oxide Hybrid Rocket Motor with Passive Mixing Device . [Abstract] [PDF]
[PDF Plus]
18. Anna Petrarolo, Mario Kobald. 2021. On the liquid layer instability process in hybrid rocket fuels. FirePhysChem 1:4,
244-250. [Crossref]
19. Riccardo Bisin, Alberto Verga, Daniele Bruschi, Christian Paravan. Strategies for Paraffin-based Fuels Reinforcement: 3D
Printing and Blending with Polymers . [Abstract] [PDF] [PDF Plus]
20. Carmine Carmicino, Giuseppe Gallo, Raffaele Savino. 2021. Self-consistent surface-temperature boundary condition for
liquefying-fuel-based hybrid rockets internal-ballistics simulation. International Journal of Heat and Mass Transfer 169,
120928. [Crossref]
21. Cagri Oztan, Victoria Coverstone. 2021. Utilization of additive manufacturing in hybrid rocket technology: A review. Acta
Astronautica 180, 130-140. [Crossref]
22. Cagri Oztan, Eric Ginzburg, Mert Akin, Yiqun Zhou, Roger M. Leblanc, Victoria Coverstone. 2021. 3D printed ABS/
paraffin hybrid rocket fuels with carbon dots for superior combustion performance. Combustion and Flame 225, 428-434.
[Crossref]
23. Jerome Messineo, Koki Kitagawa, Carmine Carmicino, Toru Shimada, Christian Paravan. 2020. Reconstructed Ballistic
Data Versus Wax Regression-Rate Intrusive Measurement in a Hybrid Rocket. Journal of Spacecraft and Rockets 57:6,
1295-1308. [Abstract] [Full Text] [PDF] [PDF Plus]
24. S.A. Rashkovskiy, S.E. Yakush. 2020. Numerical simulation of low-melting temperature solid fuel regression in hybrid
rocket engines. Acta Astronautica 176, 710-716. [Crossref]
25. Mustafa Basyal, Kaan Bilge, Nur B. Emerce, Utku C. Yildiz, Ugur Kokal, Arif Karabeyoglu. Synergistic Effect of Nano-
additives and Tackifier Resins on Hybrid Rocket Fuel Performance . [Abstract] [PDF] [PDF Plus]
26. Flávio D. A. Quadros, Pedro T. Lacava. 2019. Swirl Injection of Gaseous Oxygen in a Lab-Scale Paraffin Hybrid Rocket
Motor. Journal of Propulsion and Power 35:5, 896-905. [Abstract] [Full Text] [PDF] [PDF Plus]
27. Yash Pal, K. Harish Kumar, Yueh-Heng Li. 2019. Ballistic and mechanical characteristics of paraffin-based solid fuels.
CEAS Space Journal 11:3, 317-327. [Crossref]
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

28. Francesco Battista, Daniele Cardillo, Manrico Fragiacomo, Giuseppe Daniele Di Martino, Stefano Mungiguerra, Raffaele
Savino. 2019. Design and Testing of a Paraffin-Based 1000 N HRE Breadboard. Aerospace 6:8, 89. [Crossref]
29. Christian Schmierer, Mario Kobald, Konstantin Tomilin, Ulrich Fischer, Stefan Schlechtriem. 2019. Low cost small-
satellite access to space using hybrid rocket propulsion. Acta Astronautica 159, 578-583. [Crossref]
30. G.D. Di Martino, S. Mungiguerra, C. Carmicino, R. Savino. 2019. Computational fluid-dynamic modeling of the internal
ballistics of paraffin-fueled hybrid rocket. Aerospace Science and Technology 89, 431-444. [Crossref]
31. Anna Petrarolo, Mario Kobald, Stefan Schlechtriem. 2019. Optical analysis of the liquid layer combustion of paraffin-based
hybrid rocket fuels. Acta Astronautica 158, 313-322. [Crossref]
32. Giuseppe Daniele Di Martino, Carmine Carmicino, Stefano Mungiguerra, Raffaele Savino. 2019. The Application
of Computational Thermo-Fluid-Dynamics to the Simulation of Hybrid Rocket Internal Ballistics with Classical or
Liquefying Fuels: A Review. Aerospace 6:5, 56. [Crossref]
33. Suhang Chen, Yue Tang, Wei Zhang, Ruiqi Shen, Hongsheng Yu, Yinghua Ye, Luigi T. DeLuca. 2019. Innovative Methods
to Enhance the Combustion Properties of Solid Fuels for Hybrid Rocket Propulsion. Aerospace 6:4, 47. [Crossref]
34. Kamila P. Cardoso, Luiz F. A. Ferrão, Elizabete Y. Kawachi, Jessica S. Gomes, Márcio Y. Nagamachi. 2019. Ballistic
Performance of Paraffin-Based Solid Fuels Enhanced by Catalytic Polymer Degradation. Journal of Propulsion and Power
35:1, 115-124. [Abstract] [Full Text] [PDF] [PDF Plus] [Supplementary Material]
35. G. D. Di Martino, S. Mungiguerra, C. Carmicino, R. Savino, D. Cardillo, F. Battista, M. Invigorito, G. Elia. 2019. Two-
Hundred-Newton Laboratory-Scale Hybrid Rocket Testing for Paraffin Fuel-Performance Characterization. Journal of
Propulsion and Power 35:1, 224-235. [Abstract] [Full Text] [PDF] [PDF Plus]
36. Dinesh Mengu, Rajiv Kumar. 2018. Development of EVA-SEBS based wax fuel for hybrid rocket applications. Acta
Astronautica 152, 325-334. [Crossref]
37. Yi Wu, Xilong Yu, Xin Lin, Sen Li, Xiaolin Wei, Chuan Zhu, Linlin Wu. 2018. Experimental investigation of fuel
composition and mix-enhancer effects on the performance of paraffin-based hybrid rocket motors. Aerospace Science and
Technology 82-83, 620-627. [Crossref]
38. Giuseppe D. Di Martino, Stefano Mungiguerra, Carmine Carmicino, Raffaele Savino. Computational Fluid-Dynamic
Simulations of the Internal Ballistics of Hybrid Rocket Burning Paraffin-based Fuel . [Citation] [PDF] [PDF Plus]
39. Anna Petrarolo, Mario Kobald, Stefan Schlechtriem. Visualization of Combustion Phenomena in Paraffin-Based Hybrid
Rocket Fuels at Super-Critical Pressures . [Citation] [PDF] [PDF Plus]
40. Anna Petrarolo, Mario Kobald, Stefan Schlechtriem. 2018. Understanding Kelvin–Helmholtz instability in paraffin-based
hybrid rocket fuels. Experiments in Fluids 59:4. . [Crossref]
41. M. Kobald, U. Fischer, K. Tomilin, A. Petrarolo, C. Schmierer. 2018. Hybrid Experimental Rocket Stuttgart: A Low-
Cost Technology Demonstrator. Journal of Spacecraft and Rockets 55:2, 484-500. [Abstract] [Full Text] [PDF] [PDF Plus]
42. Mario KOBALD, Christian SCHMIERER, Ulrich FISCHER, Konstantin TOMILIN, Anna PETRAROLO. 2018.
A Record Flight of the Hybrid Sounding Rocket HEROS 3. TRANSACTIONS OF THE JAPAN SOCIETY FOR
AERONAUTICAL AND SPACE SCIENCES, AEROSPACE TECHNOLOGY JAPAN 16:3, 312-317. [Crossref]
43. Yue Tang, Suhang Chen, Wei Zhang, Ruiqi Shen, Luigi T. DeLuca, Yinghua Ye. 2017. Mechanical Modifications of
Paraffin‐based Fuels and the Effects on Combustion Performance. Propellants, Explosives, Pyrotechnics 42:11, 1268-1277.
[Crossref]
44. Nikita V. Muravyev, Konstantin A. Monogarov, Dmitry Prokopyev, Anatoly A. Bragin, Luciano Galfetti, Luigi T. DeLuca,
Alla N. Pivkina. 2017. Macro- vs Microcrystalline Wax: Interplay of Evaporation and Decomposition under Pressure
Variation. Energy & Fuels 31:8, 8534-8539. [Crossref]
45. Anna Petrarolo, Mario Kobald, Stefan Schlechtriem. Liquid Layer Combustion Visualization of Paraffin-based Hybrid
Rocket Fuels . [Citation] [PDF] [PDF Plus]
Downloaded by 151.68.26.87 on November 20, 2023 | http://arc.aiaa.org | DOI: 10.2514/1.B36207

You might also like