You are on page 1of 15

Aerospace Science and Technology 89 (2019) 392–406

Contents lists available at ScienceDirect

Aerospace Science and Technology


www.elsevier.com/locate/aescte

Performance comparison of oxidizer injectors in a 1-kN paraffin-fueled


hybrid rocket motor
M. Bouziane a,∗ , A.E.M. Bertoldi b , P. Milova c , P. Hendrick c , M. Lefebvre a
a
Department of Chemistry, Royal Military Academy, Avenue de la Renaissance 30, 1000 Brussels, Belgium
b
Faculty of Gama, University of Brasília, Área Especial de Indústria e Projeção A, 72444-240 Brasília, Brazil
c
Aero-Thermo-Mechanics department, Université Libre de Bruxelles, F.D. Roosevelt Avenue 50, 1050 Brussels, Belgium

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

Article history: To investigate the effects of oxidizer injection on the performance of hybrid rocket motors (HRMs), we
Received 11 September 2018 have designed, manufactured and tested four types of injector: showerhead (SH), hollow-cone (HC),
Received in revised form 5 March 2019 pressure-swirl (PSW) and vortex (VOR). This study is motivated by the fact that the experimental
Accepted 5 April 2019
measurements of N2 O/paraffin firings are poorly presented in the open literature. Besides few publications
Available online 10 April 2019
are dedicated to the characterization of novel types of injectors in hybrid rocket propulsion application
Keywords: such as HC, PSW, and VOR. It is advantageous that the study was conducted in the same motor
Hybrid rocket motors configuration, with the advantage that it allowed to compare the performance of different types of
Injector design injectors.
Injection system This paper analyzes the influence of the oxidizer injector design on the main performance parameters,
Regression rate such as fuel regression rate, specific impulse and combustion efficiency. First, in order to observe injector
O / F ratio spray qualities, a series of cold tests using liquid water and liquid nitrous oxide are carried out, providing
Specific impulse
a good understanding of the spray profiles. Then, the motor performance data is obtained by a series of
firing tests using N2 O as oxidizer and paraffin as fuel. Comparison of the various injectors data is made
with the same average oxidizer mass flux and feeding pressure. The showerhead is used as a benchmark
in this study. During this experimental analysis, the VOR injector exhibits the highest regression rate,
followed by HC and SH. Because the assessment of the regression rate was not enough to explain all
the effects of the injectors on the motor performance, firing tests of small-scale hybrid motor have
been carried out. In terms of spray properties, the PSW exhibits significant differences compared with
the other injectors; it generates the smallest Sauter Mean Diameter (SMD) in the formed spray and
achieves good atomization. In spite of the fact that the PSW injector leads to the lowest regression rate,
it provides good specific impulse, increases the oxidizer-to-fuel ratio (O / F ), as well as a uniform and
smooth consumption of the paraffin fuel grain. VOR leads to the highest specific impulse. In terms of
stability, VOR, HC and SH injectors exhibit lower oscillations in the chamber pressure. Some observations
are made on exhaust plume intensity developed during combustion, and in firing tests with SH a blow-
out phenomenon occur often. Similarly to liquid engines, it is possible in hybrid motors to increase the
global propulsive performance using alternative designs of the injection system.
© 2019 Elsevier Masson SAS. All rights reserved.

1. Introduction cost. However, this technology has some development challenges


in comparison with the other more mature propulsion systems, as
Recently hybrid rocket motor propulsion systems have received liquid and solid rockets [1,2]. One of the principal drawbacks of
substantial renewed interest as possible design alternatives to hybrid technology is the low regression rate of a solid fuel, hence
presently used liquid and solid propulsion systems. This renewed relatively poor combustion efficiency and too low specific impulse
interest is principally due to the inherent advantages over the [3,4].
other chemical propulsion systems, in particular the safety aspects
Moreover, the oxidizer injection characteristics play a substan-
and capability to control and throttle the motor thrust at low
tial role in hybrid rocket motor performance [5]. In typical config-
urations, a liquid oxidizer is injected into the combustion chamber
by means of an atomizer, and a spray is formed. The liquid ox-
* Corresponding author.
E-mail address: mohammed.bouziane@dymasec.be (M. Bouziane). idizer droplets are vaporized in the pre-chamber, flown through

https://doi.org/10.1016/j.ast.2019.04.009
1270-9638/© 2019 Elsevier Masson SAS. All rights reserved.
M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406 393

Nomenclature

Variable Rx Exit radius in VOR injector


R inj Injector port radius
At Throat nozzle area
A inj Area of orifice in injector plate ṙ Average regression rate
CD Discharge coefficient Sg Geometrical swirl number

c exp Experimental characteristic velocity t Film thickness in PSW injector

cth Theoretical characteristic velocity tb Burning time
CF Thrust coefficient
ρ Density
Da Air core diameter in PSW injector
ρa Air density
ξ Efficiency of the motor
D inj Diameter of orifice in injector plate
θ Cone semi angle in PSW injector
df Final port diameter of fuel grain
Ds Swirl chamber diameter in PSW injector
μ Dynamic viscosity
D0 Discharge diameter in PSW injector
σ Surface tension
F Thrust Abbreviations
F Average thrust
HC Hollow-cone injector
I sp Specific impulse
HDPE High density polyethylene propellant
G ox Average oxidizer mass flux HRM Hybrid rocket motor
G ox_i Average initial oxidizer mass flux LOX Liquid oxygen
g0 Gravitational acceleration constant N2 O Nitrous oxide
Lg Fuel grain length VOR Vortex injector
Lp Tangential entry passage length of PSW injector PMMA Poly-methyl methacrylate
Ls Swirl chamber length of PSW injector PSW Pressure-swirl injector
L0 Nozzle length of PSW injector PVC Polyvinyl-chloride
ṁ Total average mass flow RMA Royal Military Academy of Belgium
ṁox Average oxidizer mass flow rate SH Showerhead injector
ṁ f Average fuel mass flow rate SMD Sauter Mean Diameter
m f Burnt fuel mass ULB Université Libre de Bruxelles
N inj Number of orifices in injector plate UnB University of Brasilia
O /F Oxidizer-to-fuel ratio
Subscripts
O /F Experimental average oxidizer-to-fuel ratio
u ox Oxidizer velocity f Fuel
uz Axial component of the oxidizer velocity inj Injector
Pc Combustion chamber pressure G Gas
Pc Average combustion chamber pressure L Liquid
P tbt Test bench tank pressure of N2 O ox Oxidizer

the combustion port, and then react with the fuel grain to achieve merically that the presence of swirl velocity component increases
stable combustion. Thus, the combustion process will be severely the velocity magnitude in the flow field and causes an increment
influenced by the incoming oxidizer flow pattern. In fact, the flow of the velocity gradient at the fuel surface. Heat and mass trans-
characteristics can significantly affect the overall behavior of the port on the fuel surface increases, and consequently regression rate
motor in terms of thrust, fuel consumption, combustion efficiency augments [12].
and combustion stability [6]. Therefore, there is a real need for
Another atomizer, reliable and simple to design, is the pressure
studying the injector influence to develop reliable tools to pre-
swirl injector (PSW), and its application in hybrid rocket motors
dict the fuel regression rate under different oxidizer flow condi-
is relatively new. In Ref. [13], de Morais Bertoldi et al. observed
tions.
Hybrid rockets can use any injector initially designed for liquid an increase of the regression rate by 20% using PSW compared to
engines, with required modifications due to the absence of one showerhead injector.
propellant component – usually the liquid fuel. An injector type In this work, four different injectors, showerhead (SH), hollow-
that is widely used in hybrids is the vortex injector due to the cone (HC), pressure-swirl (PSW) and vortex atomizer (VOR) were
relatively broad existing investigations, both experimental and nu- designed and manufactured to perform a series of firing tests using
merical. Vortex injector induces a rotational motion on the flow a 1-kN lab-scale hybrid rocket motor developed at Université Libre
with thus a substantial tangential velocity component of the flow. de Bruxelles (ULB) in collaboration with Royal Military Academy of
Knuth et al. stated that it is possible to induce higher solid fuel Belgium (RMA).
regression rates, around 8 times higher than the similar classi-
Significant differences in the effect of the injectors on the motor
cal hybrid engines when using vortex injectors [7–9]. Yuasa et al.
performance are observed. SH is taken as a reference configuration
tested also gaseous oxygen injected by vortex injector in the front
due to its simplicity. It shows relatively good performance in con-
head of the grain, achieving a regression rate 3 times faster than
the one observed in classical hybrid motors [10]. Bellomo et al. trast to HC. VOR tests agree with the literature data in terms of
investigated liquid nitrous oxide injected using a vortex configura- regression rate and relatively stable combustion [11]. Our regres-
tion: an increase in the regression rate up to 51% was measured sion rate results of PSW are not in line with Ref. [13]. Nevertheless,
while the instability in the combustion chamber was lower than our results of PSW show relevant and consistent droplet size, pre-
in axial injection configuration [11]. Kumar et al. have proved nu- heating phenomenon and reduction of the flame blow-out.
394 M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406

Fig. 1. Schematic of ULB-HRM test bench.

2. Set up and methodology

2.1. Lab-scale test bench design

A new lab-scale test bed, the ULB-HRM test bench, with au-
tomatic control system and data acquisition was designed and
implemented to conduct experimental investigations on a liquid
nitrous oxide/paraffin-based fuel 1-kN hybrid motor. Fig. 1 shows
a schematic of the ULB-HRM test stand, which is described with
more details in Ref. [14]. The test stand consists of a horizontal
bench that allows quick and secure mounting of the hybrid motor
and its subsystems, such as liquid nitrous oxide tanks, feed system,
pyrotechnic ignition device, and data acquisition system.
The test bench allows the measurement of the thrust using a
load cell SENSY model 2965. Taking the sensor sensitivity and our
applications into account, a careful calibration has been performed
up to 1 kN, and, in the calibration range, the load cell appears
to be quite linear. The weight of the oxidizer tank is given by a Fig. 2. 3D view of the ULB-HRM motor.
second load cell TEDEA model 615. Both are connected to a COND-
SGA charge amplifier. Different pressures are measured, first one is
on the feed line, second one is the pressure of test bench oxidizer
tank, then is measured before the injector and in the pre-chamber. global view of the set up. The second camera is placed aside from
Pressure transmitters are strain-gage base transducers with a range the nozzle, to record the exhaust plume.
of 100 bar. A Kistler piezoelectric pressure sensor coupled with a
charge meter type 5015 is used to measure the chamber pressure 2.2. Motor design
at the fore-end of the motor. The temperature is recorded in differ-
ent locations on the lines by using thermocouples type K, plugged
into a datalogger TM500. The hybrid motor (Fig. 2) is composed of three main parts: (i)
A LabView program is integrated with the data acquisition sys- 100 mm length pre-chamber (compartment between the mouth of
tem using the NI USB-6218 and NI USB-9215 cards, and a control the injector and the front area of the grain fuel). The pre-chamber
box, enabling remote control of the whole test bench sequence of contains the ignition cartridge and the injector plates. This modu-
events. lar construction allows an easy substitution of its various compo-
Specific details regarding the design and development of ULB- nents.; (ii) combustion chamber with an internal diameter equal to
HRM motor design, control system, ignition system, and testing 140 mm, containing the fuel grain and holding the post-chamber;
procedure are presented in Ref. [14]. and (iii) a convergent-divergent nozzle with a 22 mm critical sec-
Two video cameras, GoPro 5 black, are employed and are re- tion diameter.
motely controlled by a smartphone. The first is located, for safety The total motor length is 440 mm with an external diameter of
purposes, behind the test bench and enables the operator to get a 154 mm.
M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406 395

Table 1
Theoretical parameters of the ULB-HRM.

Parameter Parameter
Oxidizer N2 O O/F shift (theoretical) ∼7.9
Fuel Paraffin Oxidizer mass flow rate (g/s) 550
Nominal thrust (kN) 1.0 Average fuel mass flow rate (g/s) 70
Chamber pressure (bar) 20 to 30 Total mass flow rate (g/s) 620
Nozzle expansion ratio 5.2 Operation time (s) 5 to 10

Table 2
Characteristics of the injectors.

Injector D inj (mm) N inj u z (m/s) SMD (μm)


SH 1.4 11 42.06 1140
HC 1.4 11 40.63 1140
PSW 3 6 14.93 26
VOR 3 6 / / Fig. 4. Illustration of the initial interior diameter of the fuel grain compared to the
SH injector plate orifices.

Table 3
Nitrous oxide properties.

Liquid density Gas density Surface tension Dynamic viscosity


ρox_L (kg/m3 ) ρox_G (kg/m3 ) σ (N/m) μL (Pa.s)
772.25 1.83 0.24 325 × 10−6

Fig. 3. Injectors configurations: a – Shower head plate injector (SH), b – Hollow-cone


plate injector (HC), c – Pressure-swirl plate injector (PSW), and d – Vortex plate
injector (VOR).
Fig. 5. Illustration of HC injector direction jet on the grain fuel.

2.3. Oxidizer injection design


is estimated by equation (3) as proposed by Tanasawa and Toy-
oda [22]. The results are summarized in Table 2, and properties of
The four injectors used in this research are tested under the
nitrous oxide are summarized in Table 3.
same motor conditions (fixed combustion chamber and post-
ṁox
chamber geometry). All injectors are manufactured in aluminum uz = (2)
and designed based on the theoretical parameters of the ULB-HRM ρox · A inj
motor (see Table 1) aiming to delivery 550 g/s and a pressure drop
 0.25  
D inj σ μL
of 25 bar. Equation (1) provides the correlation between these pa- SMD = 47 1 + 331 (3)
u ox ρox_G (ρox_L · σ · D inj )0.5
rameters. The injectors are characterized by a specific geometry
and the number of holes to produce the four desired flow patterns The hollow-cone (HC) injector has 11 holes, similar to the SH
(Table 2). The four configurations are shown in Fig. 3. injector, each orifice has a diameter of 1.4 mm and it is inclined
15◦ in order to drive the flow to the lean side of the fuel grain (i.e.

ṁox = C d · N inj · A inj · 2 · ρox ·  P (1) the edge of the fuel grain closer to the injector), as it is shown in
Fig. 5. Equations (2) and (3) are used for HC injector too, including
The showerhead (SH) has 11 orifices with 1.4 mm diameter, the angle of 15◦ , and the results are mentioned in Table 2.
which are spread equally in two different radiuses and one ori- The pressure-swirl (PSW) injector has 6 atomizing elements
fice in the center in order to deliver a homogenous distribution with a discharge diameter of 3 mm each. The use of PSW in hy-
of the oxidizer into the chamber. Because the showerhead shows brid rocket motors is relatively new, but it is commonly used in
simplicity, has a large database, and is easy to manufacture, it will gas turbines and liquid propellant rockets [22]. In PSW atomizers,
be used as a reference in this work. An extensive research about Fig. 6, the liquid is fed to the injector through tangential passages
showerhead (direct injection) was done by Carmicino et al. [4–6, giving the liquid a high angular velocity, and forming, in the swirl
15–18]. Also, various numerical and experimental research groups chamber, a liquid layer with a free internal surface, thus creating
use axial injection as reported in Refs. [11,19–21]. The configura- a gas-core vortex. The liquid is then discharged from the nozzle in
tion of the injector with the initial grain port diameter is shown the form of a hollow conical sheet which then breaks up into small
in Fig. 4. Based on the theory of a plain orifice injector, the axial droplets [22,23]. The PSW generates the smallest Sauter Mean Di-
injection velocity (u z ) is given by the equation (2) and the SMD ameter (SMD) in the formed spray and is known to achieve good
396 M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406

injector characteristics, a single PSW orifice, located in the center


of the PSW plate has been used. For specific details regarding the
injectors design and characterization, one may refer to Ref. [24].

2.4. Fuel grain preparation

In this work, pure paraffin is used as fuel. The fuel grains are
manufactured in a way to produce crack-free and void-less grains.
The paraffin is purchased in granular form from Brenntag NV com-
pany. The commercial paraffin is heated until melting point tem-
perature. The melted fuel is poured into two polyvinyl-chloride
cartridges, one over the other, fixed on a metal base. A central tube
makes the size and shape of the internal port of the fuel grain.
An additional quantity of melted paraffin is poured during the so-
lidification process to avoid cavities and cracks due to significant
shrinking of paraffin when cooling down. After cooling, the grain
is removed from the metal mold and goes to the workshop for de-
Fig. 6. Cross section of PSW injector.
tailed visual inspection. The residual paraffin fuel can be re-melted
and included in a new grain.
During the manufacturing process, the fuel grains are produced
atomization. Experimentally, SMD value using water ranges from
in such way that the variation between the different grain lengths
100 to 40 μm for pressure-drop ranging from 10 to 40 bar as re-
does not exceed 1 mm. The grain length is calculated in opti-
ported in Ref. [24]. Note that, with the laser device available in the mum conditions to have the best performance of the 1-kN HRM.
ATM department, we are not able to measure the SMD using ni- Note that more details of the optimization method are described
trous oxide. The axial component of the oxidizer speed and SMD in Ref. [14]. The length of fuel grains used equals to roughly 108
are given by the equations (4) and (5) respectively [22], where t mm with a total mass of 1250 g. The fuel density, ρ f , is estimated
is the film thickness and is calculated using equation (6) proposed to be 0.88 g/cm3 .
by Badami et al. [22,25]. Multiple tests can be carried out per day, just by changing the
fuel cartridge. The polyvinyl-chloride case acts as thermal protec-
ṁox
uz = (4) tion element for the combustion chamber internal wall too.
ρox · ( A 0 − Aa )
 0.25
σ μL 3. Theory and calculation
SMD = 4.52 (t cos θ)0.25
ρa  P 2
 0.25 For each test, data about thrust, oxidizer consumption, temper-
σρL ature and pressure are collected.
+ 0.39 (t cos θ)0.75 (5)
ρa  P Pressure data is monitored in four locations: in the combustion
    0 .5  chamber, in the test bench tank (feeding pressure), upstream and
D0 tan2 θ downstream of the injector plate. The data acquisition sample rate
t= 1− (6)
2 1 + tan2 θ is set to 8192 Hz (213 ) to obtain good signal accuracy.
Temperature data is sampled at 1 Hz in three locations: near
The vortex (VOR) injector induces a rotational motion on the
the commercial bottles of N2 O, in the oxidizer test bench tank and
flow. Several configurations of vortex injector can be found in the
before the injector.
technical literature, depending on the velocity components given to
A scale is used to weigh the pre- and post-test mass of the solid
the flow at the inlet. In our case, 6 orifices of 3 mm diameter are fuel grain.
selected with an injection angle of 45◦ , giving an axial and tan- The liquid N2 O is transferred from the commercial cylinders to
gential velocity component. The configuration of the orifices was the test facility tank and pressurized until 60 bar using N2 gas
chosen based on the Ref. [11] to compare firing tests results. The pressure feed.
geometrical swirl number, S g , is given by equation (7) which is The calculated regression rate is the average rate determined by
valid for a full-tangential (90◦ ) injector. In the case of a mixed the diameter variation of the fuel combustion port during the total
axial–tangential injector (as our VOR), the swirl number should be burning time, and is given by equation (8):
multiplied by the sine of the injection angle [11]. The geometrical
swirl number for the configuration considered in this paper equals d f − di
ṙ = (8)
to 31. 2tb
( R x − R inj ) · R x The definitions for the burning time, action time, and pres-
Sg = (7)
N inj · R 2inj sure rise or ignition rise time are defined in Fig. 9 [3]. A common
method is the aft tangent bisector method, as shown in Fig. 9 to
First, to observe the injectors spray qualities, a series of cold determine end of burning time, but this time can also be deter-
tests using water and nitrous oxide are carried out. A good under- mined by computer analysis [3]. Based on this, in our application,
standing of the spray profiles can be observed in these cold tests. the burning time tb is calculated observing the ascending and de-
Fig. 7 shows the flow pattern of the water discharged through in- scending peaks of the chamber pressure. A typical graph of cham-
jectors SH, HC, PSW, and VOR respectively, with a 30 bar pressure- ber pressure is presented in Fig. 10 where one notices the points
drop. Fig. 8 presents the injection of liquid N2 O, which is self- defining the burning time. In addition to the pressure traces, the
pressurized inside commercial bottles at around 40 bar in labo- recorded video helped to visualize the burning time using the im-
ratory conditions through the same series of injectors. Note that ages of the beginning and the end of the combustion. In general,
during the cold condition pre-testing, and to better visualize the the starting burning time is from 5 to 10% of the initial maximum
M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406 397

Fig. 7. Water discharged at 30 bar through a – SH, b – HC, c – PSW, and d – VOR injectors.

Fig. 8. Liquid N2 O discharged through a – SH, b – HC, c – PSW, and d – VOR injectors.

Fig. 9. Definitions of burning time and action time [3].


Fig. 10. Typical chamber pressure record (SH injector).

value, and the end of burning time represents 20 to 40% of the In this relation m f is the total mass of the burnt fuel and L g
initial maximum value. It varies because the combustion is not to- is the fuel grain length.
tally similar from one test to another. In addition, the variation of The liquid N2 O mass flow rate is measured by two means. First,
choosing the end of burning time is too small because the pressure a calibrated venturi is installed in the feed line, but due to a leak-
or thrust termination descends is quite fast in our measurements. age in the connections the device has been removed. To overcome
And it is not accurate to use the valve actuation time, because this the problem, a load cell is used to compute the mass oxidizer con-
does not account for the response time of the valve shutoff and sumption during the test.
of the overall oxidizer line; indeed, these response times affect the The average oxidizer mass flow rate is calculated by dividing
observed burning time.
the total injected mass of N2 O by the burning time. A typical graph
The initial port diameter, di , is an input data, measured before
showing these values is depicted in Fig. 11. These data, together
the tests, and is equal to 30 mm for all tests presented in this
with the measured fuel mass, are used to evaluate the total mass
paper. This helps to enable an easy and fair comparison between
flow rate. The average oxidizer-to-fuel ratio O / F is calculated us-
firings, such as the initial port diameter has a significant effect on
ing equation (10):
regression rate in HRMs as reported in Refs. [15,26,27].
The final port diameter, d f , cannot be measured directly due to ṁox
the complicated (slightly deformed) fuel geometry after combus- O /F = (10)
ṁ f
tion. A more precise way to estimate the final port diameter is to
use the fuel mass variation expressed by equation (9). The oxidizer mass flux is defined as the instantaneous oxidizer
mass flow rate over grain port cross-sectional area [3]. Then, its
 1/2
4m f average formula is given by equation (11), according to [28]. And
d f = d2i + (9)
π · ρ f · Lg for initial values is calculated using equation (12).
398 M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406

Fig. 11. Weight of N2 O injected during a firing test.


Fig. 12. Characteristic velocity graph showing the calculation method of the efficien-
16ṁox cies. (For interpretation of the colors in the figure(s), the reader is referred to the
G ox = (11) web version of this article.)
π (di + d f )2
4ṁox
G ox_i = (12) Thus, the first defined efficiency at O / F is calculated using a
π d2i ∗ at number “1”, and the second efficiency of optimum
value of cth
∗ at number “2”.
O / F is calculated using a value of cth
The specific impulse represents the thrust per unit of propellant
“weight” flow rate [3]. A higher number often indicates better per- Thrust coefficient C F is a dimensionless multiplication factor
formance, i.e. the motor with the higher value of specific impulse and signifies the degree to which the thrust is amplified by the
is more efficient because it produces more thrust for the same nozzle. Its typical values are ranging from just under 1.0 to roughly
amount of propellant. The specific impulse is calculated based on 2.0 and is defined by equation (16) [3]. This parameter is used in
the following equation (13): our calculation analysis also to quantify the quality of the firing
test.
F
I sp = (13) F
ṁ · g 0 CF = (16)
P c · At
The combustion efficiency of the motor is given by equation
(14). It is the ratio of the measured characteristic velocity, calcu- Each firing is recorded using a video camera to investigate the
lated by equation (15), and theoretical characteristic velocity cal- exhaust plume. The camera records with a speed of 120 fps and
culated with the EXPLO5 thermochemical software [29]. a resolution of 1080 dpi. For easy illustration and characterization
of the exhaust plumes, we defined an “intensity factor” with the

c exp following values:
ξ= ∗ (14) “2” for a large explosion flame as shown in Fig. 13a,
cth
“1” for stable flame combustion as shown in Fig. 13b,
∗ P c · At “0.5” for rather stables flames as shown in Fig. 13c, and
c exp = (15)
“0” for flame blow-out and low intensity flame as shown in

Fig. 13d.
In the data analysis, two different combustion efficiencies were We finally note that the ignition cartridge contribution to the
specified and calculated. First, it is calculated, applying formula of thrust is negligible because the fuel mass consumed during the
Eq. (15), with a cth∗ corresponding to O / F . This efficiency repre-
ignition phase is less than 15 grams which represents less than
sents a realistic estimation of the motor efficiency because it uses 1.2% of the total fuel mass.
an experimentally measured average oxidizer-to-fuel ratio. A sec-
ond combustion efficiency was also calculated related to the opti-
mum condition (i.e. using cth∗ corresponds to O / F 4. Experimental results
opt ) whatever the
result of O / F .
Fig. 12 shows the theoretical characteristic velocity at 17.5 bar The firing tests were conducted at the 1 Wing Air Base of Bel-
as combustion chamber pressure. The experimental characteristic gium at Beauvechain. No major problems have occurred during
velocity is represented by a red dot. Two additional points are these tests, except a few tests where mass of the N2 O tank or
highlighted in this graph: thrust have not been correctly recorded. These tests are not post-
processed.
– “1”: the theoretical characteristic velocity corresponding to In some other cases, the grain fuel moves inside the combustion
O /F . chamber and the signal of the Kistler pressure sensor in the com-
– “2”: the theoretical characteristic velocity corresponding to bustion chamber is not recorded. In these cases, the pre-chamber
O / F opt . pressure is used for the assessment of the P c .
M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406 399

Fig. 13. Exhaust plumes characterized by an intensity factor with values of: a – “2”, b – “1”, c – “0.5”, and d – “0”.

Table 4
Test results obtained with SH injector (average values).

Test n◦ P tbt , ṁox , tb , df , G ox_i , G ox , O /F ṙ, I sp , Pc, CF ξ, % at O / F ξ, % at O / F opt


bar g/s s mm g/cm2 s g/cm2 s mm/s s bar
SH-1 60 386.4 8.28 132.8 54.7 7.4 2.6 6.2 167.5 17.9 1.29 98.0 83.4
SH-2 60 380.3 8.29 132.8 53.8 7.3 2.5 6.2 147.1 17.0 1.19 93.8 79.4
SH-3 60 387.0 7.25 119.3 54.7 8.8 2.8 6.2 161.4 17.2 1.27 95.4 81.8
SH-4 65 417.4 8.10 132.8 59.1 8.0 2.7 6.3 156.2 19.1 1.21 96.7 83.3

Fig. 15. Typical exhaust plume with SH injector in a firing test.

out occurs, and the flame falls back downstream of the injector.
Meanwhile, heat transfer to the fuel grain surface continues and
produces more fuel vapor, which, when mixed appropriately with
the injected oxidizer, causes an external explosion. The second
explosion develops a shockwave that propagates upstream to re-
establish the flame. Flame blow-out can occur as a result of the
mixture ratio variation, as cited in Refs. [30,31]. Generally, one
observes a strong flame blow-out during the first half of the com-
bustion, i.e. between 4.2 to 5.8 s. The backflow is what we observe
Fig. 14. Thrust and pressures as function of time, test #SH-2. in the trace in Fig. 16 after the second explosion. The high axial
velocity produced by the SH shape propels the gaseous oxidizer
4.1. Test with SH injector through the nozzle, reduces the residence time of the latter in
the combustion chamber, prevents appropriate mixing with the
The benchmark tests are performed using SH injector. Table 4 evaporating fuel, and eventually exits the motor without proper
reports the average performance obtained with this injector. In test reaction/combustion.
#SH-4, the feeding pressure is set at 65 bar. Fig. 17 exhibits the fuel grain after combustion. The grain is
A typical graph of the generated thrust and three pressure almost fully consumed due to the high regression rate. For this rea-
records, feeding pressure, pressure before injector and chamber son, the burning time (equal to the N2 O injection time) has been
pressure are presented in Fig. 14. At the end of the thrust record, a reduced in the following tests. The paraffin fuel is burned radially,
peak appears due to the full consumption of the paraffin fuel and but, in addition, also axially reducing the length of the grain during
the reaction of the PVC case of the grain. the combustion process. The traces left by the 5 external orifices
In Fig. 15 a typical exhaust plume is shown, during the steady of the injector are clearly visible.
phase of the combustion. Fig. 16 represents the evolution of the
flame intensity value generated by the motor during test #SH-2. 4.2. Tests with HC injector
Two peaks are noticeable in Fig. 16, representing an explosive
combustion (hard ignition) in the plume. The first explosion is As detailed in paragraph 2.3, the HC injector is manufactured
caused by the injection of low temperature oxidizer; hence a blow- to drive the oxidizer to the inlet port of the grain with an angle
400 M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406

Fig. 18. Thrust and pressure as function of time, test #HC-1.

Fig. 16. Intensity value of the exhaust plume of firing test #SH-2.

Fig. 17. Fuel grain after SH firing test, a – injector side and b – nozzle side. Fig. 19. Exhaust plume with HC injector in a firing test.

of 15◦ . We assumed that this kind of incoming flow will create a


strong recirculation zone due to the impact of the oxidizer spray
with the paraffin fuel on the lean side of the grain. Table 5 summa-
rizes the average obtained results. Because of the high regression
rate of fuel, burning time has been reduced to 6 s.
Fig. 18 shows the motor thrust for test #HC-1 and the main
pressure records.
Steady state exhaust plume obtained with the HC injector is
visible in Fig. 19. The burning particles sparkling around the flame
come from the decomposition of the PVC casing. Fig. 20 shows
the variation of the flame intensity (test #HC-1). Most of the time
the flame is small, and some oxidizer flows out through the nozzle
without burning. Even during the ignition phase, no actual external
explosion is noticed, as it appears from video records. A picture-
by-picture analysis shows an amount of gas exhausting the nozzle
without actual combustion. We speculate that this is the reason
for the low specific impulse of this series of tests (Table 5).
After the tests performed with HC, the remaining fuel grain
is significantly deformed and does not allow investigation of the
combustion surface. It is obvious that the grain, softened by the
temperature, moves slightly inside the chamber during combus-
tion, resulting in the observed deformation. Fig. 20. Intensity value of the exhaust plume of firing test #HC-1.

Table 5
Test results with HC injector (average values).

Test n◦ P tbt , ṁox , tb , df , G ox_i , G ox , O /F ṙ, I sp , Pc, CF ξ, % at O / F ξ, % at O / F opt


bar g/s s mm g/cm2 s g/cm2 s mm/s s bar
HC-1 60 413.2 6.24 121.9 58.5 9.1 2.4 7.4 129.8 18.1 1.08 92.9 77.8
HC-2 60 412.3 6.36 123.4 58.3 8.9 2.5 7.3 156.5 18.3 1.28 93.0 78.8
HC-3 60 410.1 6.35 124.4 58.0 8.8 2.4 7.4 152.6 18.6 1.23 95.5 79.9
HC-4 60 431.5 6.35 124.5 61.0 9.2 2.5 7.4 141.2 19.3 1.14 94.4 79.8
HC-5 60 424.6 6.34 122.1 60.1 9.3 2.6 7.3 132.4 18.9 1.07 93.9 79.9
M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406 401

Table 6
Test results obtained with PSW injector (average value).

Test n◦ P tbt , ṁox , tb , df , G ox_i , G ox , O /F ṙ, I sp , Pc, CF ξ, % at O / F ξ, % at O / F opt


bar g/s s mm g/cm2 s g/cm2 s mm/s s bar
PSW-1 60 368.0 8.45 105.8 52.1 10.2 4.0 4.5 138.6 14.4 1.15 78.4 78.0
PSW-2 60 367.3 8.32 106.6 52.0 10.0 3.9 4.6 153.7 13.9 1.32 76.3 75.2
PSW-3 70 433.8 8.30 108.7 61.4 11.5 4.4 4.7 160.0 16.8 1.30 79.1 79.0
PSW-4 75 434.3 8.22 109.6 61.4 11.4 4.4 4.8 177.9 17.6 1.40 81.8 82.0
PSW-5 80 473.3 8.26 107.0 67.0 12.8 4.9 4.7 150.4 19.4 1.14 84.3 85.1
PSW-6 80 439.1 8.29 111.8 64.8 11.6 4.4 4.9 169.4 19.1 1.28 84.3 84.1

4.3. Tests with the PSW injector

The N2 O is injected through 6 orifices, each one vaporizes the


incoming oxidizer and discharges it as a hollow-cone. The cones
are interfering with each other and collide against the pre-chamber
wall, creating a recirculation zone and the preheating of the in-
jected oxidizer. The average data obtained using PSW is collected
in Table 6.
As for the previous tests, the feeding pressure has been set to
60 bar (#PSW-1 and #PSW-2). The resulting oxidizer mass flow
is quite low, preventing suitable comparison with the other injec-
tor configurations. Therefore, a set of tests with increased feeding
pressure has been carried out (#PSW-3 to PSW-6). In addition to
the resulting increased ṁox , this set of tests enables us to inves-
tigate the influence of the feeding pressure on the overall motor
performance.
While keeping the same inner diameter of the grain, di , increas- Fig. 21. Thrust and pressures as function of time in firing test #PSW-3.
ing the feeding pressure P tbt results in:

(i) an increase of ṁox and G ox_i ,


(ii) a significant improvement of O / F , from 3.9 to 4.9, which bet-
ter approaches the optimum value of 7.5,
(iii) a slight increase of the average regression rate, roughly 7% in
the test #PSW-5/6 compared to #PSW-1,
(iv) a noticeable increase of the chamber pressure, i.e. around 30%,
(v) an enhancement of the combustion efficiencies, as it increased
by 10%,
(vi) no actual consistent change in the specific impulse and thrust
Fig. 22. Typical exhaust plume of PSW injector in a firing test.
coefficient values.

Fig. 21 illustrates the variation of different pressures and thrust


in the test #PSW-3. Significant oscillations are noticed in the pres-
sure trace in the combustion chamber. The thrust exhibits also
significant oscillations in PSW tests. Because the thrust is linked
directly to the chamber pressure, oscillations in the latter affect
the generated thrust profile. In fact, thrust oscillations are mostly
caused by longitudinal, not transverse, pressure pulsations [32].
Note that the load cell could also be influenced by some hardware-
mounting (assembling) characteristics, because, in Figs. 14 and 18,
the pressure record is stable when thrust oscillations are still ob-
served.
A typical flame from PSW firings is presented in Fig. 22. The
flame’s intensity value from test #PSW-5 is shown in Fig. 23. The
explosion and blow-out during ignition are present as explained
in the previous paragraph (SH test). After the ignition phase, the
flame is noticeably stable, thanks to the pre-heating of the oxidizer
provided by the PSW injector.
The combustion of the paraffin fuel, as in the previous tests,
occurs radially and axially. Fig. 24 shows views of both sides of the
remaining fuel grain (injector and nozzle side). The burnt surface
appears quite smooth. Fig. 23. Intensity value of the exhaust plume of firing test #PSW-5.
402 M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406

Fig. 24. Fuel grain after PSW firing test, a – injector side, and b – nozzle side.

4.4. Test results with VOR injector

The use of a vortex injector produces a helical streamline flow,


and the swirl flow increases the mixing of the fuel [33]. Table 7
reports the summary of the averaged outcomes of VOR firing tests.
Unfortunately, the values of ṁox are significantly higher than
Fig. 25. Thrust and pressures as function of time in firing test #VOR-5.
the ones measured with the other injectors. The theoretical dis-
charge coefficient used for the design of this injector was indeed
underestimated and, consequently, the injector diameter is a bit
too large for an optimal comparison with the other injectors. The
higher value of ṁox implies a higher mass flux and has inevitably
a promoting effect on the regression rate.
Analyzing data from Table 7, ṙ increases. This confirms the
statement that helical streams enhance convective heat transfer to
solid fuel, accounting for the increased rate of fuel consumption,
as reported in [5] and [15].
In all tests, the throat nozzle diameter was 22 mm, except in
tests #VOR-4 and #VOR-5 due to logistic problems during the test
campaign. In these cases, the throat nozzle was 24 mm. No sig-
Fig. 26. Exhaust plume of VOR injector firing test.
nificant influence has been noticed on the global results, except a
slightly lower chamber pressure.
Typical traces of chamber pressure, feeding pressure, pressure
before the injector and thrust are shown in Fig. 25.
Fig. 26 depicts a typical exhaust plume of a VOR firing, which is
quite large due to the large amount of injected oxidizer. There is no
quenching of the exhaust plume, as can be seen in Fig. 27. Only a
small quick decrease of the flame intensity in the beginning is vis-
ible, due to the initially low temperatures of the injected oxidizer.
Because of this short quenching, a sudden and short blow-out oc-
curs before full stabilization of the flame (Fig. 27).
Fig. 28 shows a piece of the remaining fuel grain after a VOR
firing test. The burning surface appears quite smooth. The shown
part of the grain is the part initially located at the end of the com-
bustion chamber. The helical oxidizer jet marks the burning surface
with traces at an angle equal to 24◦ . This angle is smaller than the
design angle of the injector (45◦ ). The reduction of the angle of
the incoming flow is due to the following reasons:

(i) The fuel grain undergoes an axial combustion. Being far from
the injection point, the end of the grain is impacted by an
oxidizer jet with an angle lower than the injection angle.
(ii) The combustion accelerates the axial component of the flow
velocity and therefore straightens the streamlines. Fig. 27. Intensity value of the exhaust plume of firing test #VOR-5.

Table 7
Test results obtained with VOR injector (average values).

Test n◦ P tbt , ṁox , tb , df , G ox_i , G ox , O /F ṙ, I sp , Pc, CF ξ, % at O / F ξ, % at O / F opt


bar g/s s mm g/cm2 s g/cm2 s mm/s s bar
VOR-1 60 531.3 5.19 109.8 75.2 13.8 3.3 7.7 183.5 25.0 1.31 99.1 90.2
VOR-2 60 535.6 5.10 105.0 77.4 15.3 3.6 7.2 176.9 24.8 1.28 92.9 88.9
VOR-3 60 539.5 5.25 107.6 76.3 14.5 3.5 7.4 186.8 24.8 1.41 94.9 89.6
VOR-4 60 557.7 5.29 111.9 78.9 14.1 3.5 7.7 167.3 21.0 1.24 91.7 86.6
VOR-5 60 562.5 5.30 111.1 79.6 14.4 3.4 7.7 182.4 21.6 1.33 94.8 88.5
M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406 403

Fig. 28. Remaining fuel grain after a VOR firing test.

(iii) We may also speculate that the liquid paraffin in the bound-
ary layer changes a little bit the shape of the internal burning
Fig. 29. Average regression rate and O / F with the four injectors.
surface of the fuel grain during cooling.

5. Discussion and comparison

Four N2 O injectors have been designed, manufactured, charac-


terized and tested with firing tests using the 1-kN ULB-HRM motor.
The pattern flows are presented in Fig. 7 and Fig. 8 for both
water and liquid nitrous oxide as duty fluids. Discrepancies on the
stream profiles between water and N2 O injections are noticeable,
in particular in the case of HC injector. This is due to the nature of
N2 O which transforms quickly from liquid to gas in normal atmo-
spheric conditions, resulting in a lower density flow and a reduced
flow vector velocity. Note that, during the combustion, the axial
component of the velocity vector will accelerate further.
For comparison purposes, the firings have been performed with
a similar ṁox . All injectors are quite adapted to the size of the
motor to reach roughly 1-kN thrust (Figs. 14, 18 and 21). This is
less the case for our tested VOR configuration (Fig. 25). The latter
exceeds the 1-kN thrust due to features related with the manufac-
tured VOR injector and the obtained high ṁox .
The average results for firing tests are summarized in Ta-
bles 4–7. Figs. 29 and 30 present a graphical comparison of the
averaged outcomes between the four injectors.
Fig. 30. Average thrust coefficient, ξ at experimental O / F and ξ at optimal O / F
As shown in Fig. 29, the highest regression rate is achieved with with the four injectors.
VOR injector, with an increment of 58% and 23% compared to PSW
and SH, respectively. This is because the VOR provides a centrifu-
gal force to the injected oxidizer and produces a higher convective Moreover, in contrast with Ref. [13], a series of firings were
heat transfer to the solid fuel. This result fairly agrees with many carried out at the university of Brasilia using N2 O/paraffin. They
results in the literature, such as Refs. [31] using N2O/HDPE propel- found an increase of ṙ up to 20% by using a pressure-swirl injec-
lant, [34] using O2 /PMMA propellant, and also [11]. tor compared to direct injection system (showerhead). Our results
The tests #HC also present a high regression rate, whereas, at with PSW show a ṙ lowered by 20% in comparison to SH. The
the same time, it has shown the lowest efficiency and thrust coef- major difference between both motors is the length of the pre-
ficient (Fig. 30). This issue seems to occur because of a major ero- chamber. ULB-HRM motor has roughly the double of the length of
sion of the fuel grain due to strong recirculation and impingement the one used at UnB, which helps to lower the vector velocity of
of the liquid N2 O on the rather soft paraffin fuel. Consequently, un- the incoming gaseous oxidizer. Knuth et al. wrote in [35] that “in-
burned paraffin fuel is ejected through the motor’s nozzle without jector mass velocity had a strongly positive effect on the regression
proper combustion, accounting for an increased regression rate but rates”. So, it seems that pressure-swirl injectors require a short dis-
keeping the specific impulse low. tance between injector and inlet of the fuel grain. It is because the
It appears that the average regression rates measured are a bit spray provided by PSW is finely atomized (ours reached a value of
bigger than those found in [11] and [13]. In addition to Ref. [28] 40 μm as described in [24]).
where they used LOX, except in tests #4L-06 and #4L-07, the val- The oxidizer-to-fuel ratios (Fig. 29), in general for all tested in-
ues are close. jectors, are still lower than the optimum value which is worth
404 M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406

Fig. 31. Injector after typical firing test.

around 7.5. The highest one is achieved with PSW due to its rela-
tively low regression rate, and it represents around 60% of O / F opt .
Two different combustion efficiencies are given in Fig. 30:

∗ corresponding to O / F .
– First efficiency is calculated using a cth
It has shown good values, higher than 90%, except for PSW
which equals around 80% due to high O / F shown by PSW, Fig. 32. Theoretical and experimental I sp in function of O / F .
thus a high corresponding cth ∗ .

– The second efficiency of the motor calculated at O / F opt for all


injectors varies from 80 to 90%, with higher values shown with
VOR. This is due to the increase of the residual time conducted
by the helical streamline of VOR flow. It agrees with the con-
clusions of [34] and the relatively high mass flux used in VOR
tests helps to enhance the efficiency. The range of efficiency of
the different injectors using N2 O/paraffin found in this work is
in line with a motor tested using LOX/paraffin in [28].

Fig. 31 shows a typical injector plate after a firing test. Fuel is


transported back to the injection area through the flow recircula-
tion of the combustion port and deposited on the injector plate,
so this unburned fuel does not contribute to the combustion effi-
ciency.
In Fig. 32 two theoretical specific impulses are presented. The
first one is calculated at 25 bar corresponding to the average
chamber pressure of the VOR injector tests. The second one is cal-
culated at 17.5 bar corresponding to the average chamber pressure
of SH, HC and PSW injector tests.
VOR, SH, and PSW give higher specific impulse comparatively
to the one measured with HC. The measured values of specific im-
pulse in tests #PSW-4 and #VOR-3 represent around 78% of the Fig. 33. Exhaust plume intensity developed during combustion.
theoretical maximum I sp .
We suppose that the high observed average regression rate and
the low observed specific impulse (especially with HC injector) are quenching, a sudden and short blow-out occurs before full sta-
partly due to the relatively large amount of grain chunks exhaust- bilization of the flame due to the pre-heating allowed by the
ing the nozzle. Indeed, as a consequence of the poor mechanical PSW injector.
properties of the fuel grain, fuel fragment loss can occur [4]. We – In VOR signal, the preheating allowed by the helical flow and
may speculate that fuel fragment loss is influenced by the injector the relatively high oxidizer mass flow rate prevent the blow
orifice diameter (i.e. incoming oxidizer droplet sizes which attack out during combustion. This is the best flame behavior ob-
fuel surface). Larger diameter could lead to significant effects. To tained.
confirm this proposal, more experimental investigations in simi- – In HC signal, it is clear that the exhaust plume is small, and
lar operating conditions are needed, with smaller injector orifices. the flame blow-out is absolutely absent during steady phase.
Here, VOR provides the biggest droplets through an exit orifice of All injectors present a significant explosion during ignition ex-
3 mm. SH injector has a 1.4 mm exit orifice diameter. Furthermore, cept the HC.
in terms of droplet size, an improvement of VOR and SH configura- – In SH signal, a remarkable blow-out occurs. As explained in
tions should be considered to better estimate the actual regression section 4.1, this is due to the high axial velocity produced by
rate and combustion efficiency. the SH shape that propels the gaseous oxidizer through the
Fig. 33 brings together the traces representing the exhaust nozzle, reduces the residence times of the latter in the com-
plume intensity generated by the motor for different firing tests bustion chamber, prevents appropriate mixing with the evap-
with the four injectors: orating fuel and eventually exits the motor without proper
reaction/combustion.
– In PSW signal, only a small quick decrease of the flame in-
tensity in the beginning is visible, due to the initially low Finally, in terms of instability, Fig. 34 shows the combustion
temperatures of the injected oxidizer. Because of this short chamber pressure fluctuations for typical firing tests. Substantial
M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406 405

terns due to the difference of liquid nitrous oxide characteristics


compared to water.
In general, the average regression rate measurements appear
higher compared to the open literature. The highest regression rate
is achieved using the VOR injector. It exhibits an increment of ṙ up
to 58% and 23% compared to PSW and SH, respectively. The tests
of PSW deliver the lowest regression rate, 20% less in comparison
to SH. It is in contrast with the results given in Ref. [13]. Injection
through PSW gives a good atomized spray, and therefore a low
vector velocity of the gaseous oxidizer. Using a long pre-chamber
(like in our application where it equals 100 mm) helps to lower
the incoming oxidizer velocity; it causes a poor convective heat
transfer, which induces a decrease of the regression rate. When
the feeding pressure is increased, tests performed with PSW pro-
vide some improvements. A 30% increase of the feeding pressure
leads to an increase of the average regression rate by roughly 7%
and of the combustion efficiencies by 10%.
Two types of efficiency were specified and investigated:

– The efficiency at O / F is quite high: around 95% on average for


Fig. 34. Combustion chamber pressure evolution in stable working regime. the injectors SH, HC, and VOR. However, for the injector PSW,
the efficiency is only around 80%.
fluctuations are noticed with the PSW signal compared with the – The range of efficiency at O / F opt of the tested hybrid motor
other injectors, followed by VOR, and more stable signals are varies from 80 to 90%. It appears that VOR tests tend to gen-
recorded using SH and HC injectors with only a tiny difference be- erate higher efficiency. This is sustained by two factors: vortex
tween these three ones. The pressure oscillation in the combustion movement of the flow which increases the residual time, and
chamber observed using PWS injector is due to the characteristics the motor that operates at a higher mass flux when using VOR
of the spray and the oxidizer average velocity in the pre-chamber. injector.
In addition, the large pre-chamber length introduces some fluctu-
ations in the pressure signal as well. The increase of the residence A phenomenon of explosion (hard ignition) during ignition, fol-
time potentially generates combustion instability, as reported in lowed by a flame blow-out, hence a re-explosion, occurred at the
Ref. [36]. VOR injector has large orifices and provides relatively big stout of all firings. In addition, for SH tests, a series of flame blow-
droplets, which increases the time lag of vaporization, accounting outs have been seen during the first seconds of the combustion.
for the associated oscillations in the chamber pressure. Under these conditions of motor testing, PSW create substantial
It is important to notice that, after investigation of the recorded fluctuations of chamber pressure compared to all other injectors,
chamber pressure oscillations, the amplitude of the oscillations followed by VOR. The most stable signals are recorded using SH
over the mean value of the chamber pressure was less than 5% at and HC.
bulk mode (low frequency) for all tests [37]. These results demon- It is suggested that further research should be conducted on
strate the stability in operation of the ULB hybrid motor for the droplet size of the incoming jet. Development of SH and VOR in-
presented injectors and chamber pressure. jectors with smaller orifices is suitable too. Note that with the VOR,
The instability, resulting from the coupling with feed system, “small orifices” means “more orifices”, in order to keep the same
is different from the intrinsic hybrid low frequency instability and oxidizer mass flow rate, and therefore the swirl number increases
just occurs in motors that use liquid oxidizer. The mechanism of which is favorable for vortex injection. And an increase of the angle
the feeding system coupled instabilities and techniques to over- design of the VOR injector to have a radial injection. An investiga-
come the problem will not be discussed in this work, but it is tion of the influence of the distance between PSW and inlet of the
currently under investigation by the UnB/ULB team. fuel grain is also needed to confirm previous statements on PSW
results.
6. Conclusions
Conflict of interest statement
In this paper, a series of firing tests is presented. They investi-
gate the influence of the liquid oxidizer flow pattern on the overall None declared.
performance of hybrid rocket combustion. The flow characteristics
were changed using four types of injectors: showerhead, hollow- Acknowledgements
cone, pressure-swirl, and vortex. The tests were carried out using
the ULB-HRM test bench developed in collaboration between Uni- M. Bouziane would like to express his gratitude to the Algerian
versité Libre de Bruxelles (ULB) and Royal Military Academy of Ministry of Defense for the scholarship of his PhD student posi-
Belgium (RMA). tion at RMA. The authors gratefully acknowledge the support of
First of all, this study helps to enrich the literature with exper- the Beauvechain Air Base military staff allowing the use of the fa-
imental measurements of N2 O/paraffin firings, which are relatively cilities to perform the firing tests. They thank also the technicians
rare in the open literature. Second, it provides an experimental of the ATM department of ULB and the Chemistry department of
characterization of novel types of injectors in hybrid rocket propul- RMA.
sion application. In addition, the study gives a reliable compari-
son of different types of injectors’ performance as they have been References
tested in the same motor configuration.
The cold tests show that some kind of discrepancies may be [1] J.-Y. Lestrade, J. Anthoine, G. Lavergne, Liquefying fuel regression rate modeling
observed, depending on the type of the injector, in the flow pat- in hybrid propulsion, Aerosp. Sci. Technol. 42 (2015) 80–87.
406 M. Bouziane et al. / Aerospace Science and Technology 89 (2019) 392–406

[2] G. Cai, C. Li, S. Zhao, H. Tian, Transient analysis on ignition process of catalytic [20] M. Invigorito, G. Elia, M. Panelli, An improved approach for hybrid RocNet in-
hybrid rocket motor, Aerosp. Sci. Technol. 67 (2017) 366–377. jection system design, World Acad. Sci. Eng. Technol. Int. J. Mech. Aerospace,
[3] G.P. Sutton, O. Biblarz, Rocket Propulsion Elements, John Wiley & Sons, 2016. Ind. Mechatron. Manuf. Eng. 10 (4) (2016) 686–696.
[4] C. Carmicino, F. Scaramuzzino, A.R. Sorge, Trade-off between paraffin-based and [21] E. Doran, J. Dyer, K. Lohner, Z. Dunn, B. Cantwell, G. Zilliac, Nitrous ox-
aluminium-loaded HTPB fuels to improve performance of hybrid rocket fed ide hybrid rocket motor fuel regression rate characterization, in: 43rd
with N2 O, Aerosp. Sci. Technol. 37 (2014) 81–92. AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, 2007, p. 5352.
[5] C. Carmicino, A.R. Sorge, Performance comparison between two different injec- [22] A.H. Lefebvre, V.G. McDonell, Atomization and Sprays, CRC Press, 2017.
tor configurations in a hybrid rocket, Aerosp. Sci. Technol. 11 (1) (2007) 61–67. [23] H. Silva Couto, P.T. Lacava, D. Bastos-Netto, A.P. Pimenta, Experimental evalua-
[6] C. Carmicino, A.R. Sorge, Role of injection in hybrid rockets regression rate be- tion of a low pressure-swirl atomizer applied engineering design procedure, J.
haviour, J. Propuls. Power 21 (4) (2005) 606–612. Propuls. Power 25 (2) (2009) 358–364.
[7] W. Knuth, D. Gramer, M. Chiaverini, J. Sauer, Development and testing [24] M. Bouziane, A.E. de Morais Bertoldi, D. Lee, P. Milova, P. Hendrick, M. Lefeb-
of a vortex-driven, high-regression rate hybrid rocket engine, in: 34th vre, Design and experimental evaluation of liquid oxidizer injection system for
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, 1998, p. 3507. hybrid rocket motors, in: 2017 EUCASS Conference, 2017, p. 133.
[8] W. Knuth, M. Chiaverini, D. Gramer, J. Sauer, Experimental investigation [25] M. Badami, V. Bevilacqua, F. Millo, M. Chiodi, M. Bargende, GDI swirl injector
of a vortex-driven high-regression rate hybrid rocket engine, in: 34th spray simulation: a combined phenomenological-CFD approach, 2004.
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, 1998, p. 3348. [26] G. Cai, P. Zeng, X. Li, H. Tian, N. Yu, Scale effect of fuel regression rate in hybrid
[9] W. Knuth, D. Gramer, M. Chiaverini, J. Sauer, R. Whitesides, R. Dill, Prelimi- rocket motor, Aerosp. Sci. Technol. 24 (1) (2013) 141–146.
nary CFD analysis of the vortex hybrid rocket chamber and nozzle flow field, [27] D.R. Greatrix, Regression rate estimation for standard-flow hybrid rocket en-
in: 34th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, 1998, gines, Aerosp. Sci. Technol. 13 (7) (2009) 358–363.
p. 3351.
[28] A. Karabeyoglu, G. Zilliac, B.J. Cantwell, S. DeZilwa, P. Castellucci, Scale-up tests
[10] T. Takashi, S. Yuasa, K. Yamamoto, Effects of swirling oxidizer flow on fuel
of high regression rate paraffin-based hybrid rocket fuels, J. Propuls. Power
regression rate of hybrid rockets, in: 35th Joint Propulsion Conference and Ex-
20 (6) (2004) 1037–1045.
hibit, 1999, p. 2323.
[29] “EXPLO5”, Retrieved from http://www.ozm.cz/en/explo-5-software/. (Ac-
[11] N. Bellomo, F. Barato, M. Faenza, M. Lazzarin, A. Bettella, D. Pavarin, Numeri-
cessed 23 July 2018).
cal and experimental investigation of unidirectional vortex injection in hybrid
[30] T. Boardman, D. Brinton, R. Carpenter, T. Zoladz, An experimental investigation
rocket engines, J. Propuls. Power 29 (5) (2013) 1097–1113.
of pressure oscillations and their suppression in subscale hybrid rocket motors,
[12] C.P. Kumar, A. Kumar, Effect of swirl on the regression rate in hybrid rocket
in: 31st Joint Propulsion Conference and Exhibit, 1995, p. 2689.
motors, Aerosp. Sci. Technol. 29 (1) (2013) 92–99.
[31] J. Pucci, The effects of swirl injector design on hybrid flame-holding combus-
[13] A.E. de Morais Bertoldi, C.A.G. Veras, P. Hendrick, Experimental evaluation of
tion instability, in: 38th AIAA/ASME/SAE/ASEE Joint Propulsion Conference &
pressure-swirl injection system over solid fuel regression rate in hybrid rockets,
Exhibit, 2002, p. 3578.
in: 2017 EUCASS Conference, 2017, p. 661.
[14] M. Bouziane, A.E. De Morais Bertoldi, P. Milova, P. Hendrick, M. Lefebvre, De- [32] Q. Zhang, Z. Wei, W. Su, J. Li, N. Wang, Theoretical modeling and numerical
velopment and testing of a lab-scale test-bench for hybrid rocket engines, in: study for thrust-oscillation characteristics in solid rocket motors, J. Propuls.
2018 SpaceOps Conference, 2018, p. 2722. Power 28 (2) (2012) 312–322.
[15] D. Bianchi, F. Nasuti, C. Carmicino, Hybrid rockets with axial injector: port di- [33] H. Li, L. Ye, X. Wei, T. Li, S. Li, The design and main performance of a hydrogen
ameter effect on fuel regression rate, J. Propuls. Power 32 (1) (2016) 984–996. peroxide/kerosene coaxial-swirl injector in a lab-scale rocket engine, Aerosp.
[16] F. Scaramuzzino, C. Carmicino, G. Festa, A. Viviani, A. Russo, Fuel regression- Sci. Technol. 70 (2017) 636–643.
rate characterization on a lab-scale hybrid rocket burning N2 O and paraffin- [34] S. Yuasa, O. Shimada, T. Imamura, T. Tamura, K. Yamoto, A technique for im-
based propellants, in: 49th AIAA/ASME/SAE/ASEE Joint Propulsion Conference, proving the performance of hybrid rocket engines, in: 35th Joint Propulsion
2013, p. 4039. Conference and Exhibit, 1999, p. 2322.
[17] D. Bianchi, F. Nasuti, C. Carmicino, Numerical analysis of port diameter ef- [35] W.H. Knuth, M.J. Chiaverini, J.A. Sauer, D.J. Gramer, Solid-fuel regression rate
fect on hybrid rocket fuel regression rate with axial injection, in: 51st behavior of vortex hybrid rocket engines, J. Propuls. Power 18 (3) (2002)
AIAA/SAE/ASEE Joint Propulsion Conference, 2015, p. 3835. 600–609.
[18] C. Carmicino, D. Pastrone, Novel comprehensive technique for hybrid rocket [36] B. Greiner, R. Frederick Jr., Hybrid rocket instability, in: 29th Joint Propulsion
experimental ballistic data reconstruction, J. Propuls. Power 34 (1) (2017) Conference and Exhibit, 1993, p. 2553.
133–145. [37] A.E.M. Bertoldi, Combustion Instability in Hybrid Rocket Propulsion Systems,
[19] E. Toson, A.M. Karabeyoglu, Design and optimization of hybrid propulsion sys- PhD Dissertation, Mechanical Engineering Department, University of Brasília,
tems for in-space application, in: 51st AIAA/SAE/ASEE Joint Propulsion Confer- Brasília, DF, Brazil, July 2018 (in Portuguese).
ence, 2015, p. 3937.

You might also like