You are on page 1of 18

AIAA Propulsion and Energy Forum 10.2514/6.

2020-3763
August 24-28, 2020, VIRTUAL EVENT
AIAA Propulsion and Energy 2020 Forum

CFD Simulation of a 1kN Paraffin-Fueled Hybrid


Rocket Engine

B. Dequick∗ , M. Lefebvre†
Royal Military Academy, Avenue de la Renaissance 30, Brussels, 1000, Belgium

P. Hendrick‡
Université Libre de Bruxelles, Avenue F.D. Roosevelt 50, Brussels, 1050, Belgium

Hybrid rocket engines (HREs) are rocket engines that combine a solid fuel and a liquid
(or gaseous) oxidizer, or vice versa. A setup with a solid fuel is most commonly used. For
some years now, researchers at Université Libre de Bruxelles (ULB) are running tests with a
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

test bench HRE. This engine has a target thrust of 1 kN and uses a paraffin fuel with liquid
N2 O as oxidizer. A first computational fluid dynamics (CFD) simulation of this engine is
presented here. It is a steady-state RANS simulation, using the standard 𝑘-𝜖 turbulence model
together with the Eddy Dissipation Model (EDM) for the turbulence chemistry interaction
(TCI). Oxidizer and fuel enter the domain in a gaseous phase. An ambient area is included in
the computational domain as well. This avoids imposing a boundary condition at the nozzle
exit section, and therefore a more correct pressure profile can be formed. Furthermore, it
allows to compare and investigate the exhaust plume structure. The results are compared with
test firing data and show an average deviation of less than 10%. Also, a parametric study is
done regarding the oxidizer and fuel inlet temperatures and the oxidizer to fuel mass ratio (O/F
ratio). Using the CFD model, future work will aim on improving the engine’s performance by
focusing on the design influence of the post combustion chamber and nozzle.

I. Introduction

A. Hybrid Rocket Engines


Rocket propulsion systems can be classified in many ways. In the specific field of chemical rocket propulsion,
characterized by the presence of a fuel and an oxidizer which react together at some point, it is most common to distinct
between three main categories: solid, liquid and hybrid rockets [1]. The first one has a solid grain in which a fuel and an
oxidizer are premixed at molecular level, held together by a binder. In the second type, the fuel and the oxidizer are
liquid and stored completely separately. For the hybrid case, the fuel and oxidizer are stored separately as well, but in
different phases. Usually, the fuel is solid and the oxidizer is liquid or gaseous.
For both solid and liquid engine types, research is in a mature state and these types of engines can be found in space
launch vehicles and a wide range of other commercial applications. This is not yet the case for hybrid engines, but they
are promising because of a series of advantages compared to the solid or liquid engine types, mainly related to safety
and costs [2].
The difficulty with Hybrid Rocket Engines (HREs) is to generate enough thrust, which is needed for most applications.
For this, the solid fuel regression rate or burning surface should be increased. The latter can be obtained with a
multi-port configuration, but this has several drawbacks such as the negative effect on structural strength and fuel
efficiency. Increasing the regression rate is therefore the preferred way to increase thrust. This allows for a simple and
robust single port fuel grain to be used.
For more than a decade now, paraffin-based fuels have proven to offer much higher regression rates than, for instance,
HTPB fuels (Hydroxyl-Terminated PolyButadiene). But these paraffin-based fuels give rise to a new challenge, being the
particularity of the mechanics by which they burn. The formation of a liquid film at the fuel surface and the entrainment
of droplets are typical and key to the increased regression rate [3], but it makes the understanding and modeling of the
internal ballistics more difficult.
∗ PhD Candidate
† Professor
‡ Professor

Copyright © 2020 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
A recent and detailed overview of the state of the art and current challenges with respect to HREs in which liquefying
fuels such as paraffin-based fuels are used, is provided by [4].

B. Computational Fluid Dynamics


Fluid dynamics research can be theoretical (analytical), experimental or numerical. In the numerical approach, the
partial differential equations (PDEs) that describe the fluid flow are approximated by a system of algebraic equations
which can be solved on a computer. Hence the name Computational Fluid Dynamics (CFD). To obtain the system of
algebraic equations, a discretization method must be chosen, by which the fluid domain or computational domain is
divided in discrete elements [5].
As for other applications, CFD has gained importance in HRE research because of the continuous improvement and
accessibility of computational power. CFD simulations could help indicate possible improvements of HRE performances
and their behavior. However, in the specific case of paraffin-fueled HREs, CFD models are hard to develop because of
the complexity of physical phenomena occurring in these engines, as mentioned in the previous section. Depending on
the purpose, CFD simulations of paraffin-fueled HREs still include some assumptions and simplifications. This also
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

applies for the study presented in the current paper. Details are provided in section III.
A recent and detailed overview of the state of the art of CFD techniques for HREs with classical and liquefying fuels
such as paraffin-based fuels is provided by [6]. It is clear that improved CFD modeling, including model validation
using experimental data, is necessary.

II. Context and purpose of current study

A. 1 kN Hybrid Rocket Engine at Université Libre de Bruxelles


The numerical study presented in this paper is based on the lab-scale HRE at Université Libre de Bruxelles (ULB),
and it will be referred to as the ULB-HRE in the remainder of the text.

1. Latest design
The ULB-HRE was first developed in 2010 and has meanwhile undergone several modifications. The engine is
designed to have a theoretical thrust of 1 kN and a burning time of 10 s. The latest design is shown in Fig. 1 [7], on
which four typical parts of a HRE can be recognized from left to right: the pre-chamber with the oxidizer injector and
ignition cartridge, the combustion chamber with a single port fuel grain, the post-chamber, and the nozzle.

Fig. 1 3D cut of the ULB-HRE, indicating its main parts [7].

2
The dimensions were obtained based on classical internal ballistics theory for hybrid propulsion. Details are
presented in [7]. The total internal axial length measured from the injector plate to the nozzle exit is ≈360 mm. At
the pre-chamber and post-chamber, the internal diameter is at its maximum of ≈130 mm. The fuel grain consists
out of pure paraffin and has a single port circular configuration. Its total length is a relatively short 110 mm and the
initial port diameter varies between tests, ranging from 20 to 50 mm. The nozzle throat and exit diameter are 22 and
50 mm, respectively. The oxidizer is liquid N2 O, which is fed through the injector plate after which it atomizes in the
pre-chamber.

2. Previous experimental work


In recent experimental work, different injector flow patterns have been studied: shower head, hollow-cone, pressure-
swirl and vortex injector [8]. It becomes clear that the engine’s performance is strongly affected by the injector
configuration [9, 10]. During operation and depending on the chosen configuration, the chamber pressure of the
ULB-HRE varies between 15 and 25 bar, and the oxidizer pressure before the injector plate varies between 30 and
55 bar. The burning time is controlled and usually lies within the range of 4 to 9 s. The specific impulse (Isp ) ranges
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

from 145 to 180 s.

3. Previous CFD work


One of the challenges in the CFD modeling of the ULB-HRE is to incorporate some additional physical phenomena,
inherent to the use of paraffin as fuel, compared to more classic non-liquefying solid fuels. In this scope, important CFD
work has been done by Milova, Blanchard, and Galfetti [11]. The formation of a liquid fuel layer, and consequently
the burning of fuel droplets formed by the entrainment effect, were studied. This was done numerically by injecting
gaseous C2 H4 and droplets of liquid paraffin into the computational domain. The focus lied on the numerical model
development itself, not taking into account any of the actual ULB-HRE design parameters.

B. Main purpose
The research conducted with the ULB-HRE has mainly been experimental so far. The development of a numerical
model of the entire engine is therefore desirable, such that it adds a predictive capacity to the research if it proves to be
sufficiently reliable.
The main purpose of the current paper is to establish a first CFD model of the ULB-HRE. The available experimental
results are used for comparison purposes. This allows to evaluate the accuracy of the model.
Future work will focus on the post-combustion chamber and nozzle design influence, in order to propose potential
engine performance improvements.

III. CFD model

A. Computational domain
The computational domain is 2D axisymmetric with respect to the main flow direction and includes the interior of
the engine and the exhaust plume region. The dimensions of the latter are based on the general size of the observed
plume during firing tests. The simulations are steady-state and so the fuel port diameter is fixed at 70 mm, which
represents a typical intermediate value during engine operation. Ignition features are not included, as this is not of
interest in the current work. The geometry of the circular plate and bolts which keep the nozzle in place are ignored as
well, since their impact on the exhaust plume structure is assumed to be negligible. Figure 2 shows an overview.

B. Mesh
A series of different meshes were created in order to study mesh convergence and the influence of inflation layer
cells near the boundaries. All meshes are structured-like and have quadrilateral cells. Triangular cells do appear in
some areas, but they represent less than 0.02% in number.

3
Fig. 2 Computational domain
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

1. Inflation layer cells


From the chosen modeling approach, as yet to be explained in detail in section III.C.2, it follows that a so called
near-wall treatment (NWT) must be selected to account for the viscous sublayer which exists near the wall [12]. A
NWT requires a certain near wall mesh resolution, characterized by the scaled distance from the walls to their adjacent
cell centroids "P" [13]. This scaled distance is denoted by 𝑦 + or by 𝑦★, depending on the definition. In the current
model, the latter is used and it is defined as follows.

𝜌𝐶 𝜇0.25 𝑘 0.5
𝑃 𝑦𝑃
𝑦★ ≡ (1)
𝜇
Here, 𝜌 and 𝜇 are the fluid density and dynamic viscosity, 𝑘 𝑃 is the turbulence kinetic energy in P, and 𝑦 𝑃 is the
actual distance from P to the wall. 𝐶 𝜇 is a constant equal to 0.09.
As said before, NWTs generally require 𝑦★ to be in a certain range. Depending on the NWT, 𝑦★ should typically be
less than 1 or greater than 30. However, the NWT that will be chosen here (see III.C.2) is claimed to be insensitive to
the values of 𝑦★. If this is correct, results should not depend on the thickness (within reasonable limits) of the cell
layers near the walls. In the current paper, these results are represented by the obtained global operating values such
as chamber pressure and nozzle exhaust speed. They depend less on what is happening in the near-wall region than
parameters such as drag coefficients. Therefore it is predicted, especially with the chosen insensitive NWT, that the
operating parameters are not strongly affected by the near-wall mesh resolution.
This will now be checked by comparing the results obtained with 3 meshes, all with equal cell sizes of ≈ 1 mm.
The only difference between these meshes are the values of 𝑦★ and the number of cell layers in the near-wall regions.
Figure 3 shows a top-left corner of each mesh to visualize the inflation layers.

(a) No refinement (b) Low refinement (c) High refinement

Fig. 3 Top-left corner of three meshes with different levels of near-wall refinement.

The 3 simulations are done with boundary conditions that fall within the range of those that will be used for the
actual simulations of the experiments which are discussed in section IV.
The results are compared in two ways. First, by looking at the resulting global operating values. They are summarized
together with some mesh properties in Table 1. Next, by looking at the obtained field values at two sections of the engine:
a section in the post-combustion chamber and the nozzle exit section. The obtained field values are shown in Fig. 4.

4
Table 1 Comparison of global operating parameters obtained with 3 meshes with different levels of near-wall
refinement.

Mesh Near-wall Number of cells 𝑦★ 𝑃post-ch 𝑇 post-ch 𝑣 nozzle exit 𝐹 𝐼sp


mesh refinement [bar] [K] [m/s] [N] [s]
1 None 82774 182 22.31 1298 1756 1316 192
2 Low 90284 36 22.21 1317 1749 1312 191
3 High 135987 0.42 22.18 1300 1744 1310 191
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

(a) (b) (c)

(d) (e) (f)

Fig. 4 Profile of axial velocity, temperature and static pressure along a section in the post-chamber and along
the nozzle exit section, for 3 levels of near-wall mesh refinement. The legend shown in (a) is valid for all charts.

5
From Table 1 it can be concluded that overall differences for typical operating parameters between the three
meshes are small, especially between mesh 2 and mesh 3, where differences between 𝑃post-ch , 𝑇 post-ch and 𝑣 nozzle exit are
respectively 0.14%, 1.3%, and 0.29%. This is confirmed in Fig. 4, from which it is clear that mesh 2 and mesh 3 result
in very similar field value profiles.
Therefore, in the context of the current study and modeling approach presented later in section III.C.2, a moderate
near-wall mesh refinement is considered to be sufficient.

2. Mesh convergence
Different series of meshes were created during the initial mesh design process. The differences between these series
are essentially related to the near-wall mesh refinement (discussed in section III.B.1). Within each series, meshes only
differ in global cell size.
In order to obtain mesh-independent results, mesh convergence must be reached. This means refining the mesh cell
sizes until there is no significant change in the solution when all other parameters are kept unchanged. Within each
series, consecutive mesh refinements were done and the resulting engine operating values were monitored.
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

Figure 5 shows the evolution of the chamber pressure for 5 series of meshes, all with a single representative set
of boundary conditions. Initially, a first order differencing scheme was used for the turbulence field values (𝑘 and
𝜖). For these mesh series (shown in gray), a high overall refinement was necessary to observe some convergence just
above 22 bar. On the other hand, when using second order differencing for the turbulence field values (series shown in
black), the chamber pressure always falls within the range of 22 to 23 bar, with a general trend towards just above 22 bar
for finer meshes. Any fluctuations are solely related to the mesh, since all simulations have converged perfectly, as
explained in section III.C.4.
Based on these observations, a global mesh size of 1 mm × 1 mm is considered to be sufficient when using second
order differencing for all field values. This mesh resolution provides results that differ only 1% from those of the
0.4 mm × 0.4 mm mesh (for series with near-wall mesh refinement). The final mesh used for the CFD simulations in
section IV is shown in Fig. 6. It has 120602 cells.

Chamber pressure Mesh convergence


[bar] for chamber pressure
24 Series Differencing Near-wall
scheme for mesh
A
turbulent field refinement
23 values
B
C A 2nd order none
22 B 2nd order low
C 2nd order high
D 1st order high
21 E 1st order none

D
20 E

19

18 Cell edge size [mm]


0.1 1

Fig. 5 Mesh convergence for chamber pressure for 5 series of meshes, each with increasing global refinement
from right to left. The impact of the differencing scheme for the turbulence field values becomes clear.

6
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

Fig. 6 Mesh used for the simulations in section IV.

C. Flow modeling

1. Flow aspects of paraffin-fueled HREs


The combustion in HREs is a non-premixed combustion. Fundamental work in understanding the involved physics
has been done by Marxman, Wooldridge, and Muzzy [14], describing the boundary layer combustion model for classical
HREs. The injected oxidizer flows through a fuel port in which, after ignition, a diffusion-limited flame is formed in a
relatively wide turbulent boundary layer. This has already been visualized by experiments in 1963 [15]. The flame is fed
by the axial oxidizer inflow and the radial vaporized fuel inflow. Convective and radiative heat transfer from the flame to
the grain surface causes the fuel to sublimate into gasses. The mass flow of these fuel gasses in turn block a part of the
heat transfer (blowing or blocking effect), thus creating some equilibrium. Furthermore, it has been shown that the
regression rate of the fuel grain mainly depends on the oxidizer flux through the fuel port. As the fuel grain is consumed,
the port diameter increases. This leads to a decreased oxidizer flux (assuming constant oxidizer mass flow rate) and
an increased fuel surface. Both effects do not entirely compensate each other, resulting in an O/F ratio shift during
operation. An overview of the numerous physicochemical processes occurring in classical HREs is provided by [16].
In the special case of liquefying fuels such as paraffin, where a liquid film is formed on the grain surface, important
work has been done by Karabeyoglu, Cantwell, and Altman [3]. Here, the solid fuel first melts and, after a cracking and
pyrolysis process, enters the main flow as low carbon number gasses. Because of liquid film instabilities such as roll
waves, a part of the fuel can enter the main flow as entrained droplets which do not contribute to the blocking effect.
This in turn leads to higher regression rates compared to non liquefying fuels. Note that the regression rate, for any
HRE, is usually not uniform and depends on the axial coordinate.
Other important aspects of the internal ballistics relate to the oxidizer injection system, such as the injector flow
pattern, the oxidizer phase, and droplet size distribution in case of a spray. These can all affect recirculation zones,
impingement zones of oxidizer droplets on the fuel surface, the amount of unreacted oxidizer exiting the engine, and
thus the overall engine performance. Examples of experimental and CFD studies on the oxidizer injection in HREs
include [9] and [17, 18], respectively.

2. Modeling strategy and setup


A figure is provided at the end of this section in which the essential elements of the flow modeling setup are
summarized.
In order to model turbulent flows such as in the ULB-HRE, one could consider three strategies. They correspond to
Direct Numerical Simulation (DNS), Large-Eddy Simulation (LES) or Reynolds-Averaged Navier-Stokes (RANS) [19].
The last will be discussed further, as it is the preferred approach for this work.

7
RANS is the most common way by which the turbulent flows inside HREs are modeled today [4, 6]. In the associated
equations, any fluctuations caused by turbulence are time averaged and they appear as an extra stress term which needs to
be modeled by some turbulence model. This is why RANS simulations provide a steady-state solution which represents
an average flow (when boundary conditions are constant in time). Running a series of steady RANS simulations can
represent snapshots of an evolving flow, such as in a HRE with increasing port diameter during operation. Some
examples of such a quasi-steady approach include [6, 18, 20] and many more. In this paper however, only one fuel port
diameter is considered to establish the first CFD model of the ULB-HRE. Many researches have obtained representative
results by using RANS to model paraffin-fueled HREs. Recent examples include [6, 21–24]. The computational cost is
low compared to LES and DNS, and a 3D mesh is not required. Instead, equations can be solved in a 2D axisymmetric
domain. With this in mind, it seems that the RANS approach is a good choice to establish a first CFD model of the
ULB-HRE.
As said before, the RANS equations need to be closed by some turbulence model. Commonly used models are the
𝑘-𝜖 [25], 𝑘-𝜔 [26], 𝑘-𝜔 SST [27] and Spalart-Allmaras [28] models. For now, the model used for this study is the 𝑘-𝜖
model, as the use of other models do not yet provide realistic results. In some cases, no diffusion flame is formed at all
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

in the (relatively short) combustion chamber. Along with a turbulence model, a proper wall treatment must be chosen
to take into account the viscous sublayer in which turbulence effects are blocked because of the presence of the wall.
The Menter-Lechner near-wall treatment is used here, as it is presented as a 𝑦★-independent model by the software
developer [13].
Next, a combustion model has to be chosen. Two common approaches are based on which factor is considered to
limit the chemical reaction rate: the reaction kinetics or the turbulent mixing. If the mixing of the reactants is considered
fast, then the reaction progress is limited by some finite-rate reaction mechanism, described by Arrhenius laws. In this
case, reaction progress depends on local averaged field values and does not take into account any turbulence-chemistry
interaction (TCI). On the other hand, if the chemical reactions are considered fast compared to the transport phenomena,
then the reaction progress is limited by the turbulent mixing. One way to implement this TCI is the Eddy Dissipation
Model (EDM) [29], in which the different species source terms are calculated from the turbulent kinetic energy, the eddy
dissipation rate and the species mass fractions. This implies that transport equations for each species must be solved,
but it avoids including complex chemical kinetics. The EDM is typically suited for a one-equation global reaction, as
the reaction rates for multiple reactions would all be based on the same turbulence field values, which would lead to
incorrect results [13]. The model, which is of the type mixed-is-burnt, is one of the popular choices for simulating the
turbulent non-premixed combustion in HREs. Recent examples include [21, 23, 30]. The EDM will be applied in the
current study as well. An alternative way to model combustion with TCI, thus also assuming fast chemistry, would be
the mixture fraction approach [31]. In this case, no transport equations for the different species are solved. Instead,
transport equations for the averaged mixture fraction 𝑓 = 𝑓 (𝑂/𝐹) and it’s variance 𝑓 02 are solved. By assuming a
Probability Density Function (PDF) coupled with chemical equilibrium, a look-up table can be constructed from which
the temperature and species mass fractions are obtained. Examples include [6, 11, 18, 32]. Like the EDM, it is also a
mixed-is-burnt model.
To model the fuel entering the combustion chamber, a fuel mass source is added in a very thin layer of cells adjacent
to the axial grain wall, all at a fixed temperature of 740 K, which is the average of Tboiling and Tcritical of paraffin. The
grain sides are assumed non-reactive. Fuel mass flows are obtained from experimental data, and thus the regression
rate is not solved and there is no gas-surface interaction (GSI). The low carbon gasses resulting from the pyrolysis of
paraffin fuel consist mainly out of ethylene (C2 H4 ) and a small fraction of hydrogen [33] of about ≈ 0.5 m% when
C28 H58 is considered as the paraffin wax hydrocarbon [11]. As a first approximation, the fuel mass source is set to 100%
of gaseous ethylene. Future work may include the liquid fuel droplet entrainment such as presented in [11]. For a DNS
of the liquid film instabilities the reader is referred to [34].
The inlet oxidizer spray is simulated as gaseous N2 O entering the domain at a specified mass flow rate and
temperature. It is thus assumed that the liquid oxidizer atomizes and vaporizes upon entering the flow domain. This is
a common approach, recent examples include [6, 21, 22, 24, 32]. Because of the 2D axisymmetric domain, all real
inlet nozzle geometries are reduced to a disk or annulus shaped area, although the injection velocity vectors can differ
from the normal direction. In the current study, simulations where the oxidizer is entirely injected as ∅ 30 𝜇m liquid
droplets of N2 O at 280 K, suggest that they evaporate completely in the pre-chamber. These simulations however are not
presented here, as they are preliminary. An example of a CFD model considering such a multiphase oxidizer injection is
presented in [23].

8
The ambient area is bounded by a wall on the left (next to the nozzle exit), a 1 atm pressure inlet (298 K, air) at the
top, and a 1 atm pressure outlet on the right. The pressure inlet is due to the exhaust plume which draws in air from the
sides. With this setup, no reversed flows are present in the converged solution. Furthermore, the presence of an ambient
area allows the formation of a more correct nozzle exit pressure profile.
Heat losses through the motor housing are ignored, and thus, at this stage, all walls are set to be adiabatic except for
the grain surface wall, which is kept at a fixed temperature as mentioned before. Furthermore, a no slip condition is valid
at every wall. Radiation effects are ignored for the sake of simplicity. It is assumed that their contribution is negligible.
Figure 7 shows a summary of this section.
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

Fig. 7 Model overview

3. Equations
This section provides an overview of the fluid equations that apply for the modeling strategy as discussed in section
III.C.2.
Starting from an arbitrary, stationary, and non-deformable control volume, one can derive a general transport
equation for any extensive property 𝜙 (some property per unit mass),
𝜕
(𝜌𝜙) + ∇ · (𝜌®𝑣 𝜙) = ∇ · (Γ∇𝜙) + 𝑆 (2)
𝜕𝑡
which is known as the convection-diffusion equation. From left to right there is a non-steady term, a convective
term, a diffusive term, and a source term. Γ is known as the diffusivity. If the transported property is mass, then 𝜙
equals 1 (mass per unit mass), Γ equals zero, 𝑆 equals zero and Eq. 2 becomes
𝜕𝜌
+ ∇ · (𝜌®𝑣 ) = 0 (3)
𝜕𝑡
which is known as the continuity equation (conservation of mass). If the transported property is momentum, then 𝜙
equals 𝑣® (momentum per unit mass), and Eq. 2 becomes
𝜕
(𝜌®𝑣 ) + ∇ · (𝜌®𝑣 𝑣®) = ∇ · (Γ∇®𝑣 ) + 𝑆 (4)
𝜕𝑡
As surface forces on a fluid element, represented by a stress tensor 𝜏, ¯ relate to diffusion of momentum, and as body
® provide a source of momentum, it can be shown that Eq. 4 becomes
forces (𝑏)
𝜕
(𝜌®𝑣 ) + ∇ · (𝜌®𝑣 𝑣®) = ∇ · 𝜏¯ + 𝜌 𝑏® (5)
𝜕𝑡
which is known as Navier’s equation of equilibrium, valid for all fluids. For Newtonian, isotropic, homogeneous,
and Stokian fluids, 𝜏¯ becomes
 
2
𝜏¯ = −𝑃𝐼 + 𝜇 (∇®𝑣 + ∇®𝑣𝑇 ) − ∇ · 𝑣® 𝐼 (6)
3

9
and in this case Eq. 5 is known as the Navier-Stokes equations, as one equation for each component of 𝑣® can be
written. As explained in section III.C.2, to model turbulence, the Navier-Stokes equation is averaged in time which
leads to an extra stress term. This term is modeled using the Boussinesq Hypotheses. In this case Eq. 5 becomes
 
𝜕 𝑇 2 2
(𝜌®𝑣 ) + ∇ · (𝜌®𝑣 𝑣®) = −∇𝑃 + ∇ · (𝜇 + 𝜇𝑡 ) (∇®𝑣 + ∇®𝑣 − ∇ · 𝑣® 𝐼) + ∇ · (−𝜌𝑘 𝐼) (7)
𝜕𝑡 3 3
in which the turbulent viscosity 𝜇𝑡 and the turbulent kinetic energy 𝑘 are modeled via the standard 𝑘-𝜖 turbulence
model. Therefore, two additional transport equations are solved: one for 𝑘 and one for 𝜖 (turbulent dissipation rate).
 
𝜕 𝜇𝑡
(𝜌𝑘) + ∇ · (𝜌®𝑣 𝑘) = ∇ · (𝜇 + )∇𝑘 + 𝐺 𝑘 + 𝜌𝜖 (8)
𝜕𝑡 𝜎𝑘

𝜖2
 
𝜕 𝜇𝑡 𝜖
(𝜌𝜖) + ∇ · (𝜌®𝑣 𝜖) = ∇ · (𝜇 + )∇𝜖 + 𝐶1𝜖 𝐺 𝑘 + 𝐶2𝜖 𝜌 (9)
𝜕𝑡 𝜎𝜖 𝑘 𝑘
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

in which 𝐺 𝑘 represents the generation of turbulence kinetic energy due to the mean velocity gradients, and 𝜎𝑘 and 𝜎𝜖
are the turbulent Prandtl numbers for 𝑘 and 𝜖, respectively. 𝐶1𝜖 and 𝐶2𝜖 are constants: 𝐶1𝜖 = 1.44 and 𝐶2𝜖 = 1.92.
The turbulent viscosity 𝜇𝑡 is calculated as

𝑘2
𝜇𝑡 = 𝜌𝐶 𝜇 (10)
𝜖
in which 𝐶 𝜇 is a constant equal to 0.09.
Furthermore, the transport equation is also solved for energy, from which the temperature field can be obtained.
Equation 2 then becomes
𝜕
(𝜌ℎ) + ∇ · (𝜌®𝑣 ℎ) = ∇ · (𝜅∇𝑇) + 𝑆 ℎ (11)
𝜕𝑡
in which ℎ is the sensible enthalpy, 𝜅 the thermal conductivity, and 𝑆 ℎ the heat source (for example, due to chemical
reactions).
To model the chemical reactions and the TCI, the EDM has been chosen. As a first approximation at this stage of the
CFD study, a simple global chemical reaction model has been considered:

𝐶2 𝐻4 + 6𝑁2 𝑂 = 2𝐶𝑂 2 + 2𝐻2 𝑂 + 6𝑁2 (12)


The reaction progress is entirely driven by the turbulence field values 𝑘 and 𝜖, as well as the concentrations of each
species. There are six species: five are shown in Eq. 12, and there is also O2 as part of the air in the exhaust plume
area. This means that five additional transport equations as shown below must be solved, namely for each species mass
fraction 𝑌𝑖 , with 𝑖 ∈ [1, 5]. Note that 𝑌6 = 1 − 5𝑖=1 𝑌𝑖 .
Í

 
𝜕 ∇𝑇
(𝜌𝑌𝑖 ) + ∇ · (𝜌®𝑣𝑌𝑖 ) = ∇ · 𝜌𝐷 𝑖,𝑚 ∇𝑌𝑖 + 𝐷 𝑇 ,𝑖 + 𝑅𝑖 (13)
𝜕𝑡 𝑇
Here, 𝐷 𝑖,𝑚 is the mass diffusion coefficient and 𝐷 𝑇 ,𝑖 is the thermal diffusion coefficient for species 𝑖 in the mixture.
𝑅𝑖 is the net rate of production of species 𝑖 by the chemical reaction. For the EDM, 𝑅𝑖 is the smallest of the two
following equations.
 
00 0 𝜖 𝑌𝑅
𝑅𝑖 = (𝜈𝑖 − 𝜈𝑖 )M𝑖 𝐴𝜌 min 0 (14)
𝑘 𝑅 𝜈𝑖 M 𝑅
Í !
00 0 𝜖 𝑃 𝑌𝑃
𝑅𝑖 = (𝜈𝑖 − 𝜈𝑖 )M𝑖 𝐴𝐵𝜌 Í 00 (15)
𝑘 𝑗 𝜈𝑗 M𝑗

Here, 𝜈𝑖0 and 𝜈𝑖00 are the reaction coefficients for reactant 𝑖 and product 𝑖, respectively. M is the molecular weight, 𝑌𝑅
is the mass fraction of any reactant and 𝑌𝑃 is the mass fraction of any product. 𝐴 and 𝐵 are model constants equal to 4
and 0.5 respectively.

10
From basic rocket propulsion theory, the equations for thrust (𝐹) and specific impulse (𝐼sp ) are also of interest.

𝐹 = ( 𝑚𝑣)
¤ exit + (𝑃exit − 𝑃ambient ) 𝐴exit (16)

𝐹
𝐼sp = (17)
𝑚¤ exit 𝑔0
in which the subscript exit relates to the nozzle exit section.

4. Solver and convergence


The selected solver uses a finite volume and pressure based method. A pseudo-transient coupled algorithm is used
for the initial solution, followed by the SIMPLEC algorithm if necessary, as it shows to improve numerical stability. The
under-relaxation factors (URFs) and the time step are altered during the simulations to pursue the best compromise
between convergence speed and stability.
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

In this way, highly converged solutions are obtained, with all residuals (scaled) below 10−6 and a series of monitored
field values leveling out ’completely’. The deviation between mass inflow and mass outflow is always less than 0.001%.

IV. Results

A. Selected experimental results


At this stage of the CFD study, the oxidizer inlet is modeled as gaseous N2 O entering the domain in axial direction
at 280 K, with a constant mass flux over a circular area which has a diameter of 70 mm. It is assumed that this setup
corresponds best to a showerhead (SH) injector. As mentioned in section II.A.2, experimental work has been done with
4 different injector types, but only the results obtained with the SH injectors (also 4 different ones) will be used for
comparison purposes. All SH injectors have orifices spread over a ∅ 70 mm circular area: 11× ∅ 1.4 mm for SH1, 11×
∅ 1.9 mm for SH2, 21× ∅ 1.4 mm for SH3, and 71× ∅ 0.8 mm for SH4. This is shown in Fig. 8. Note that the resulting
N2 O mass flux just downstream of these injectors is actually not uniform in space. Nevertheless, as mentioned just
before, the first CFD model uses a uniform inlet mass flux. In future work, other inlet mass flow profiles can be explored.

Fig. 8 Overview of the 4 types of showerhead injectors used during the experiments.
(a) SH1: 11× ∅ 1.4 mm (b) SH2: 11× ∅ 1.9 mm (c) SH3: 21× ∅ 1.4 mm (d) SH4: 71× ∅ 0.8 mm for SH4
.

The space-time averaged results of the SH experiments are summarized in Table 2. The reader is referred to [35] for
details about how these values were obtained.

11
Table 2 Summary of experimental results with 4 different showerhead injectors, used for comparison with the
CFD results [35].
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

B. Numerical results

1. Compared with experimental data


First, the experimental and numerical operating chamber pressures are compared. On Fig. 9a, these pressures are
plotted on the abscissa and the ordinate respectively, together with the perfect match identity line and its 10% and 20%
offset lines. It can be concluded that the numerical model underpredicts the experimental chamber pressure slightly.
The overall average deviation is −7.0%. This is −6.2% for SH1, −9.1% for SH2, −7.3% for SH3, and −7.4% for SH4.
The next important operating characteristics are the thrust and specific impulse. Numerically, they are obtained from
the exit velocity, mass flow rate and static pressure, using Eqs. 16 and 17. The experimental thrust was measured. The
experimental specific impulse was deducted from the average thrust and mass flow rate, using Eq. 17. A comparison
of the numerical and experimental results is shown in Fig. 9b and 9c. It can be concluded that the numerical model
overpredicts the thrust and specific impulse slightly. The overall average deviation is 9.2% for both thrust and specific
impulse. This is 5.2% for SH1, 14.9% for SH2, 12.2% for SH3, and 13.7% for SH4.
At this stage it is of course difficult to identify the reasons for the observed model deviations from the experimental
results, as there are so many experimental and numerical aspects involved. Moreover, it is quite certain that all these
aspects have some influence, negligible or not.
It can be speculated that the CFD inlet modeling for both oxidizer and fuel can be of importance. For example,
a CFD simulation has already been done in the context of the current study, in which the oxidizer enters the domain
as ∅ 30 𝜇m liquid droplets. Preliminary results show an increase in chamber pressure compared to the case with a
gaseous oxidizer inlet. However, this model is not presented yet as it needs further development.
Also note that the experimental values in Table 2, to which the CFD results are compared, can deviate to some
extent from reality. This is because of some uncertainty in the measurements themselves, and because of the space-time
averaging of these values. For example, during the experimental tests, the SH1-injector has shown to cause flame
blow-out regularly, therefore reducing the experimental time-averaged chamber pressure. It can be speculated that this
explains why the CFD results deviate less from the SH1 test data than from the SH2-4 test data.
Figure 10 shows the temperature and Mach number contours for the simulation of test SH1-01.
Figure 11 shows a typical experimental and numerical exhaust plume structure, and they seem to correspond well.
Both images are resized so that the scales would match.

12
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

(a) (b)

(c)

Fig. 9 Comparison between experimental and numerical chamber pressure (a), thrust (b),
and specific impulse (c).

13
(a)
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

(b)

Fig. 10 Contours of temperature (a) and Mach number (b) for the simulation of test SH1-01.

Fig. 11 Comparison between experimental and numerical exhaust plume structure.

14
Chamber pressure Chamber pressure
23.0 23.0
22.9 22.9
22.8 22.8
Chamber pressure [bar]

Chamber pressure [bar]


22.7 22.7
22.6 22.6
22.5 22.5
22.4 22.4
22.3 22.3
22.2 22.2
22.1 22.1
22.0 22.0
200 300 400 500 600 200 400 600 800 1000 1200
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

Oxidizer inlet temperature [K] Fuel inlet temperature [K]

(a) Oxidizer inlet (b) Fuel inlet

Fig. 12 Charts showing how the chamber pressure is affected by the inlet temperatures.

2. Parametric study
This section shows how the model responds to the oxidizer and fuel inlet temperature (Fig. 12), and mass ratio (O/F)
(Fig. 13). The monitored field value is the chamber pressure. The following observations are made.
Figure 12a. Within the range of simulated oxidizer inlet temperatures (220 − 600 K), the maximum and minimum
chamber pressure only differ 0.25 bar. A minimum is observed in the region of 400 K. In reality, the oxidizer inlet
temperature can be controlled in the sense that the storage tanks are usually at ambient temperature. Small temperature
variations in the feeding lines are assumed to be negligible. The simulated chamber pressure difference between, for
example, −20 ◦ C (253 K) and 40 ◦ C (313 K) is only 0.1 bar or 0.48%.
Figure 12b. As the fuel inlet temperature is increased, the chamber pressure increases as well. Assuming that in
reality, the temperature of the expelled gasses from the paraffin wax pyrolysis process lies somewhere within the boiling
and critical temperature region (700 − 780 K [11]), the model introduces a chamber pressure uncertainty of 0.051 bar or
0.23%. Note that the fuel gasses inlet temperature is of course not directly controllable during experiments.
Figure 13 shows how the modeled chamber pressure is affected by the oxidizer to fuel mass flow ratio. The total
mass flow rate is kept constant at 0.7 kg/s, a typical value for the SH4-injector test runs. So, for example, when the O/F
ratio is 1, both oxidizer and fuel mass flow rates are set to 0.35 kg/s. According to Eq. 12, the stoichiometric O/F ratio
is 6.57. This is consistent with Fig. 13, on which the modeled chamber pressure reaches a maximum for an O/F ratio
between 5 and 7. The experimental O/F ratios as shown in Table 2 are lower than this, ranging from 2.4 to 3.6. This
is possibly because in reality, an important part of the fuel can leave the engine as partially unburnt droplets due to
entrainment effects, contributing far less to the gaseous fuel phase in the combustion chamber. If this is true, keeping in
mind the range of the model’s optimal O/F ratio (5 − 7) and the range of the experimental O/F ratio (2.4 − 3.6), the mass
percentage of entrained fuel during the experiments can be estimated to range from 28% to 66%. These estimations can
be realistic values according to [3], taking into account the experimental port mass fluxes considered in this study of
min. 98 𝑚𝑘𝑔2 𝑠 at the upstream port section, and max. 184 𝑚𝑘𝑔2 𝑠 at the downstream port section.

V. Conclusion and future work


In this work, a first Computational Fluid Dynamics (CFD) model of the test bench Hybrid Rocket Engine at
Université Libre de Bruxelles (ULB-HRE) has been developed. Although this first model is relatively straight forward
and does not yet include more complex chemical and physical phenomena, the simulation results are promising and are
quite close to the available experimental results. The average deviation is 7% for chamber pressure, and 9.2% for thrust
and specific impulse. This offers a good foundation for further development of the CFD model, and, ultimately, for
proposing improvements of the ULB-HRE.

15
24
23
22

Chamber pressure [bar]


21
20
19
18
17
16
15
14
1 2 3 4 5 6 7 8 9 10
O/F mass flow ratio
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

Fig. 13 Chart showing how the chamber pressure is affected by the O/F mass ratio.

Furthermore, a parametric study has been carried out to monitor the model’s sensitivity to the oxidizer and fuel
inlet temperatures. Within a broad, but physically realistic range of inlet temperatures for both fuel and oxidizer, the
differences in operating chamber pressures stay below 0.1 bar or 0.5%. Therefore it can be concluded that the impact
of these parameters is relatively limited. Also, a parametric study of the inlet oxidizer to fuel mass flow ratio (O/F
ratio) has been done in order to determine which ratio leads to a maximum chamber pressure in the model. Simulations
show a maximum chamber pressure for an O/F ratio between 5 and 7, which is consistent with the stoichiometric O/F
ratio of 6.57. Looking at the lower experimental O/F ratios which are between 2.4 and 3.6, the experimental fuel mass
entrainment is estimated to range from 28% to 66%.
Elements which can be explored in the future model include the following.
To model entrained fuel droplets, a spray of liquid paraffin originating from the grain surface can be added. The
droplet temperature could be set to 515 K, which is the average of Tmelting and Tboiling of paraffin. Both phases in the
flow are coupled as the droplet size cannot be ignored. The two-phase flow is then solved in an Eulerian-Lagrangian
framework. This strategy has also been adopted by [11], although in this case, the gaseous fuel entered the domain
through a velocity inlet rather than as a mass source.
To improve the oxidizer inlet modeling, other inlet geometries can be considered, as well as a two-phase flow such
as proposed above for the fuel entrainment. Switching from an axisymmetric 2D to a full 3D computational domain
would allow to model the injector’s geometry exactly.
The chemical reaction can be altered to a more precise version, or, the entire combustion model can be altered by,
for example, switching to a mixture fraction approach instead of using the Eddy Dissipation Model (EDM).
The CFD model will be used for future parametric studies, which will definitely include modifying the dimensions
of the post combustion chamber and nozzle. As this will influence the exhaust plume structure, the use of a thermal
camera during the experiments can be interesting for comparison purposes with the CFD model’s temperature contours.

16
References
[1] Sutton, G. P., and Biblarz, O., Rocket Propulsion Elements, 9th ed., 2017, pp. 5–9.

[2] Altman, D., and Holzman, A., Fundamentals of Hybrid Rocket Combustion and Propulsion, 2007, Progress in Astronautics and
Aeronautics, Vol. 218, Chap. 1.

[3] Karabeyoglu, M. A., Cantwell, B. J., and Altman, D., “Development and Testing of Paraffin-based Hybrid Rocket Fuels,” 37th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Salt Lake City, UT, July, 2001.

[4] Leccese, G., Cavallini, E., and Pizzarelli, M., “State of Art and Current Challenges of the Paraffin-based Hybrid Rocket
Technology,” AIAA Propulsion and Energy 2019 Forum, Indianapolis, IN, August, 2019.

[5] Ferziger, J. H., and Perić, Computational Methods for Fluid Dynamics, 1999, Chap. 2.

[6] Di Martino, G. D., Carmicino, C., Mungiguerra, S., and Savino, R., “The Application of Computational Thermo-Fluid-Dynamics
to the Simulation of Hybrid Rocket Internal Ballistics with Classical or Liquefying Fuels: A Review,” Aerospace, Vol. 6,
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

Number 5, May, 2019.

[7] Bouziane, M., Bertoldi, A. E. M., Milova, P., Hendrick, P., and Lefebvre, M., “Development and Testing of a Lab-scale
Test-bench for Hybrid Rocket Motors,” 2018 SpaceOps Conference, Marseille, France, May, 2018.

[8] Bouziane, M., Bertoldi, A. E. M., Dahae, L., Milova, P., Hendrick, P., and Lefebvre, M., “Design and Experimental Evaluation
of Liquid Oxidizer Injection System for Hybrid Rocket Motors,” 7th European Conference for Aeronautics and Space Sciences
(EUCASS), Milan, Italy, May, 2017.

[9] Bouziane, M., Bertoldi, A. E. M., Dahae, L., Milova, P., Hendrick, P., and Lefebvre, M., “Experimental Investigation of
Injectors Design and Their Effects on 1kN Performance Hybrid Rocket Motor,” 69th International Astronautical Congress
(IAC), Bremen, Germany, October, 2018.

[10] Bouziane, M., Bertoldi, A. E. D., Milova, P., Hendrick, P., and Lefebrve, M. H., “Performance Comparison of Oxidizer Injectors
in a 1-kN Paraffin-Fueled Hybrid Rocket Motor,” Aerospace Science and Technology, Vol. 89, 2019, pp. 392–406.

[11] Milova, P., Blanchard, R., and Galfetti, L., “A Parametric Study of the Effect of Liquid Fuel Entrainment on the Combustion
Characteristics of a Paraffin-based Hybrid Rocket Motor,” 6th European Conference for Aeronautics and Space Sciences
(EUCASS), Kraków, Poland, July, 2015.

[12] Bredberg, J., “On the Wall Boundary Condition for Turbulence Models,” Chalmers University of Technology, Department of
Thermo and Fluid Dynamics, Göteborg, Sweden, 2000.

[13] ANSYS Fluent Theory Guide (2019 R3), ANSYS, Inc., 2019, Chaps. 4, 7.

[14] Marxman, G. A., Wooldridge, C. E., and Muzzy, R. J., “Fundamentals of Hybrid Boundary Layer Combustion,” AIAA
Heterogeneous Combustion Conference, Palm Beach, FL, December, 1963.

[15] Muzzy, R. J., “Schlieren and Shadowgraph Studies of Hybrid Boundary-Layer Combustion,” AIAA Journal, Vol. 1, Number 9,
1963.

[16] Chiaverini, M. J., and Kuo, K. K., Fundamentals of Hybrid Rocket Combustion and Propulsion, 2007, Progress in Astronautics
and Aeronautics, Vol. 218, pp. 603–606.

[17] Waxman, B. S., Cantwell, B. J., and Zilliac, G., “Effects of Injector Design and Impingement Techniques on the Atomization of
Self-Pressurizing Oxidizers,” 48th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Atlanta, GA, July, 2012.

[18] Di Martino, G. D., Malgieri, P., Carmicino, C., and Savino, R., “A Simplified Computational Fluid-Dynamic Approach to the
Oxidizer Injector Design In Hybrid Rockets,” Acta Astronautica, Vol. 129, 2016.

[19] Echekki, T., and Mastorakos, E. (eds.), Turbulent Combustion Modeling, 2011, Fluid Mechanics and its Applications, Vol. 95,
pp. 28–34.

[20] Bianchi, D., Nasuti, F., and Carmicino, C., “Hybrid Rockets with Axial Injector: Port Diameter Effect on Fuel Regression Rate,”
Journal of Propulsion and Power, Vol. 32, Number 4, July, 2016.

[21] Di Martino, G. D., Mungiguerra, S., Carmicino, C., and Savino, R., “Two-Hundred-Newton Laboratory-Scale Hybrid Rocket
Testing for Paraffin Fuel-Performance Characterization,” Journal of Propulsion and Power, Vol. 35, Number 1, 2019.

17
[22] Leccese, G., Bianchi, D., and Nasuti, F., “Modeling and Simulation of Paraffin-Based Hybrid Rocket Internal Ballistics,” AIAA
Propulsion and Energy 2018 Forum, Cincinnati, OH, July, 2018.

[23] Paccagnella, E., Gelain, R., Barato, F., Pavarin, D., van den Berg, P., and Barreiro, F., “CFD Simulations of Self-Pressurized
Nitrous Oxide Hybrid Rocket Motors,” AIAA Propulsion and Energy 2018 Forum, Cincinnati, OH, July, 2018.

[24] Bianchi, D., Nasuti, F., and Delfini, D., “Modeling of Gas-Surface Interface for Paraffin-Based Hybrid Rocket Fuels in
Computational Fluid Dynamics Simulations,” Progress in Propulsion Physics, Vol. 11, 2019.

[25] Jones, W. P., and Launder, B. E., “The Prediction of Laminarization with a Two-Equation Model of Turbulence,” Int. Journal
for Heat and Mass Transfer, Vol. 15, 1972.

[26] Wilcox, D. C., “Reassessment of the Scale-Determinig Equation for Advanced Turbulence Models,” AIAA Journal, Vol. 26,
Number 11, 1988.

[27] Menter, F. R., “Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications,” AIAA Journal, Vol. 32,
Number 8, 1994.
Downloaded by CARLETON UNIVERSITY on August 23, 2020 | http://arc.aiaa.org | DOI: 10.2514/6.2020-3763

[28] Spalart, P. R., and R, A. S., “A One-Equation Turbulence Model for Aerodynamic Flows,” AIAA 30th Aerospace Sciences
Meeting and Exhibit, Reno, NV, January, 1992.

[29] Magnussen, B. F., and Hjertager, B. H., “On Mathematical Modeling of Turbulent Combustion With Special Emphasis on Soot
Formation and Combustion,” 1977.

[30] May, S., Karl, S., and Bo𝑧ˇi𝑐,


´ “Development of an Eddy Dissipation Model for the use in Numerical Hybrid Rocket Engine
Combustion Simulation,” 7th European Conference for Aeronautics and Space Sciences (EUCASS), Milan, Italy, 2017.

[31] Sivathanu, Y. R., and Faeth, G. M., “Generalized State Relationships for Scalar Properties in Nonpremixed Hydrocarbon/Air
Flames,” Combustion and Flame, Vol. 82, 1990.

[32] Di Martino, G. D., Mungiguerra, S., Carmicino, C., and Savino, R., “Computational Fluid-dynamic Simulations of the Internal
Ballistics of Hybrid Rocket Burning Paraffin-based Fuel,” AIAA Propulsion and Energy 2018 Forum, Cincinnati, OH, July,
2018.

[33] Karabeyoglu, M. A., Cantwell, B. J., and Stevens, J., “Evaluation of the Homologous Series of Normal Alkanes as Hybrid
Rocket,” 41st AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, Tucson, AZ, July, 2005.

[34] Adachi, M., and Shimada, T., “Liquid Films Instability Analysis of Liquefying Hybrid Rocket Fuels Under Supercritical
Conditions,” AIAA Journal, Vol. 53, Number 6, 2015.

[35] Bouziane, M., “Influence of the Oxidizer Injector Design on the Performance of a 1-kN Paraffin-Fueled Hybrid Rocket Motor,”
Ph.D. thesis, Université Libre de Bruxelles, Brussels, Belgium, 2019.

18

You might also like