You are on page 1of 43

DEGREE PROJECT IN MECHANICS,

SECOND CYCLE, 30 CREDITS


STOCKHOLM, SWEDEN 2023

Simulation of
Combustion in a
Hybrid Rocket Engine
MSc Thesis Report

Oscar Andersson

KTH ROYAL INSTITUTE OF TECHNOLOGY


SCHOOL OF ENGINEERING SCIENCES
TRITA-SCI-GRU 2023:035

Author
Oscar Andersson oscaa@kth.se
Aerospace Engineering
KTH Royal Institute of Technology

Location
FOI Grindsjön
Stockholm, Sweden

Examiner
Anders Dahlkild
KTH Royal Institute of Technology

Supervisor

Christian Ibron
FOI Grindsjön

ii
Abstract

A numerical investigation on the combustion mechanics of a hybrid rocket engine


is performed through unsteady Reynolds-averaged Navier-Stokes simulation. The
hybrid rocket engine model is based on an experimental laboratory scale engine design
operating on GOX and HDPE as a propellant. A simple convection heat flux model is
used to determine the heat transfer to the fuel wall. The project is done with the goal of
finding the fuel regression rate in mind, as it is an essential parameter for determining
engine performance. The results show early results of the fluid- and thermodynamics
occurring in the combustion chamber. Propellant mixing is shown to not be optimal
as a significant part of the exit flow consists of high concentrations of oxidizer that
has not reacted with the fuel. The flame temperature is shown to be relatively high
inside the combustion chamber. It is concluded from the simulations that the model
needs further improvement in order to accurately compute the flow as well as the heat
transfer to the fuel. To determine the regression rate, radiation should be implemented
into the heat transfer model.

Keywords

Thesis, Hybrid Rocket Engine, CFD, computational fluid dynamics, combustion

iii
Sammanfattning

En numerisk undersökning av förbränningsmekaniken i en hyrbidraketmotor är


gjord genom unsteady Reynolds-averaged Navier-Stokes-simulering. Modellen av
hybridraketmotorn är baserad på en experimentell laboratorieskaleraket med GOX
och HDPE som framdrivningsmedel. En enkel ”heat flux”-modell med konvektion
används för att beräkna värmeöverföringen till bränsleväggen. Projektet görs med
målet i åtanke att kunna bestämma bränsleregressionen, då detta är en grundläggande
parameter för att ta reda på motorns prestanda. Resultaten visar tidiga resultat av
de strömnings- och termodynamiska fenomen som sker i förbränningskammaren.
Blandningen av bränslet visar sig inte vara optimalt då en väsentlig andel av
utflödet består av höga koncentrationer av oxidatorn som ej reagerat med bränslet.
Flamtemperaturen visas vara relativt hög i förbränningskammaren. Den slutsats
som dras från simuleringarna är att modellen behöver ytterligare förbättringar för
att, med noggrannhet, kunna beräkna flödet och värmeöverföringen till bränslet.
För att beräkna bränsleregressionen bör hänsyn till strålningsvärme implementeras
i värmeöverföringsmodellen.

Nyckelord

Examensarbetet, hybridraketmotor, CFD, numeriska strömningsmekaniska


beräkningar, förbränning

iv
Acknowledgements

I am deeply indebted to my supervisor Christian Ibron for his aid with providing his
expertise as well as patience and encouragement over this period, without which this
thesis would not have been possible. I am extremely grateful to my examiner Anders
Dahlkild, for his supporting me and feedback throughout this project.

I would like to thank FOI, the Swedish Defence Research Agency, and KTH Royal
Institute of Technology for giving me this opportunity to work on such an interesting
project. It has been a truly rewarding endeavour that has given me insight and
experience I could not have gotten anywhere else.

Special thanks to Oskar, Niklas, Niklas, Stefan and Kevin for sharing knowledge in
their respective areas.

Lastly, I want to thank my family for their continuous support and belief in me over
the time of this project.

v
Nomenclature

Symbols
ṙ regression rate
ϵ emissivity
µ dynamic viscosity
ρ density
ε dissipation rate of the turbulent kinetic energy
B blowing parameter
G mass flux
hv heat of gasification
k turbulent kinetic energy
p pressure
Qtot total heat transfer
Re Reynolds number
T temperature
x axial distance from grain head end

Subscripts
b burned gas
e core flow
f fuel
w wall

Abbreviations
BC boundary condition
CFD computational fluid dynamics
DNS direct numerical simulation
GOX gaseous oxygen

vi
NOMENCLATURE

HDPE high-density polyethylene


HRE hybrid rocket engine
HTPB hydroxyl-terminated polybutadiene
LES large eddy simulation
O/F oxidizer to fuel ratio
RANS Reynolds-averaged Navier-Stokes
URANS unsteady Reynolds-averaged Navier-Stokes

vii
Contents

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Fundamental Principles of Hybrid Rocket Combustion . . . . . . . . . . 3
1.3 Fuel Regression Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Numerical Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.5 Problem Statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2 Theory 9
2.1 Numerical Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.1 Turbulence Modelling . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.2 Wall Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 Injector Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Combustion Instabilities . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4 Kinetics/Chemical Reaction . . . . . . . . . . . . . . . . . . . . . . . . . 13

3 Method 14
3.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3.2 Mesh Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.3 Turbulence Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.4 Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.5 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.5.1 Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.5.2 Flow Rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.5.3 Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.5.4 Starting the Simulation . . . . . . . . . . . . . . . . . . . . . . . 22
3.6 The Kinetics of Combustion . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.7 Heat Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

viii
CONTENTS

4 Results 24

5 Conclusion 29
5.1 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.1.1 Regression Rate . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5.1.2 CFD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
5.1.3 Flow and Heat Transfer . . . . . . . . . . . . . . . . . . . . . . . 31

References 33

ix
Chapter 1

Introduction

1.1 Background
The development of hybrid rockets have been documented since its birth in the 1930s
[2]. Although not being the most common type of engine used for the majority of
rockets it possesses unique properties that separate it from solid and liquid engines,
proving to be beneficial for certain mission profiles. With the commercialization
and general growth of the space industry and with a market in demand there has
been a resurgence in interest of the concept. Some areas of interest for military
and commercial use are sounding rockets, tactical rockets, spacecraft propulsion and
launch boosters [3].

Combining technologies from both solid and liquid rockets gives the hybrid engine
unique features to its advantage, some of which being [3, 4, 17]:

1. Reduction of system complexity, since only one component of the propellant is


stored as a fluid and thus only half of the plumbing is required, as opposed to
liquid rockets.

2. Throttleability as well as engine termination and restart possibilities through


inlet flow rate control.

3. Flexibility due to several areas of system application.

4. Safer to handle and operate due to oxidizer and fuel being stored separately. The
propellant is generally nonexplosive. The increased safety in manufacturing,
handling and reliability would consequently lead to lower operational and

1
CHAPTER 1. INTRODUCTION

maintenance costs.

5. Faults or grain defects such as crack formation are not of critical risk for
operation, making the system more robust.

6. High specific impulse near or on par with that of liquid rockets.

Some drawbacks of the hybrid rocket propulsion systems [3]:

1. Low regression rate causing low performance.

2. O/F shift resulting in decreased performance if not dealt with appropriately.

3. Combustion inefficiency due to poor mixing, as oxidizer and propellant react in


a diffusion flame.

4. Transients and combustion instabilities.

Low regression rate is a major drawback particular to the hybrid rocket. A


characteristic that is closely linked to the nature of the internal ballistics of any hybrid
design. As a consequence the performance is relatively low compared to that of solid
and liquid rocket designs, when it comes to scaling, which has made it an unfeasible
choice for larger vehicles. The mechanisms driving the regression rate of hybrid
rockets differs from the other conventional rocket propulsion systems in that they
are heavily dependent on the port gas flow, eddies, turbulence and recirculation [17].
Understanding the regression rate is essential for the improvement of hybrid rockets
engines.

The regression rate, similarly to solid rockets, is defined as the rate at which the surface
of the grain inside the combustion chamber moves perpendicular to the flow moving
in the axial direction. Solid rockets do however have the propellant premixed as a
grain, and the surface undergoes combustion as it is heated by the flame [3], whereas
hybrid rockets have oxidizer and fuel stored separately. Hence solid rocket engines
have a consistent propellant mixture. For most solid rocket propellant combinations,
one can approximate the regression rate as a function of pressure:

r = apn1 (1.1)

where r is the regression rate, a is an empirically obtained temperature constant, pn1


is the operating pressure, and n the pressure exponent. For more detailed analysis

2
CHAPTER 1. INTRODUCTION

experimental data is required [17]. As hybrid rocket propellants are not premixed
the solid fuel must undergo phase change and combustion only occurs in a diffusion
flame zone. Efficient mixing of the propellant components has to occur in order to
achieve satisfactory combustion efficiency and performance. This leads to necessary
flow analysis needed to be done on hybrids for understanding and improving mixing
of the propellant.

1.2 Fundamental Principles of Hybrid Rocket


Combustion
Hybrid rocket engines work by having the propellant components stored in to different
states. That is, the fuel as a solid mold inside the combustion chamber and a liquid
oxidizer in a separate container. Having the oxidizer as a fluid and fuel as a solid is
commonly referred to as a classical hybrid. A visual representation is provided by
Fig. 1.2.1. One could have a configuration the other way around where the oxidizer is
stored as a solid and have fuel injected as a fluid into the chamber, although this is not
as common. Combustion occurs as oxidizer is injected into the combustion chamber
where mixes with the pyrolyzed fuel grain from the fuel wall. The pyrolysis is caused
by the heat transfer from radiation and convection from the flame. As fuel leaves the
wall and enters the flow, it has a blocking effect on the heat transfer to the wall. This
in turn, decreases the blowing effect, thus increasing the heat transfer, making the fuel
mass flux self-regulating [4].

Oxidizer Fuel

Figure 1.2.1: Hybrid rocket engine concept.

The unique interaction between the oxidizer and propellant in contrast to other
propulsion systems is the existence of a boundary flame between the ”free stream”
oxidizer flow and the fuel wall. The flame resides within the boundary layer and
originates from the fact that the fuel surface undergoes a phase change from solid (to
liquid) to gas. Fuel is expelled from the fuel wall into the flow. This leads to having

3
CHAPTER 1. INTRODUCTION

a fuel rich region between the flame and the fuel wall and an oxidizer rich region
above the flame in the free stream. The pyrolysis of the fuel is primarily driven by
the heat transfer to the fuel wall by convection and radiation. The combustion process
is illustrated in Fig. 1.2.2 as based on the description in [10].

Oxidizer rich region

Bulk
flow Flame zone

Fuel rich region

Fuel wall

Figure 1.2.2: Sketch of the diffusion combustion process.

The system inherently has an oxidizer-to-fuel ratio (O/F) shift during combustion due
to the diffusion flame running down the length of the chamber, changing the contents
of the flow at different positions. O/F shift is also caused by the change in size of
exposed surface area of the fuel over time. To maximize efficiency of the rocket, the
latter must be accounted for through control of the oxidizer flow rate.

The choice of propellant of the combustion and the fuel affects how big of a role
each phenomenon of the heat transfer plays. It is therefore important to know the
characteristics of the oxidizer and fuel in order to accurately model the mechanics
taking place. Several viable propellants for HREs that have been experimented with
and documented. The choice of what oxidizer and fuel to use will greatly impact several
parameters of the engine, e.g. performance and exhaust properties. Some propellant
combinations have the opportunity of providing high performance with the addition
of a more environmentally friendly exhaust. A popular hybrid fuel, HTPB (hydroxyl-
terminated polybutadiene), in combination with LOX (liquid oxygen) is one example
that has a nontoxic and relatively low smoke exhaust [17]. HTPB is suitable for large
scale engines as a replacement for boosters. Other fuels commonly used for laboratory
scale experiments are paraffin and polyethylene.

Convection has been shown to be the primary driver of heat transfer to the fuel wall for
nonmetallized fuels [17]. For metallized fuels or carbon-heavy fuels, radiation effects
could be significant whereas for nonmetallized fuels the effect might be negligible
[10] in early analysis. A study on a particular hybrid rocket design by Lazzarin et

4
CHAPTER 1. INTRODUCTION

al.[8] concluded that the regression rate was underestimated by 50% before accounting
for radiation. With radiation this was reduced to 30%. Nevertheless the effects of
radiation should be considered.

1.3 Fuel Regression Rate


Enchancing the regression rate is one of the major challenges for hybrid rocket
development. Various methods are being studies that could help in this area.
Some concerning experimentation with fuel and oxidizer compositions and their
combination with each other, and different geometric designs of the solid component
such as multiport configurations and spiraling ports.

The 1960s resparked an interest in hybrid rockets and investigations were made to
better understand the fundamentals of the combustion. Research and development
was done at several places, some noteworthy investigating organizations being
ONERA, Volvo Flygmotor as well as the United States.

To combat the low regression rate, fuel additives have been experimented with such
as NH4 CO4 (ammonium perchlorate), NH4 NO3 (ammonium nitrate). The use of
metals e.g. Mg, Al and B to enhance the regression rate by reducing the effective
heat of gasification have been investigated [4]. Strand et al. [16] found that adding
Al and coal to HTPB (Hydroxyl-terminated polybutadiene) increased the regression
compared to fuel consisting of pure HTPB, but does not necessarily indicate an increase
in performance since it may come at the expense of decreased combustion efficiency
[4].

Other ways of increasing regression rate have been through manipulation of the flow.
One example of this is having a varying fuel surface by adding troughs periodically
along the port to act as turbulence generators [3]. Another approach to generate
turbulence in order to increase the fuel regression is by implementing a circular
injector. This is installed in the aft chamber and has been proven to significantly
increase the regression rate for commonly used polymeric substances [3].

To this day works by Marxman and Gilbert [13] and Altman and Humble [1], and
others, which have been expanded upon [7, 10], are still referenced and used as a
foundation for modelling contemporary designs. In the early works, experimental data
was used to analytically derive semi-empirical formulas for describing the regression

5
CHAPTER 1. INTRODUCTION

rate. Marxman initially provided an expression for the regression rate by applying
energy flux balance at the surface in [11], by analyzing combustion with oxidizer flowing
over a slab of fuel:

ρf ṙhv = Q̇w (1.2)

where ρf is the solid fuel density, ṙ the regression rate, hv the heat of gasification and
Q̇w the total heat flux to the wall. This is the basis for the derivation of what would
become ”Marxman’s diffusion-limited model”. The model assumes that fuel regression
is controlled by heat transfer from the flame. In addition to that, Reynolds analogy is
assumed to hold. The momentum and thermal diffusivity are assumed to be similar,
and so is the thermal and mass diffusivity. Thus, the Lewis and Prandtl numbers
are taken to be unity [12]. Marquardt and Majdalani have provided a derivation
of Marxman’s diffusion-limited model [10], arriving at the expression for regression
rate

Gx −0.2 0.23
ρf ṙ = 0.036G( ) B (1.3)
µe

where ρf is the solid fuel density, G is the mass flux, µe is core flow dynamic
viscosity, and B the blowing parameter that corrects for difference in heat flux between
combusting and noncombusting cases. Collecting the constant terms the expression
above can alternatively be written as

ṙ = a1 G0.8 x−0.2 (1.4)

where a1 as a constant is related to the propellant. One can note from eq. (1.4) that
the local mass flux, G, is the primary parameter governing the regression rate. In the
space-time averaged case

ṙ = a2 Gn (1.5)

which can be used when working from experimental measurements.

For cases where heat transfer from radiation has to be taken into account, Marxman

6
CHAPTER 1. INTRODUCTION

and Gilbert added a grey-body radiation term to Eq. 1.3 such that

Gx −0.2 0.23 σϵ(ϵg Tb4 − Tw4 )


ρf ṙ = 0.036G( ) B + (1.6)
µe hv

The accuracy of the expression above has been considered to not be an accurate way
of describing the radiation, as it does not take into account the blowing and blocking
effects [10].

Other equations for describing regression rate have been presented over the years,
each with their own benefits, several of which can be found in [4]. These models vary
in complexity and in which physical phenomena are accounted for. A problem for
classical models such as Marxman’s diffusion-limited model is that they do not account
for details in geometry and the details of chamber flow. Comparisons of simulations
with experimental data has shown significant variations between the results [4].

1.4 Numerical Modelling

Regression rate is dependent on several variables, primarily the choice of propellant


and engine design geometry, often requiring multiple experimental tests to gather
data on engine parameters resulting in high research and development costs. Since
experimental studies also rely on simplified modelling the are limited in that they
cannot provide the quantitative data required for late-stage engine design. As
computational power has increased over the years, CFD has become a resourceful tool
for simulating the combustion physics and calculating regression rate. Whilst earlier
relying on experimental data and analytical models, one can now through numerical
methods simulate the effects of advective and radiative heating on regression. This has
proven to be cost-effective and design parameter analysis can be done easily. Recent
investigations have shown that numerical models to be a viable method and when
comparing numerical with experimental results, prove to provide additional, valuable
information of the combustion mechanics in HREs [5].

The level of accuracy of the flow required for the solution and the computational
resources available determine the choice of numerical model for resolving the flow. The
choice of model governs what phenomena of the flow are captured. For many cases,
e.g. industry and early research, the general flow characteristics suffice without the

7
CHAPTER 1. INTRODUCTION

need for computing the flow down to the smallest eddies. Hence Reynolds-averaged
Navier-Stokes modelling (RANS) is feasible for these tasks.

Unsteady RANS (URANS) solves the ensemble average taking into account time-
varying velocity in the momentum equation. In other words, URANS has the ability
to capture velocity changes and large scale properties of the flow but not any present
turbulence. Large eddy simulation (LES) on the other hand simulates large eddies and
the general flow at a higher resolution and is able to capture the turbulent mixing of
the flow. URANS provides a way for analysing the general behaviour of the flow over
time at a lower computational cost compared to LES, at the expense of not capturing
the turbulent eddies and accuracy. Furthermore, unlike LES, the URANS model has
wall functions which are necessary in order to determine the heat transfer to the fuel
wall.

1.5 Problem Statement


This thesis presents a study on the combustor flow mechanics of a of a hybrid rocket
engine design through CFD simulation in order to investigate what key physical
processes determine heat transfer through high-density polyethylene fuel (HDPE) and,
in turn, regression rates of said fuel. The oxidizer for the case is gaseous oxygen (GOX).
The simulation is done with an unsteady RANS (URANS) model in order to capture the
heat transfer to the wall. The desired oxidizer mass flux is 300 g/cm2 s.

8
Chapter 2

Theory

2.1 Numerical Modelling

The governing equations for Reynolds-averaged Navier-Stokes, compressible flow, are


given by [19]

Mass conservation
∂ ρ̄
+ div(ρ̄Ũ) = 0 (2.1)
∂t

Momentum equations

[ ′2 )
]
∂(ρ̄u′ v ′ ) ∂(ρ̄u′ w′ )
∂(ρ̄Ũ )
∂t
+ div(ρ̄Ũ Ũ) = − ∂∂xP̄ + div(µ grad Ũ ) + − ∂(ρ̄u
∂x
− ∂y
− ∂z
+ SM x (2.2)

[ ′ v′ )
]
∂(ρ̄v ′2 ) ∂(ρ̄v ′ w′ )
∂(ρ̄Ṽ )
∂t
+ div(ρ̄Ṽ Ũ) = − ∂∂yP̄ + div(µ grad Ṽ ) + − ∂(ρ̄u
∂x
− ∂y
− ∂z
+ SM y (2.3)

[ ′ w′ )
]
∂(ρ̄v ′ w′ ) ∂(ρ̄w′2 )
∂(ρ̄W̃ )
∂t
+ div(ρ̄W̃ Ũ) = − ∂∂zP̄ + div(µ grad W̃ ) + − ∂(ρ̄u
∂x
− ∂y
− ∂z
+ SM z (2.4)

9
CHAPTER 2. THEORY

Scalar transport equation


[ ]
∂(ρ̄Φ̃) ∂(ρ̄u′ ϕ′ ) ∂(ρ̄v ′ ϕ′ ) ∂(ρ̄w′ ϕ′ )
+ div(ρ̄Φ̃Ũ) = div(ΓΦ grad Φ̃) + − − − + SΦ (2.5)
∂t ∂x ∂y ∂z

2.1.1 Turbulence Modelling

Simulating the turbulent flow field correctly of the flow occurring is key to make sure
that the solution is accurately representing the flow. There are several models available
commonly found in both research and industry, each with their own strengths and
weaknesses. Selecting the appropriate model relies heavily on subject knowledge, prior
experience and what is to be extracted from the results.

The k-ωSST, two-equation eddy-viscosity model [14], utilizes equations of both the k-ω
and k-ε models depending on where in the domain the flow is calculated. The near
wall region region is computed with the k-ω model whereas the outer wake and free
stream regions are calculated using the k-epsilon model [14]. This way, boundary layer
and flow separation is better handled by the k-ω model without sacrificing solution
accuracy of the solution in the free stream flow where k − ε is active instead.

The model equations for the turbulence specific dissipation rate equation, turbulence
kinetic energy and turbulence viscosity are given by [6]

D ργG 2
(ρω) = ∇ · (ρDω ∇ω) + − ργω(∇ · u) − ρβω 2 − ρ(F1 − 1)CDkω + Sω (2.6)
Dt ν 3

D 2
(ρk) = ∇ · (ρDk ∇k) + ρG − ρk(∇u) − ρβ ∗ ωk + Sk (2.7)
Dt 3

respectively. The turbulence viscosity is given by [6]

k
ν t = a1 (2.8)
max(a1 ω, b1 F2 S)

a1 and b1 and β ∗ are default coefficients, F2 is what is known as a second blending


function and S is the invariant measure of the strain rate. The details of the model can
be found in [6, 15].

10
CHAPTER 2. THEORY

2.1.2 Wall Treatment

To be able to determine the heat flux to the wall the boundary layer flow needs
to handled properly. There are various methods for capturing the flow inside the
boundary layer. In order to not have to solve for the entire boundary layer numerically,
one can instead use the semi-empirically derived theory, Law of the wall. The law uses
a formula for calculating the average velocity field at the boundary as proportional to
the logarithm of the distance to the wall. Using this method of approach, the flow field
can accurately be captured with a lower mesh resolution at the wall boundary than
what would have been necessary with DNS or LES, where the viscous sublayer would
otherwise need to be resolved.

Universal flow characteristics for the flow velocity in the boundary layer can be divided
into three regions with regard to the distance to the wall. From nearest to farthest from
the wall, these are the viscous sublayer, the buffer layer and the overlap layer. Closest
to the wall the velocity profile is linearly proportional to the wall distance which is given
by

u+ = y + (2.9)

for 0 < y + < 5, where u+ and y + are the nondimensionalized average flow velocity
tangential to the wall and the nondimensionalized normal distance from the wall, given
by [20]

u
u+ = (2.10)
u∗

yu∗
y+ = (2.11)
ν

respectively, where y is the distance between the wall boundary and the adjacent cell’s
center, u∗ is the friction velocity and ν is the kinematic viscosity. The friction velocity,
u∗ , is defined as [20]

( )1/2
∗ τw
u = (2.12)
ρ

11
CHAPTER 2. THEORY

The overlap layer the velocity following a logarithmic profile in relation to y. This is
given by [20]

1
u+ = ln y + + B (2.13)
κ

where κ ≈ 0.41 and B ≈ 5.0 for smooth walls.

The layer contains the region where the Law of the wall can be applied, the range of
which is 30 < y + < 500 [18]. The size of the cells adjacent to the wall needs the correct
dimensioning such that the cell center lies within the log-law region in the boundary
layer, to properly utilize the wall function.

2.2 Injector Effects

The way oxidizer is injected into the combustion chamber plays a significant role in the
flow and combustion characteristics and will as a consequence affect the regression
rate and combustion efficiency. Some configurations that have been investigated
are head-end, as well as aft-end axial, radial and swirl injectors. Axial injectors
on laboratory-scale motors have been found to have higher regression rates and
combustion efficiencies than radial injectors but at the expense of less even regression
rate along the surface [4].

Attempts on increasing the regression rate through swirl injectors have been made and
experiments have shown that swirl injectors can increase the regression rate to over
that of axial injectors with only a slight decrease in combustion stability [4].

2.3 Combustion Instabilities

Combustion engines experience pressure oscillations around the natural frequencies


of the combustion chamber, from flame holding, and/or the oxidizer feed system’s
induced oscillations. Analysis of combustion instabilities will not be performed in this
thesis, nevertheless it is an important topic of study for further investigation.

12
CHAPTER 2. THEORY

2.4 Kinetics/Chemical Reaction


The kinetics are modelled based on the chemical reactions between GOX and HDPE.
Being a polymer, polyethylene, in its solid state is made up by series of bonds of carbon
atoms and the pyrolysis the fuel results in constituents of ethylene monomers as well
as polymer chains of varying length.

The rate at which reactions occur depends to a large degree on the temperature. This
leads to some reaction paths appearing more frequently than others. Nevertheless, the
total amount of reactions that can occur are many. In the case of CFD, this means
that a large amount of computations need to be done if one is to use a kinetics model
including all possible reaction paths. To reduce the computational costs a skeletal
kinetic mechanism can be used instead, only performs calculations for the major
reactions.

It has been observed that the nature of the chemical reaction in heterogeneous
combustion appears to depend on pressure. For lower pressures gas-phase reactions
are slower, or ”kinetically limited”. A greater mixing of oxidizer and fuel has time to
take place. Over a certain pressure threshold reaction rates are higher and the reactions
are ”diffusion limited” instead [4].

The rate of reaction taking place between the oxidizer and the fuel is governed by the
Arrhenius equation,

k = Ae(−Ea /RT ) (2.14)

where k is the reaction rate between the reactants, A is a constant, Ea is the activation
energy, R is the universal gas constant and T is the temperature.

Depending on the accuracy of solution required, kinetics models can be modified


to only account for reactions deemed to have a significant impact on the overall
combustion. Since the reaction rate to a large degree is dependent on the temperature,
combustion taking place in a high temperature region will more closely resemble the
stoichiometric reaction, offering a reduction in the number of reactions needed in the
kinetics model without sacrificing significant accuracy.

13
Chapter 3

Method

The case model is a hybrid rocket engine with a cylindrical fuel geometry and no swirl,
axial injection. A second model similar to the first one but with a circulation zone added
near the injector, to be compared to the first model. The GOX oxidizer is assumed to
have a port mass flux of 30 g/cm2 s. Fuel regression is not modelled in these cases, so
the port diameter remains static throughout the simulations. In these simulations it
is assumed that the only product from pyrolysis is the monomer ethylene. The rate of
the fuel mass flux to come from the fuel wall is proportional to the stoichiometric O/F
and is assumed to be in gaseous state.

The simulation is run with RAS, the software’s URANS solver in OpenFOAM, in the
kOmegaSST model. The application used is rhoReactingFoam, a density-based solver
for compressible flows.

3.1 Geometry

The geometry to be simulated over is based on an experimental hybrid rocket engine by


FOI. A CAD image of the section view of the engine is shown in Fig. 3.1.1. A simplified
model is made of the combustion chamber including the nozzle as shown in Fig 3.1.2,
along with relevant measurements. The combustion chamber consists of a single port
of the solid fuel. The injector is simplified to be an axial cylindrical inlet and a rough
convergent-divergent nozzle shape is present in order to simulate choked flow.

14
CHAPTER 3. METHOD

Figure 3.1.1: Experimental hybrid rocket engine by FOI.

Figure 3.1.2: Drawing of combustion with nozzle with measurements in [mm].

The geometry simulated, shown in Fig. 3.1.4, is the volume generated by a 30° rotation
about the axis of symmetry of the area highlighter in red as seen in Fig. 3.1.3. The
general shapes of the profile are simplified. The nozzle is represented by straight lines
with no smooth curvature in the throat. The injection port is modeled as a cylindrical
port coaxial with the axis of symmetry.

Figure 3.1.3: The highlighed area of the first case engine’s section view that the
geometry is created from.

15
CHAPTER 3. METHOD

(a) Geometry section of engine to be (b) 2D side view of engine geometry to be


simulated. simulated.

Figure 3.1.4: Section geometry of engine to be simulated.

For the second case, part of the fuel wall has been removed at the fore section (closest to
the inlet), making up the recirculation zone. This is shown in Fig. 3.1.5. The resulting
geometry is presented in Fig. 3.1.6.

Figure 3.1.5: The highlighed area of the second case engine’s section view that the
geometry is created from.

(a) Geometry section of engine to be (b) 2D side view of engine geometry to be


simulated simulated

Figure 3.1.6: Section geometry of engine with added recirculation zone.

3.2 Mesh Generation


Mesh generation is done through the ”blockMesh” utility giving a geometry as
represented by hexahedral cells. The cell size is set such that the Courant number,
Co < 0.7, with respect to the cells within the combustion chamber (all cells before the
nozzle throat). to make sure that the solution is converges correctly. The overall cell
size i kept fairly uniform with no mesh grading before the nozzle. In the vicinity of the
nozzle mesh grading is done to accommodate for the geometry changes, to maintain

16
CHAPTER 3. METHOD

othrogonal quality and minimize skewness as blockMesh keeps the number of cells
uniform over the axes for each given volume block. Edges in the azimuthal direction
are curved according to the rotational shape of the model.

To validate the simulation results the y + -values at the fuel wall are computed with
OpenFOAM’s ”yPlus” post-processing utility. Cells adjacent to the fuel wall boundary
are sized such that y+ is in the inner region of the log-law layer. That is, about
30 < y + < 200. Mesh refinements are then done where needed until a satisfactory
y + is achieved by generating a new (refined) mesh and mapping the current flow field
onto the new mesh and updating y + . The final meshes for the case without and with
recirculation are given in Fig. 3.2.1 and 3.2.2, respectively.

(a) View from inlet. (b) Side view.

(c) Fore section. (d) Aft section.

Figure 3.2.1: Mesh without recirculation zone.

17
CHAPTER 3. METHOD

(a) View from inlet. (b) Side view.

(c) Fore section. (d) Aft section.

Figure 3.2.2: Mesh with recirculation zone.

The y + values at the fuel wall boundary for the two cases are given in Fig. 3.2.3 and
3.2.4.

Figure 3.2.3: y + at the axial and aft face of the fuel wall.

18
CHAPTER 3. METHOD

Figure 3.2.4: y + at the recirculation zone, the axial and aft face of the fuel wall.

3.3 Turbulence Model

The turbulence model is set to k-ωSST. In order to start the simulation initial values
for k and ω are required. These are estimated with equations [6, 9]

3
k = (I|uref |)2 (3.1)
2

and

k 0.5
ω= (3.2)
Cµ0.25 L

respectively. The turbulence intensity, I, is set to 10%. Setting the reference velocity,
ur ef , to the port flow velocity it is approximately 60 m/s. Inserting these values into
eq. (3.1) gives a turbulence kinetic energy k = 13.5 J/kg. Following with eq. (3.2),
Cµ = 0.09 is a constant and setting the reference length to the port radius, L = 0.017
m, gives ω = 394.6.

3.4 Schemes

The numerical schemes applied are chosen primarily based on maintaining numerical
stability over accuracy. The numerical schemes selected are presented in 3.4.1.

19
CHAPTER 3. METHOD

Table 3.4.1: A list of the numerical schemes selected in OpenFOAM.

Term Scheme

ddtSchemes, ∂/∂t, ∂ 2 /∂t2


default Euler

gradSchemes, ∇
default Euler

divSchemes, ∇·
default none
div(phi, U) Gauss limitedLinearV 1
div(phi, Yi) Gauss limitedLinear01 1
div(phi, h) Gauss limitedLinear 1
div(phi, K) Gauss limitedLinear 1
div(phid, p) Gauss limitedLinear 1
div(phi, epsilon) Gauss upwind
div(phi, omega) Gauss upwind
div(phi, Yi_h) Gauss limitedLinear01 1
div(phi, k) Gauss upwind
div(((rho*nuEff)*dev2(T(grad(U))))) Gauss linear

Laplacian schemes, ∇2
default Gauss linear corrected

Interpolation schemes
default linear

snGradSchemes
default corrected

wallDist
method meshWave

20
CHAPTER 3. METHOD

3.5 Boundary Conditions

To get a good solution for the flow proper boundary conditions (BCs) need to be
applied. Here one has to keep in mind that since this is an investigation of high velocity
flow, the fluid needs to be treated as compressible. Key boundary conditions that give
a well behaving flow are presented here.

3.5.1 Pressure

With the presence of the convergent-divergent nozzle, the flow will be choked and
become supersonic passing through. For the pressure condition, the nozzle exit
boundary is set to ”zeroGradient”, meaning that the gradient values at the boundary
are set to zero. Hence the fluid will exit the boundary without interference. This
is necessary since no ambient environment outside of the engine, and therefore no
ambient pressure is included. Wall and fuel wall BCs are set to ”zeroGradient”.

The inlet pressure BC is given the ”fixedValue” condition and its initial value set to 10%
of the target pressure of 20 bar. After the simulation has run for 4 ms, it is gradually
increased in steps of 25% every 5 ms up to 100%.

3.5.2 Flow Rates

Since the oxidizer mass flux is predefined to be 30 g/cm2 s, the inlet velocity BC is set
to ”flowRateInletVelocity” with the ”massflowRate” set to the mass flow rate 0.0227
kg/s, corresponding to the desired mass flux and the 30° circular sector. This creates
uniform boundary velocity that is regulated by the BCs applied to the same boundary
to maintain the set flow rate.

The fuel wall BC is set to ”fixedValue” with a constant mass flow rate at 0.0078.

The mass flow rate of the oxidizer through the inlet, and fuel from the solid fuel surface
is initially estimated to be stoichiometric. Taking into account the molecular mass the
ideal O/F is calculated.

The stoichiometric reaction is

3O2 + C2 H4 = 2H2 O + 2CO2 (3.3)

21
CHAPTER 3. METHOD

With oxygen and ethylene having a molar mass of approximately 31.999 g/mol and
28.054 g/mol respectively, The O/F of the propellant is given by the product of the
molar mass fraction and the stoichiometric ratio:

MO2 3
O/F = · (3.4)
M C2 H 4 1

which gives an O/F = 3.4218. The resulting mass flow rate from the fuel wall is
corresponding to the oxidizer mass flow and the O/F becomes 0.0066 kg/s.

3.5.3 Temperature

The inlet BC is given a ”fixedValue” temperature of 292 K. Walls are set to


”zeroGradient” except for the fuel wall which is set to ”fixedValue” at 600 K.

3.5.4 Starting the Simulation

In order to initiate the combustion the internal volume temperature is set to 2200
K and GOX to fill the entire volume at the start. Combustion is then sustained by
the incoming oxidizer mass flux from the inlet and the fuel mass flux from the fuel
walls.

The boundary conditions set for k-ωSST are presented in Table 3.5.1.

3.6 The Kinetics of Combustion

A reduced chemical kinetic mechanism model is implemented into the numerical


solver. To more accurately simulate the kinetics of the combustion a skeletal structure
of the chemical mechanism is utilized based on the works of Zettervall et al, [21]. The
model is then modified to work with pure O2 as oxidizer rather than with air which it
was made for originally.

Turbulence will generally influence the diffusion and therefore the rate of chemical
interactions. The effects of the so called turbulence chemistry interaction are omitted
in this early stage modelling.

22
CHAPTER 3. METHOD

Table 3.5.1: Boundary conditions applied for the k-ωSST turbulence model.

Boundary label Boundary Condition

Turbulent kinetic energy, k


Inlet turbulentIntensityKineticEnergyInlet
Outlet zeroGradient
Walls kqRWallFunction
Fuel wall kqRWallFunction

Turbulent dissipation rate, ω


Inlet fixedValue
Outlet inletOutlet
Walls omegaWallFunction
Fuel wall omegaWallFunction

Turbulent dissipation rate, ε


Inlet turbulentMixingLengthDissipationRateInlet
Outlet inletOutlet
Walls epsilonWallFunction
Fuel wall turbulentMixingLengthDissipationRateInlet

Turbulent viscosity, νt
Inlet fixedValue
Outlet inletOutlet
Walls nutkWallFunction
Fuel wall nutkWallFunction

3.7 Heat Transfer


For this thesis the effects of radiation are not modeled. This could be an area of interest
for further investigations into to the heat transfer for this case. The heat transfer
formulation applied to this model is simplified to solely take into account convection
to the fuel surface. The heat flux at the fuel wall is computed in the post processing
using the ”wallHeatFlux” utility in OpenFOAM.

23
Chapter 4

Results

Here various parameters of interest from the steady-state combustion results are
presented. The pressure and velocity figures are presented for two ranges to make
variations visible for both the entire model and for value range solely within the
combustion chamber, not accounting for the nozzle.

The pressure results show an overall reduction after adding a recirculation zone as can
be seen in Fig. 4.0.1, in Pa.

(a) Without recirculation (b) With recirculation

Figure 4.0.1: Pressure, in Pa, inside the combustion chamber and nozzle.

The axial velocity of the flow for the two cases are presented in m/s in Fig 4.0.2. The
flow in the nozzle exit shows a significant increase in velocity as you move in the radial
direction from the axis of symmetry, as shown in Fig. 4.0.2, and can be observed to
have more than twice velocity in the axial direction compared to the fluid near the
center line, before decreasing again as it gets closer to the nozzle wall.

The flow pattern inside the combustion chamber shows variations for the two cases
as can be seen in Fig. 4.0.3. In Fig. 4.0.3b one can observe how the flow enters the
recirculation zone with a negative axial velocity component and changing to a positive

24
CHAPTER 4. RESULTS

direction as it approaches the outer wall.

(a) Without recirculation (b) With recirculation

Figure 4.0.2: Overall axial velocity magnitude and velocity variations in the diverging
section of the nozzle in m/s.

(a) Without recirculation (b) With recirculation

Figure 4.0.3: Axial velocity with range with respect to values from inlet to nozzle throat
in m/s.

The flame zone, possesses a wave-like pattern as can be identified through the
temperature plots in Fig. 4.0.4. Temperatures are the coldest in vicinity to the
centerline and increases closer to the flame where it is the highest. The maximum
flame temperature and the general flow pattern can be seen to be similar for both
cases, although there appears to be a slight variation in the thickness and position of
the flame.

(a) Without recirculation (b) With recirculation

Figure 4.0.4: Overall temperature in K highlighting the flame zone.

In the proximity of the center line, most of the fluid flow consists of oxidizer, gaseous
oxygen (O2), that has not gone through any chemical reaction, as can be observed in
Fig. 4.0.5.

25
CHAPTER 4. RESULTS

(a) Without recirculation (b) With recirculation

Figure 4.0.5: O2 concentration through the chamber.

The concentraion of the products from the stoichiometric equation (eq. 3.3) are shown
in Fig. 4.0.6, which appear to follow the pattern of the flame as observed from Fig.
4.0.4.

(a) Without recirculation (b) With recirculation

Figure 4.0.6: CO2 and H2O concentrations through the chamber.

The carbon monoxide (CO) and hydroxide (OH) concentrations are presented in Fig.
4.0.7. The concentration of CO appears spread out between the flame and the fuel
wall with the highest concentration closer to the flame for both cases. Comparing Fig.
4.0.7a with Fig. 4.0.7b the case with a recirculation zone can be observed to have a
higher concentration in the head end than the case without one.

(a) Without recirculation (b) With recirculation

Figure 4.0.7: CO and OH concentrations through the chamber.

Concentrations of some of the chemical components of the reaction are shown in

26
CHAPTER 4. RESULTS

Fig. 4.0.8 and 4.0.9 and how they change along the flame. The values presented are
between the axis of symmetry and the parallel part of the fuel wall at five equally spaced
sections.

Without recirculation zone


C2H4
CH2O
CO
CO2
H2O
Radial distance

O2
OH
Temperature

Axial distance

Figure 4.0.8: Molecular concentration between fuel wall (bottom) and centerline (top)
without a recirculation zone.

With recirculation zone


C2H4
CH2O
CO
CO2
H2O
Radial distance

O2
OH
Temperature

Axial distance

Figure 4.0.9: Molecular concentration between fuel wall (bottom) and centerline (top)
with a recirculation zone.

The flame temperature over the fuel wall running parallel with the bulk flow
is presented in Fig. 4.0.10 for the two cases. One can observe fairly similar
temperatures, although the case without a recirculation zone appears to have stronger
fluctuations.

The convective heat transfer to the fuel wall surface parallel with the bulk flow for the
two cases is shown in Fig. 4.0.11.

27
CHAPTER 4. RESULTS

Figure 4.0.10: Maximum


Figure 4.0.11: Heat flux through wall
flow temperature between center line
parallel to center line for both cases.
and fuel wall for both cases.

28
Chapter 5

Conclusion

This thesis project presents the results for an early stage analysis of a hybrid rocket
engine in under steady state operation. This is done in order to acquire data on the
behaviour of the combustion in the engine to determine fuel regression to be able
to predict engine performance more accurately than by the use of semi-empirical
regression rate models. Preliminary results for the combustion mechanics for a
GOX/HDPE propellant combination have been presented. Results show a relatively
high flame temperature in the combustion chamber. Adding the recirculation zone
moves the flame closer to the fuel wall, increasing the convective heat flux to the fuel
surface and increases temperature near the inlet. Furthermore, the results indicate
that some of the oxidizer flows through the combustion chamber without mixing with
the fuel. Differences in exit velocity can be seen in the flame position is and where
there is unmixed oxidizer. Further improvements need to be made to the model before
accurate, quantitative regression rate results can be gathered.

5.1 Discussion

5.1.1 Regression Rate

Regarding the fuel regression rate, it would be interesting to either investigate the
results for different port areas or even implement fuel regression into the CFD model.
Two different ways of approach could be taken here. The first one is to run simulations
of an engine design with a static fuel wall, but different fuel wall diameters and
compare or interpolate results to find an average regression rate. The second approach

29
CHAPTER 5. CONCLUSION

would be by implementing a regression rate model and calculating the regression rate
and continuously update the model’s fuel wall diameter over time. Modelling the
regression rate would require a thorough look at the behaviour of the flow at the fuel
wall surface and the pyrolysis of the fuel grain. Emphasis should be put on the fact that
in this thesis, the only pyrolysis product coming from the wall is ethylene whereas in
reality it would be a combination of several polymers of varying length.

The results of the chemical components of the flow show oxidizer flowing through the
chamber and out the nozzle without reacting with the fuel. This seems to be due to
poor mixing of the propellant. The exit velocity can be seen to be significantly higher
closer to the nozzle walls than the flow around the center line. There appears to be a
correlation between the concentration of different molecules and the exhaust velocity.
Where there is a high concentration of oxygen in the center of the flow, molecular mass
is higher than for the flow surrounding it, where molecules of lower molar mass reside.
Further investigation into the injector design would be of interest in how it would effect
mixing and consequently regression rate. Modifying the inlet flow by changing the inlet
design, e.g. by having a swirl injector, could possibly improve propellant mixing.

Furthermore, it might be of interest to vary the oxidizer flux to see what effects on
the flow and combustion properties that would have. With studies showing how mass
flux affects the governing mechanisms of the regression rate, it would be of interest in
finding optimum mass flux in order to maximize performance.

5.1.2 CFD

Initiating the simulation with combustion was a cause for numerical instabilities before
steady-state flow could be achieved. The causes of which were primarily high gradients
from the transients occurring at the start of the simulation due to the sudden oxidizer
flow into the domain and the initial ignition. Various approaches were tried to find
ways to circumvent this issue. In the end, having the oxidizer flow rate increase
gradually at the start helped with this and worked best.

Other approaches that showed to be promising were to start with a ”cold flow”
simulation. That is, starting the simulation by injecting oxidizer and fuel into the
domain with chemistry turned off. After letting the simulation run until the flow is
stable, then, chemistry was turned on and the domain temperature increased to 2000 K
in order to start the combustion. This, unfortunately resulted in instabilities as well and

30
CHAPTER 5. CONCLUSION

this method of approach was abandoned. A solution that possibly could mitigate this
problem could be to not increasing the temperature in the entire domain to such high
degrees and instead have local high temperature zones at location where the propellant
has started to mix.

The CFD model used in this project has been of relatively low resoultion in terms
of geometry detail and mesh. Further analysis will require a refined more model
achieving adequate results. As geometry complexity increases other mesh generating
utilities than ”blockMesh” might be more suitable. Increasing cell density in local
regions of the mesh without distorting the cells in other regions can be somewhat
challenging when working with only hexahedral cells, with the requirement that each
and every cell face is only connected to one face (that of the adjacent cell).

5.1.3 Flow and Heat Transfer

A recirculation zone after the oxidizer injection is observed to affect the flow in several
ways. It is interesting to find that the recirculation vortex seems to have a flow direction
opposite to the expected one. Changing the injection of the flow and/or changing the
geometry may be done to modify the recirculating flow.

As indicated by Fig 4.0.11, there is a significant increase in convective heat flux to


the wall with the addition of the recirculation zone after the inlet. As shown in Fig.
4.0.10, the flame temperatures are relatively similar for both cases. However, the
flame position appears closer to the fuel wall for the case with the recirculation zone,
as indicated by Fig. 4.0.4. Furthermore, as Fig. 4.0.3 suggests, the flow pattern differs
between the cases differs. These two factors could arguably explain the difference in
heat flux.

The effects of radiation and its contribution to the heat transfer to the fuel surface
would be worth investigating. One could argue the radiation to add a relatively small
contribution to the heat transfer since the fuel is nonmetallized. On the other hand,
the flame temperature appears relatively high which could imply high radiative heat
transfer. Soot is also a possible cause of radiation.

Using a skeletal reaction mechanism for the kinetics is expected to give a more accurate
representation of the combustion taking place than having a simple stoichiometric
model. It is observed that inside the combustion chamber temperatures are reaching

31
CHAPTER 5. CONCLUSION

up to and over 4000 K. One could reason that a stoichiometric reaction model should
suffice for early stages of analysis. It would likely result in a different temperature
profile and change the heat flux. It might be of interest to compare different kinetics
models and see how results would differ.

32
Bibliography

[1] Altman, D and Humble, R. “Hybrid rocket propulsion systems”. In: Space
propulsion analysis and design 379 (1995).

[2] Altman, David. “Hybrid rocket development history”. In: 27th Joint Propulsion
Conference. 1991, p. 2515.

[3] Altman, David and Holzman, Allen. “Overview and history of hybrid rocket
propulsion”. In: Progress in Astronautics and Aeronautics 218 (2007), p. 1.

[4] Chiaverini, Martin. “Review of solid-fuel regression rate behavior in classical


and nonclassical hybrid rocket motors”. In: Progress in Astronautics and
Aeronautics 218 (2007), p. 37.

[5] Faenza, M, Barato, F, Lazzarin, M, and Pavarin, D. “Hybrid rocket motors


regression rate prediction through CFD simulations”. In: 6th Eur. Conf. for
Aerospace Sciences. 2015.

[6] k-omega Shear Stress Transport (SST). https : / / www . openfoam . com /
documentation / guides / latest / doc / guide - turbulence - ras - k - omega -
sst.html. Accessed: 2022-11-16.

[7] Kuo, Kenneth K and Chiaverini, Martin J. Fundamentals of


hybrid rocket combustion and propulsion. American Institute of Aeronautics
and Astronautics, 2007.

[8] Lazzarin, M, Barato, F, Bettella, A, and Pavarin, D. “Computational fluid


dynamics simulation of regression rate in hybrid rockets”. In: Journal of
Propulsion and Power 29.6 (2013), pp. 1445–1452.

[9] Marić, Tomislav, Höpken, Jens, and Mooney, Kyle G. The OpenFOAM®
Technology Primer. sourceflux UG, 2014.

33
BIBLIOGRAPHY

[10] Marquardt, Timohty and Majdalani, Joseph. “A Primer on Classical Regression


Rate Modeling in Hybrid Rockets”. In: AIAA Propulsion and Energy 2020
Forum. 2020, p. 3758.

[11] Marxman, G and Gilbert, M. “Turbulent boundary layer combustion in the


hybrid rocket”. In: Symposium (International) on Combustion. Vol. 9. 1.
Elsevier. 1963, pp. 371–383.

[12] Marxman, GA, Wooldridge, CE, and Muzzy, RJ. “Fundamentals of hybrid
boundary-layer combustion”. In: Progress in Astronautics and Rocketry.
Vol. 15. Elsevier, 1964, pp. 485–522.

[13] Marxman, Gerald A.


“Boundary-layer combustion in propulsion”. In: Symposium (international) on
Combustion. Vol. 11. 1. Elsevier. 1967, pp. 269–289.

[14] Menter, F. R. “Two-equation eddy-viscosity turbulence models for engineering


applications”. eng. In: AIAA journal 32.8 (1994), pp. 1598–1605. ISSN: 0001-
1452.

[15] Menter, Florian R, Kuntz, Martin, and Langtry, Robin. “Ten years of industrial
experience with the SST turbulence model”. In: Turbulence, heat and mass
transfer 4.1 (2003), pp. 625–632.

[16] Strand, L, Ray, R, and Cohen, N. “Hybrid rocket combustion study”. In: 29th
Joint Propulsion Conference and Exhibit. 1993, p. 2412.

[17] Sutton, George P and Biblarz, Oscar. Rocket propulsion elements. eng. 9th. New
York: Wiley, 2016. ISBN: 1118753658.

[18] Tu, Jiyuan, Yeoh, Guan Heng, and Liu, Chaoqun. Computational fluid
dynamics: a practical approach. 3rd. Butterworth-Heinemann, 2018.

[19] Versteeg, Henk Kaarle and Malalasekera, Weeratunge. An introduction to


computational fluid dynamics: the finite volume method. 2nd. Pearson
education, 2007.

[20] White, Frank M. Fluid Mechanics. eng. 9th. Boston: McGraw-Hill, 2021. ISBN:
978-1-260-57554-5.

[21] Zettervall, Niklas, Fureby, Christer, and Nilsson, Elna JK. “Small skeletal kinetic
reaction mechanism for ethylene–air combustion”. In: Energy & Fuels 31.12
(2017), pp. 14138–14149.

34

You might also like