Anda di halaman 1dari 105

Numerical Modelling of Staged Combustion

Aft-Injected Hybrid Rocket Motors

by

Jeff Nijsse

A thesis submitted in conformity with the requirements


for the degree of Masters of Applied Science

Graduate Department of Aerospace Science and Engineering


University of Toronto

Copyright 2012 by Jeff Nijsse

Abstract
Numerical Modelling of Staged-Combustion Aft-Injected Hybrid Rocket
Motors
Jeff Nijsse
Masters of Applied Science
Graduate Department of Aerospace Science and Engineering
University of Toronto
2012

The staged combustion aft-injected hybrid (SCAIH) rocket motor is a promising design for the
future of hybrid rocket propulsion. Advances in computational fluid dynamics and scientific computing have made computational modelling an effective tool in hybrid rocket motor design and
development. The focus of this thesis is the numerical modelling of the SCAIH rocket motor in
a turbulent combustion, high-speed, reactive flow framework accounting for solid soot transport
and radiative heat transfer. The SCAIH motor is modelled with a shear coaxial injector with liquid
oxygen injected in the center at sub-critical conditions: 150 K and 150 m/s (Mach 0.9), and a
gas-generator gas-solid mixture of one-third carbon soot by mass injected in the annual opening
at 1175 K and 460 m/s (Mach 0.6). Flow conditions in the near injector region and the flame
anchoring mechanism are of particular interest. Overall, the flow is shown to exhibit instabilities
and the flame is shown to anchor directly on the injector faceplate with temperatures in excess of
2700 K.

ii

When everything goes wrong, what a joy to test your soul


and see if it has endurance and courage!

Zorba the Greek

iii

Acknowledgements

I would first like to acknowledge Professor C.P.T. Groth, for providing me with an opportunity to
study under his supervision, and for introducing me to turbulent combustion. I would also like to
extend gratitude to the University of Toronto and Prof. C.P.T. Groth for providing funding for this
research experience.
Additionally, thank you to my colleagues in the UTIAS CFD lab who have taken the time to help me
over myriad hurdles encountered along the way; principally Dr. Pradeep Jha and Dr. Marc Charest.
Computational resources for performing all of the calculations reported herein were provided by the
SciNet High Performance Computing Consortium. SciNet is funded by: the Canada Foundation for
Innovation under the auspices of Compute Canada; the Government of Ontario; Ontario Research
Fund Research Excellence; and the University of Toronto.

University of Toronto Institute for Aerospace Studies


JEFF NIJSSE

Toronto, 2012

iv

Contents

Abstract

ii

Acknowledgements

iv

Contents

List of Tables

viii

List of Figures

ix

1 Introduction

1.1 Rocket Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.1.1 Rocket Fuel Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.1.2 Hybrid Rockets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1.1.3 Staged Combustion Aft-Injected Hybrid . . . . . . . . . . . . . . . . . . . . . . . .

1.2 Research Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11


1.3 Scope of Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2 Axisymmetric Flow Model

13

2.1 Shear Coaxial Injector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13


2.2 Flame Anchoring Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 Combustion Chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Grid Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.5 The Navier-Stokes Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3 Turbulent Combustion

21
v

3.1 Turbulence Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22


3.2 Reynolds-Averaging Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.3 Favre-Averaging Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.3.1 The Favre-Averaged Navier-Stokes Equations . . . . . . . . . . . . . . . . . . . . 25
3.4 k Turbulence Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.5 Conservation of Species . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.5.1 Thermodynamic Closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.6 Chemical Kinetics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
3.6.1 Eddy Dissipation Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4 Radiative Heat Transfer

34

4.1 Radiation Transfer Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35


4.2 Optically Thin Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.3 Spectral Absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.3.1 Absorption Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
4.3.2 Statistical Narrow Band Correlated-k Model . . . . . . . . . . . . . . . . . . . . . 38
4.4 Soot Radiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4.5 Solution to the Radiation Transfer Equation . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5 Gas-Generator Propellant, Chemical Kinetics, and LOX Expansion

43

5.1 SCAIH Operating Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44


5.1.1 Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.1.2 The Critical Point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.2 Gas-Generator Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.2.1 Simplified Chemical Kinetic Scheme for Propellant Gases . . . . . . . . . . . . . 48
5.2.2 Treatment of Soot and/or Solid Carbon Particles . . . . . . . . . . . . . . . . . . 49
5.3 Liquid Oxygen Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.3.1 Peng-Robinson Equation of State . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6 Numerical Solution Method

52

6.1 The Finite Volume Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53


6.2 2-D Axisymmetric Form of Navier-Stokes Equations . . . . . . . . . . . . . . . . . . . . . 53
vi

6.3 Integral Form of Navier-Stokes Equations and FVM Formulation . . . . . . . . . . . . . 55


6.3.1 Cell Reconstruction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
6.3.2 Flux Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.3.3 Time Marching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.4 Radiation Transfer Equation Evaluation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.4.1 Discrete Ordinates Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.5 Parallel Architectures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
7 Numerical Simulation Results

64

7.1 Baseline Input Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64


7.2 Mesh Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
7.3 Flow Field Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
7.4 Flame Anchoring Mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
7.5 Chamber Wall Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.6 Effects of Solid Carbon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
7.7 Effects of Chlorine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
7.8 Effects of Radiative Heat Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
8 Conclusions and Future Research

82

8.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
8.2 Areas of Future Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
References

85

vii

List of Tables
1.1 Propellant configuration of heavy-lift launch vehicles . . . . . . . . . . . . . . . . . . . .

1.2 Rocket fuel properties when combined with LOX as the oxidizer . . . . . . . . . . . . .

1.3 Hybrid rocket propellant combinations by Isp performance . . . . . . . . . . . . . . . . .

3.1 Detailed Chemistry Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30


5.1 Thermodynamic fluid properties at the critical point listed by the mass fraction
present in the combustion chamber at the injector . . . . . . . . . . . . . . . . . . . . . . 46
5.2 Solid fuel grain gas-generator components . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.3 Generator gas composition under expansion from p = 5 MPa to p = 4 MPa . . . . . . . 47
5.4 Multi-species one-step reaction mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . 48
7.1 Numerical simulation input conditions for LOX and fuel gases . . . . . . . . . . . . . . 65
7.2 Multi-species reaction mechanism without chlorine . . . . . . . . . . . . . . . . . . . . . 78

viii

List of Figures
1.1 Government spending on NASA and CSA as a percentage of total program expenditure 2
1.2 Regression rate behavior in a standard configuration hybrid rocket motor . . . . . . .

1.3 Schematic of a swirl-injected end-burning hybrid . . . . . . . . . . . . . . . . . . . . . .

1.4 Traditional and SCAIH hybrid rocket schematic . . . . . . . . . . . . . . . . . . . . . . . . 10


2.1 Shear coaxial rocket injector photo and schematic by CTI . . . . . . . . . . . . . . . . . 14
2.2 Flame anchoring mechanism for a coaxial injector for methane/LOX . . . . . . . . . . 15
2.3 OH emission for a LOX/CH4 flame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.4 Sub-scale combustion chamber schematic used for testing a single shear coaxial injector 16
2.5 Computational domain with a detailed injector close-up . . . . . . . . . . . . . . . . . . 18
3.1 Laminar diffusion flame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2 Time scales in a turbulent combusting flow . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.1 Cylindrical coordinate system used for radiative heat transfer . . . . . . . . . . . . . . . 35
5.1 Pressure trace for a typical test firing conducted at CTI . . . . . . . . . . . . . . . . . . . 44
6.1 Two-dimensional cell on a structured grid . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
6.2 Flux discontinuity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.3 One-dimensional representation of computational cell and finite-volume procedure . 57
6.4 The Riemann problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
7.1 The near-injector region showing different levels of mesh resolution . . . . . . . . . . . 67
7.2 Mesh resolution effects on temperature, density, and O2 mass fraction . . . . . . . . . 68
7.3 Energy convergence history for a Level 4 mesh compared with a Level 5 mesh . . . . 69
ix

7.4 Temperature flow contours for steady-state, quasi-unsteady and unsteady solutions . 70
7.5 CH4 mass fraction flow contours for steady-state, quasi-unsteady and unsteady solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
7.6 Pressure flow contours for steady-state, quasi-unsteady and unsteady solutions . . . . 72
7.7 Flame anchoring mechanism showing recirculation zones and temperature on the
LOX post . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
7.8 Combustion chamber wall temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.9 Effects of carbon particulate in the fuel stream . . . . . . . . . . . . . . . . . . . . . . . . 77
7.10 Flame temperature effects of chlorine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.11 Effects of radiative heat transfer on the temperature field . . . . . . . . . . . . . . . . . 81

Chapter 1

Introduction
The Space Shuttle Atlantis flew its final mission on July 8th, 2011 marking the end of NASAs Space
Transportation System (STS). The Shuttle program has been in operation since 1982 with Atlantis
being one of seven in NASAs fleet of reusable human-rated rockets.
The cost of the STS quickly ballooned beyond its budget and could not realize its original economic
forecast nor projected mission frequency. In the 1970s, NASA planned for a total of 581 Shuttle
launches for the program to be profitable [1] but did not have the manpower to launch at this
high frequency (about once a week) and had to scale operations back to 8 launches per year for
safety, technical and training reasons [2]. Depending on the estimate, in 2008 the cost of one
Shuttle launch was between US$450 million (excluding infrastructure, research, and development
costs) [3] and US$1.5 billion (cost for the entire program divided by number of flights) [4].
Part of the reason rocket launches are expensive is due to the production and burning of large
amounts of propellant. The Space Shuttle is capable of placing large payloadsup to eight astronauts and 22700 kginto low Earth orbit and this requires 1 million kg of solid fuel and 0.735
million kg of liquid fuel. Rocket fuel is largely hydrocarbon based and in the case of fuels such as
liquid oxygen (LOX) and liquid hydrogen (LH2) extremely large amounts of energy are needed for
their extraction and production through electrolysis. Between cost and energy usage large scale
rocket programs such as the STS are being scaled down or retired completely by governments that
can no longer justify their substantial needs. Figure 1.1 shows the declining government spending on space programs for NASA and the Canadian Space Agency as a percentage of total budget
expenditure.
This decline in spending has opened an opportunity for the private sector. Presently a market is
emerging for smaller-scale rockets; used for micro-satellite launches, individual astronaut transport,

CHAPTER 1. INTRODUCTION

Government spending on NASA and CSA as a percentage of total program expenditure beginning
in 1992 when the CSA was created. Note: 2010 figures are projections. Sources: [58].

and space tourism. For example, Cesaroni Technology Incorporated (CTI) of Canada has been
developing a hybrid rocket to put small payloads into low Earth orbit. The mission objectives for
CTIs small launch vehicle are to place a payload of 100 kg into orbit at an altitude of 700 km within
a target budget of CAD$5 million per launch [9]. Satellite launching costs of this magnitude will
make it much easier and more attractive for the private sector to participate in the space industry
in the future [10].

1.1

Rocket Technology

The rocket propulsion principle is based on Newtons third lawfor every action there is an equal
and opposite reactionand is simply based on a pressure differential between a control volume
and the environment. All chemical rockets operate by mixing a fuel and an oxidizer together and
burning them in a combustion chamber. Energy stored in the chemical bonds of the propellants is
released during combustion which creates hot gas that rapidly expands through a nozzle imparting
momentum on the rocket creating propulsion. Of chemical rockets, three classifications emerge:
liquid, solid, and hybrid.

Liquid Rockets
Currently the worlds high performance rocket engines capable of placing large payloads into orbit
use LH2 as the primary source of fuel. Rocket motors such as the Space Shuttle main engine

CHAPTER 1. INTRODUCTION

(SSME), Ariane 5 main engine, and Vulcain 2 all employ LH2 as the fuel and LOX as the oxidizer.
LH2 and LOX need to be stored cryogenically in pressure vessels that can withstand the force of
their liquidized state. Additionally, complex turbomachinery is needed to mix the propellants and
to handle high mass flow rates. These factors combine to add to the cost and mass of the launch
vehicle. The handling of liquid propellants is very dangerous and safety issues must be considered
with transport, storage, and possibility of failure.
These shortcomings of liquid rockets are made up for by their throttling capability and performance.
Controlling the flow rates of fuel into the combustion chamber is a method of throttling and can
easily be done with turbo pumps. In this manner the Space Shuttle throttles down as the vehicle
approaches Max-Q, maximum aerodynamic pressure, and then once this point is passed and the
atmosphere has thinned the Shuttle throttles back up. Rocket performance will be discussed along
with fuel subsequently.

Solid Rockets
Solid rockets originated in China in the 13th century and have their fuel and oxidizer stored in a
solid state and burn in place rather than in a separate combustion chamber. Solid rockets operate
on the same principle as fire crackers. Having a consistency similar to that of ordinary candle
wax, solid rocket fuel is a mixture of fuel (usually hydrocarbon based), oxidizer such as ammonium
perchlorate (NH4 ClO4 ) , and a binder to glue everything together. This is much safer for production
and storage yet lacks the benefit of throttling. Solid rockets can be much smaller than liquid
rockets as they do not need turbomachinery or pressurant tanks. A myriad of examples of solid
rocket motor application are found in the military including surface-to-air missiles, intercontinental
ballistic missiles, and rocket propelled grenades.
Before hybrid rockets are discussed it is convenient to discuss rocket fuel performance and to outline
a comparison between liquid and solid rocket motors.

1.1.1

Rocket Fuel Performance

Specific Impulse
Specific impulse, Isp , is an attribute of propellants which is related to the available energy contained
in the fuel. Measured in seconds, specific impulse is the amount of time a unit mass of propellant

The solid rocket boosters of the Space Shuttle are a mixture known as ammonium perchlorate composite propellant
(APCP). Propellant chemical composition is discussed in Section 5.2.

CHAPTER 1. INTRODUCTION

can sustain a unit of thrust. In this manner various fuels as well as oxidizer combinations can be
compared.
The impulse per mass of propellant burned is the integral of the thrust produced
I=

tb

F dt = Muex

(1.1)

where F is the force of thrust the exhaust gas exerts on the rocket, M is the total mass of burned
fuel at time t b , and ue x is the exhaust gas velocity. Specific impulse is the amount per unit mass of
propellant per time and is generally divided by gravity to simplify the units to seconds. Additionally,
the relation to the specific energy of fuel, Esp , can be viewed as

Isp =

uex
ge

p
=

2Esp
ge

(1.2)

is the mass flow rate of propellant.


Here g e is the gravitational force at the Earths surface and m
Liquid rockets using molecular hydrogen and oxygen have the highest theoretical Isp in the chemical
regime which approaches 460 s under optimum expansion conditions in a vacuum. The maximum
is limited by the amount of the energy that can be released by breaking the chemical bonds of
the propellants. Some research is being done on metallic hydrogen which could have a theoretical
Isp as high as 1700 s [11]. This would have a profound impact on rocketry; however, metallic
hydrogen has yet to be observed in a laboratory environment and is in the early stage of research.
Solid rocket technology can be scaled easily for large launch vehicle applications such as the Atlas
V and Ariane 5 vehicles. Table 1.1 lists some current heavy-lift launch vehicles with their engine
configuration and comparative performance. All of these examples have liquid rocket engines and
only the Russian Soyuz-FG vehicle does not use solid rocket boosters (SRBs) during liftoff.
Advances in rocket propulsion have been tending towards replacing liquid hydrogen with hydrocarbon based propellants. Hydrocarbons offer a greater energy density due to the higher molecular
weight of carbon and the molecules. When comparing methane to LH2, methane has a much
greater energy density even though having a lower specific impulse. This is because of the low
molecular weight of hydrogen; very large tanks are needed to make up for the lack of density (consider the size of the SSME tank). Rocket Propellant #1 (RP-1), which is a highly refined version
of kerosene was developed by Rocketdyne in the United States in the 1950s and is widely used
in rocketry. Compared to RP-1, methane has a slightly higher Isp while occupying a midpoint in
density between LH2 and RP-1 yet still needing cryogenic storage. Since RP-1 is much cheaper to
produce than LH2 or liquid methane and can be stored at room temperature it has been a success-

CHAPTER 1. INTRODUCTION

Launch Vehicle

Engine

type

Isp (vacuum)

fuel/oxidizer

Space Shuttle

SRB
SSME

solid
liquid

292
453

APCP
LH2/LOX

Ariane 5 ECA

SRB
Vulcain 2

solid
liquid

273
431

APCP
LH2/LOX

Atlas V

RD-180
Aerojet 73F
RL 10A

liquid
solid
liquid

338
275
450

RP-1/LOX
APCP
LH2/LOX

Soyuz-FG

RD-117
RD-0124

liquid
liquid

310
359

RP-1/LOX
RP-1/LOX

Propellant configuration of heavy-lift launch vehicles. All are in current operation except the Space
Shuttle which was retired as of 21 July, 2011. Note: Solid rocket fuel contains a mixture of fuel
and oxidizer. APCP is ammonium perchlorate composite propellant. RP-1 is Rocket Propellant #1.

Hydrogen
H2
Isp (vacuum), s
melting point, K
(liquid) at Tb , k g/m3

455
13.8
70.8

Methane
CH4

RP-1
(C1 H1.96 )n

369
90.69
422.4

358
224
808

Rocket fuel properties when combined with LOX as the oxidizer. Notes: The properties of RP-1 are
an average value. Hydrocarbon fuels are a mixture of different molecules; the larger the variety the
more impure, and less expensive, the fuel. Tb is temperature at the boiling point. Sources: [1214].

ful liquid fuel. Kerosene, or jet fuel, is often represented by n-dodecane, (C12 H26 )n , and does not
perform as well as RP-1. Table 1.2 lists characteristics of these fuels.
In comparsion to standard liquid and solid chemical rockets, hybrid rockets can take advantage of
greater combinations of fuel and oxidizer while maintaining comparative performance.

1.1.2

Hybrid Rockets

Hybrid rocket technology has been around since as early as 1929 in Germany [15]. It has a number
of differences in comparison to liquid and solid technology. A hybrid rocket motor uses the combination of the two phases of propellant together. A traditional configuration has the fuel as a solid
and the oxidizer as a liquid with a reverse hybrid having the opposite.
The advantages of a hybrid rocket can be outlined as follows [16]:

CHAPTER 1. INTRODUCTION

Safety: Hybrid rocket fuel in a traditional form is solid which allows for safe manufacturing,
separate from oxidizer and, once cured, for easier and safer transport.
Throttling: Overall rocket performance can be throttled by adjusting the flow rate of the
liquid oxidizer. This is simpler than throttling of a liquid rocket where there are two flows
that must be coordinated (and an additional pump and valve control assembly). The flow of
fuel in the hybrid adjusts naturally because it is directly related to the oxidizer that comes
into contact with vaporized fuel. In an abort situation the rocket may be throttled off by
stopping the flow of oxidizer.
Grain Stability: Cracks in a solid rocket fuel grain can lead to instability, high pressures, and
explosion. Hybrid rockets avoid this scenario because burning only occurs where there is a
mixing of fuel and oxidizer (generally in the boundary layer on the surface of the fuel).
Propellant Choice: There is more versatility in propellant selection compared to liquid or
solid rockets. As seen in Table 1.1 liquid rockets generally use either RP-1 or LH2 fuel and
LOX oxidizer. Most conceivable combinations of hybrid fuels have been tested and can be
compared by Isp . Table 1.3 shows the diversification of hybrid propellant combinations ordered by their performance. Solid fuel enhancement, with substances like energetic metals,
can be done without creating a dangerous slurry, further varying the options.
Temperature: A solid rocket must be designed for a maximum expected operating pressure
which can be unsteady due to ambient launch temperature conditions. This design constraint
is lessened in a hybrid because the temperature effect on burning rate is small with liquid
combustion.
Cost: The cost of the overall launch vehicle, including development, manufacture, and transport of propellant is reduced by the safety advantage of inert fuel. Additionally, the vehicle
has fewer components than a liquid rocket that has turbo pumps for both fuel and oxidizer.
Environment: There are environmental advantages based on the exhaust gas products. The
hazardous components in solid propellants are the aluminum and chlorine which are not
present in a classical hybrid motor.

The numerous advantages of a hybrid rocket motor configuration speak for themselves and have
provided the impetus for further research into the design of hybrid systems to overcome the following deficiencies.
The disadvantages of a hybrid rocket motor configuration are as follows:

CHAPTER 1. INTRODUCTION

Fuel
Li/LiH/HTPB
HTPB
HTPB/Al
Ammonia(s)/Be
Methane(s)/Be
Methane(s)
CTI gas-generator
Kerosene
HTPB
Polyethylene
Pentane(s)
HTPB/Al
HTPB/Al
PMM
Kerosene
HTPB
HTPB/Al
Carbon(s)
Kerosene
HTPB
HTPB
Polyethylene
Cellulose (wood)
Carbon(s)
Carbon(s)

Chemical Formula

Oxidizer

O/F ratio

Isp (sea level)

Li+LiH+HTPB
(OH)2 (C4 H6 O0.1 )n
(0.6)HTPB+(0.4)Al
(0.74)NH3 +(0.26)Be
(0.64)CH4 +(0.36)Be
CH4
mixture
(C12 H26 )n
(OH)2 (C4 H6 O0.1 )n
(C2 H4 )n
C5 H12
(0.6)HTPB+(0.4)Al
(0.6)HTPB+(0.4)Al
(C5 H8 O2 )n
(C12 H26 )n
(OH)2 (C4 H6 O0.1 )n
(0.6)HTPB+(0.4)Al
C
(C12 H26 )n
(OH)2 (C4 H6 O0.1 )n
(OH)2 (C4 H6 O0.1 )n
(C2 H4 )n
C6 H10 O5
C
C

FLOX
FLOX
FLOX
LOX
LOX
LOX
LOX
LOX
LOX
LOX
LOX
LOX
N2 O4
LOX
N2 O4
N2 O4
N2 O
LOX
N2 O
N2 O
RFNA
N2 O
O2
N2 O
Air

2.8
3.3
2.5
0.5
1.3
3.0
1.4
2.5
1.9
2.5
2.7
1.1
1.7
1.5
4.0
3.5
3.5
1.9
8.0
7.1
4.3
8.0
1.0
6.3
11.3

326
314
312
307
306
291
290
281
280
279
279
274
261
259
259
258
252
249
248
247
247
247
247
236
184

Hybrid rocket propellant combinations by Isp performance. Notes: HTPB is hydroxyl-terminated


polybutadiene, RFNA is red fuming nitric acid: (0.15)N2 O+(0.85)HNO3 , PMM is polymethyl
methacrylate and FLOX is fluorine oxidizer which is 32 F2 + 31 O2 . CTI gas-generator is a unique configuration to be discussed in Section 5.2. Sources: [16, 17].

Regression Rate: The hybrid fuel regression rate is low due to the speed at which polymeric
fuels burn. A lower regression rate means the fuel grain surface area must be larger for the
rated thrust. Complex fuel grains, wagon wheels, are required with multiple ports which may
need additional supporting structure. Requirements for numerous ports and corners in grain
geometry lead to fuel slivers that form and break off the grain, remaining unburned. Slower
burning can, nevertheless, be desirable for applications such as a gas-generator.
Combustion Efficiency: There is a lower degree of mixing because of the diffusion flame between the vaporizing fuel and oxidizer which creates a lower impulse efficiency. Performance
can still be higher compared to solid rockets due to a higher theoretical Isp .
Oxidizer to Fuel (O/F) Ratio: A shift in the O/F ratio happens as new ports within the fuel

CHAPTER 1. INTRODUCTION

Regression rate behavior in a standard configuration hybrid rocket motor. Fluctuation in the rate is
seen as the flame progresses from the leading edge of the fuel slab, x/L 0.1 to x/L 0.9. From
Chiaverini et al. [18].

grain open up. This is a loss of performance because surface area and thus mass flow rate of
the fuel are in constant flux.
Ignition Transient: Slower transients for ignition and throttling response are present in
hybrids. This is viewed as a minor disadvantage.

Fuel Regression
Fuel regression rate is the speed at which a solid fuel is consumed by the combustion process.
When applied to solid rockets, combustion products are produced at the surface of the fuel grain;
however, in a hybrid situation the fuel may have to vaporize, mix with oxidizer and then burn,
resulting in products above the fuel surface, at the boundary layer. Because of the slower burning
nature of polymeric fuels, regression rate and stability are key characteristics in the design of a
hybrid system. Variable regression rates cause shifts in the O/F ratio and can lead to combustion
instabilities. Figure 1.2 shows a typical hybrid rocket regression rate fluctuating along the fuel burn
direction.
To enhance regression rates there has been research into alternate injection methods as well as
cryogenic and additive-based fuels. Vortex hybrids have been investigated by Lee [19] which inject
LOX just upstream of the nozzle and perpendicular to the nozzle expansion to create a swirl in
the combustion chamber. While increasing regression rates by a baseline factor of 2 it is yet to be
seen if this can scale outside the laboratory. Experiments with a swirl injector at the fore end of
the combustion chamber have been done by Yuasa et al. [20] and helical cut grooves by Lee [19]
have not had as promising results. An axial-injection end-burning hybrid has been developed by

CHAPTER 1. INTRODUCTION

Schematic of a swirl-injected end-burning hybrid with a single fuel disk for spacecraft maneuvering
application.

Nagata et al. [21] whereby oxidizer is injected axially through tubes inside the chamber near the
nozzle which then ablate away upstream during the burn. Nagatas motor has good regression
rates, however, scaling is again a factor. A swirl-injected end-burning hybrid by Rice et al. [22] is
shown in Figure 1.3 which could have application for planetary ascent. Haag et al. has a similar
configuration called the vortex flow pancake configuration for spacecraft maneuvering [23] which
has an injector sandwiched between two fuel disks.
Another configuration that uses two stages of combustionthe first stage for gas generation and
the second stage for propulsionand the focus of this thesis, will now be introduced.

1.1.3

Staged Combustion Aft-Injected Hybrid

A variant to the traditional hybrid configuration is called a staged combustion aft-injected hybrid
(SCAIH). The schematic for a standard hybrid rocket is seen in Figure 1.4a. This shows the oxidizer
being injected into the solid fuel grain cavity for a single stage combustion. The SCAIH rocket
motor burns a solid fuel grain independently and uses this fuel-rich product gas, mixed and burnt
with an oxidizer, as the propulsive force in a second stage of combustion [9]. Using LOX oxidizer,
the solid fuel grain differs from a standard hybrid because it is mixed with its own oxidizer to
burn independently, similar to a solid rocket. Figure 1.4b shows the SCAIH schematic in which a
solid fuel grain burns in an independent stage and then combines the product gases with LOX in
a second stage. Such a configuration is of interest to CTI for the propulsion system of a possible
launch vehicle for microsatellite application.
This configuration is not new and was explored briefly in the past few decades. In 1990 NASA
awarded a contract to Atlantic Research Corporation to study a possible replacement for the Shut-

CHAPTER 1. INTRODUCTION

(a) traditional hybrid

10

(b) SCAIH hybrid

Hybrid rocket schematic showing the traditional configuration which has the oxidizer injected into
the solid fuel grain (a) and the SCAIH configuration which has the oxidizer injected aft of the solid
fuel grain combustion (b).

tles SRBs which resulted in selection of a gas generator hybrid cycle [24]. A direct comparison
was done a year later between a fore-injected hybrid (classical) and an aft-injected hybrid (SCAIH)
showing a reduced payload delivery cost for the SCAIH [25]. In 2001, testing of subscale components intended for use as an upper stage was undergone by Lund et al. [26] and Markopoulos and
Abel [27] using hydrogen peroxide oxidizer with AP/HTPB/PPG . These few studies have demonstrated enough potential for SCAIH vehicles research to continue.
The SCAIH motor is more scalable than a conventional hybrid as it does not suffer from complex
grain geometries that deteriorate in performance as size increases. The first stage of combustion is
cooler in comparison, 1170 K, than a single stage hybrid, 3660 K, since the gas-generator burns fuel
rich [28]. In terms of theoretical performance, the SCAIH has a Isp = 290 at sea-level at an optimum
O/F ratio of 1.4 which is competitive among traditional hybrid fuel/oxidizer combinations (see
Table 1.3). This alternate configuration has the potential to address the shortcomings of classical

PPG is polypropylene glycol and is used as a binding agent

CHAPTER 1. INTRODUCTION

11

hybrids and provide an affordable solution for launch vehicles carrying small to medium sized
payloads [9].

1.2

Research Motivation

The field of computational fluid dynamics (CFD) is an important research area that has revolutionized the aerospace industry. Any design parameters that can be changed and tested computationally
provide significant savings in time and cost to the industry as well as limit safety risks. These factors alone have been enough to propel CFD research over the past 20-30 years. High pressure
rocket combustion involves branches of CFD that alone are major areas of research which are constantly being refined. With increased computing power simulations that previously were too computationally expensive are being conducted and different branches are starting to come together.
Additionally, parallel computing and efficient algorithm development play important roles in a representative CFD simulation. Application to rocketry and the aerospace industry offers some of the
greatest benefit because of the high cost associated with testing and failure of rocket components.
While hydrocarbon fueled rockets, including SCAIH rockets, have advantages over their larger
liquid fueled counterparts, there is significantly less experimental data available for their development. Experimental test cases are needed for validation with numerical methods and vice versa.
Initial research and experimental evidence by CTI has indicated excessive heating in the injector
and near-injector region during test runs. These tests were not able to run to completion (i.e., a
full-cycle burn), due to ablation of the injector and near-injector materials at temperatures in excess of 3000 K. Predicting and understanding this thermal loading provides the motivation for this
thesis and the modelling of the flame anchoring mechanism near the injector post tip. An accurate CFD combustion model of this flow field will need to incorporate chemical kinetics, radiation,
and soot modelling to be representative. Much of the hybrid rocket research discussed above has
been conducted in the 1990s onward showing the relative youth of the hybrid rocket field and the
opportunity it has for advancement.
The field of CFD applied to rockets in general is too broad to review comprehensively within the
content of this thesis. However, there are several recent numerical studies focusing on hybrid
rocket motors worth mentioning. Farmer et al. [29, 30] have modelled a hybrid combustion system
including a simplified fuel pyrolysis model. Chen [31] has developed a numerical model with realfluid properties and finite-rate chemistry. Other studies emphasizing gas-generator ignition [32]
and hybrid propellant combustion [33, 34] have been done within the past 10 years. The main
focus of these previous studies has been on solid fuel pyrolysis whereas the present work focuses

CHAPTER 1. INTRODUCTION

12

on LOX injector operation.

1.3

Scope of Research

This thesis represents the synthesis of many different subdisciplines within CFD to model the combustion of a SCAIH rocket motor. The geometric configuration and initial conditions have been
outlined by CTI with the goal of overcoming some of the issues faced by CTI during the design and
development of the SCAIH rocket motor. The main engineering challenges involve the mixing of the
gas-generator gases and liquid oxidizer, and their subsequent combustion, incorporating chemical
kinetics, turbulent combustion, and radiative heat transfer.
Thus, the primary goal of this research is using and developing numerical methods, modelling,
and simulation tools to enhance understanding of the injection, mixing and combustion processes
within a hybrid rocket motor. Secondary objectives are to refine the design of the fuel/oxidizer
injection system and to define a better picture of the flame anchoring mechanism.
The numerical combustion modelling will be a piecewise construction of many facets, beginning
with a two-dimensional (2-D) axisymmetric simplified flow model. The introductory chapter outlines rocket technology and the present state of hybrid rocket research including the SCAIH motor. Chapter 2 introduces the axisymmetric flow model including the geometry of the combustion
chamber and injector, computational grid, and Navier-Stokes equations. Chapter 3 details turbulent
combustion which includes turbulence modelling and chemical kinetics. Chapter 4 considers the
effects of radiative heat transfer. The initial state of the injected gas-generator gasses and oxidizer
and critical point considerations will be discussed in Chapter 5. The overall solution methodthe
finite volume methodis detailed in Chapter 6.
The results presented in Chapter 7 represent a two-phase reactive flow model to study the gasgenerator/LOX combustion at near- and trans-critical regimes. Chapter 8 concludes with an overview
of the thesis results including areas where future work may proceed.

Chapter 2

Axisymmetric Flow Model


Creation of a model for the fuel injector and combustion chamber of a SCAIH rocket motor first
begins with the dimensions of the components including: a shear coaxial injector, combustion
chamber and converging diverging nozzle. Next a computational grid is formed to represent the
rocket motor and finally a mathematical model is needed and formulated to represent the flow
within the grid.

2.1

Shear Coaxial Injector

In general, a shear coaxial rocket injector injects a liquid oxidizer as a centralized jet and a fuel in
the annular portion surrounding the central tube. Traditionally this is the configuration that rocket
motors use; the oxidizer is injected centrally with the fuel around it. Alternate rocket injectors
include swirl and impingement arrangements. In a swirl injector the oxidizer is injected tangentially
through slots upstream and exits the injector post in a hollow cone shape. Swirl injectors tend to
enhance flow breakup and atomization and mixing [35]. Impinging injectors use guided collision
to mix and atomize the fuel and oxidizer streams and have had success in large rocket vehicles like
Atlas and Saturn. Heavy lift vehicles such as these often use hundreds of single injector elements
in one motor; the SSME has 600 [36]. Flow breakup and atomization are important aspects of
injection systems, especially when the propellants are injected in a state at or near to their critical
point. Chapter 5 discusses thermodynamic and transport properties for multi-phase treatment.
Figure 2.1 shows a photograph of a typical rocket injector used by CTI and also the schematic and
dimensions of the SCAIH rocket injector. It is the injector and the near injector region containing
the flame stabilization and anchoring mechanism that is a key portion of the modelling of rocket
13

CHAPTER 2. AXISYMMETRIC FLOW MODEL

(a) Typical CTI injector

14

(b) SCAIH injector

Shear coaxial rocket injector photo (a) and schematic by CTI (b). On the schematic flow is from
right to left with LOX flowing in the center and gas-generator gas flowing in the annular ring.
Dimensions in inches.

combustion and is an area of focus for the SCAIH.

2.2

Flame Anchoring Mechanism

After the transitory ignition period the flame will achieve a steady- or quasi-state and remain there
due to the flame anchoring mechanism. Once steady state conditions have been reached the flame
will anchor somewhere near the vicinity of the injector face. Under certain conditions the anchoring
of the flame may be on or too close to the injector causing ablation of the injector components.
Ideally the flame should anchor at some finite distance away from the injector to avoid excessive
heating.
As mentioned in Section 1.2, the current SCAIH design is experiencing high thermal loading in the
near injector region. Experimental testing completed in-house by CTI has indicated temperature in
this region to be in excess of 3000 K. This causes their injectors to burn up before the end of a full
test burn. Excessive ablation that damages the injector before the end of the burn is unacceptable
and must still be overcome for the SCAIH rocket motor to be realized.
The flame anchoring mechanism visualized in Figure 2.2 shows the LOX/CH4 flame being anchored
between two counter rotating wake recirculation zones [37]. Within a recirculation zone such as
created by a shear coaxial injector, there is a stagnation point around which the flame generally
anchors. Anchoring of the flame slightly downstream of the injector is counter to the case of a
LOX/H2 flame which anchors much closer to the injector post because of the high diffusivity of and
low density of hydrogen [38].
To date, much of the research of this type has focused on LOX/H2 combustion and there is an

CHAPTER 2. AXISYMMETRIC FLOW MODEL

15

Schematic diagram of a flame anchoring mechanism for a coaxial injector for methane/LOX. The
injector centerline is on along the abscissa [37].

absence of LOX/CH4 data [28]. Additionally the high pressures of the SCAIH rocket combustion
are often not considered in studies of this type. The work of Zong and Yang, [37], and Lux and
Haidn, [39], have begun to investigate LOX/CH4 combustion at and above super-critical pressures.
Zong and Yang found that the flame anchored in the wake and recirculation region behind the LOX
post, which differs from the LOX/H2 flame that would anchor very close to the post. This proximity
of the anchoring mechanism to the injector is cause for additional thermal loading. Using OH
emission images, Lux and Haidn reported that the LOX/CH4 flame was similar to the LOX/H2
flame at similar operating conditions. Using the Abel transform to capture OH emission, the flame
is seen in Figure 2.3. The authors concluded that the flame anchors in the wake of the LOX post tip
at all operating conditions once the oxygen is provided in a liquid or subcooled state. Super-critical
flow conditions including mixing and breakup are introduced and discussed briefly in Chapter 5.

2.3

Combustion Chamber

The combustion chamber that has been developed for testing by CTI is a sub-scale combustion
chamber with one full scale injector element along the centerline. The layout and dimensions
are shown in Figure 2.4. Sub-scale combustion chambers are advantageous because there are
cost savings associated with manufacturing of development hardware, fewer iterations of full-size
hardware, and reduced testing costs associated with fuel and testing personnel [40]. CTIs rocket
motor has completed many tests with LOX and gas-generator fuel with burn times up to 18 seconds.
Operationally it has achieved reliable ignition and shutdown, stable combustion, and operating
pressures around 700 psi (47.6 atm) [17].

CHAPTER 2. AXISYMMETRIC FLOW MODEL

The near injector region for LOX/CH4 combustion under supercritical conditions. The image shows
averaged OH emission from the application of a deconvolution using the Abel transform. Additionally, this technique can show the magnitude of the reacting shear layer between the liquid oxygen
and gaseous methane. From Lux and Haidn [39].

Sub-scale combustion chamber schematic used for testing a single shear coaxial injector. Gasgenerator gases and LOX are injected on the left-hand side. Dimensions in inches.

16

CHAPTER 2. AXISYMMETRIC FLOW MODEL

2.4

17

Grid Generation

A grid is required to divide the computational domain into cells that can be used to provide a
discrete representation of the injector flow field of interest. Once determined, information is passed
between the cells and the flow field can evolve as the solution is updated by an appropriate solution
method. For the present case of a cylindrical rocket motor with a concentric injector along the
centerline, a structured, axisymmetric formulation of 2-D, quadrilateral cells is sufficient. More
complicated, irregular geometries can require and benefit from unstructured meshes [41].
A multi-block body-fitted structured grid topology has been developed to the specifications in Figure 2.4. The grid represents a 2-D axisymmetric coaxial injector and combustion chamber with a
nozzle (Figure 2.5a). In order to provide a robust foundation for simulations current and future, a
high level of flexibility has been built-in. Grid dimensions, block size, and cell size can all be adjusted. The multi-block design with cell and block clustering and support for block-based adaptive
mesh refinement (AMR) allows for parallelization which improves computational performance.
The upstream feed section allows for flow to develop before entering the combustion chamber. Nozzle implementation is not crucial at this point; however, its inclusion in the future could represent
a more complete overall description of the rocket combustion chamber. Both of these aspects of the
grid can optionally be turned off. The overall grid block structure including a nozzle and upstream
feed sections is in Figure 2.5b. A detailed view of the injector region showing the injector post and
the upstream feed lines is in Figure 2.5.

CHAPTER 2. AXISYMMETRIC FLOW MODEL

18

(a) SCAIH rocket motor outline. Axis units are meters.

(b) Overall computational grid showing fuel and oxidizer injector blocks (bottom left corner) and nozzle.
Axisymmetric x-axis. This example has 72 blocks with their outlines shown.

(c) Injector and combustion chamber; divisions are (d) Zoom view showing the wall boundaries (block
computational blocks.
boundaries have been removed).

Computational domain with a description of the near-injector region. Dimensions in meters.

CHAPTER 2. AXISYMMETRIC FLOW MODEL

2.5

19

The Navier-Stokes Equations

A fluid is defined as a substance with no fixed shape which is very responsive to surface forces
(i.e. pressure and stress), as well as other body forces. From an engineering perspective this is
almost always a liquid or gas or a combination therein. In the study of aerodynamics the fluid is
generally a gas; air externally flowing over a body. Internal flows are those where the fluid remains
generally confined and flows within a fixed boundary such as in internal combustion engines, gas
turbines, and rocket motors and nozzles. Modelling these flows via computer simulation has been
improving continually since the 1960s and is fundamentally based on equations developed by two
scientists in the 19th century: Claude-Louis Navier and George Gabriel Stokes [41]. The NavierStokes equations are a set of partial differential equations that govern flow of a viscous fluid. Using
vector notation in two-dimensions there are three conservation equations: mass, momentum and
energy.
The continuity equation represents conservation of mass and is given by

+ (~u) = 0

(2.1)

where is fluid density, ~u is fluid velocity and t is time.


The conservation of momentum vector representation takes the form
(~u)
t

~
~I ) =
~ + ~f
+ (~u~u + p~

(2.2)

~I is the identity tensor,


~
~ is the fluid stress tensor, and ~f is a body
where p is fluid pressure, ~
force acting on the mixture. This momentum equation embodies Newtons 2nd law, F = ma, by
recognizing that the time rate of change of momentum of a fluid parcel is equal to the net force
acting on the parcel.
And similarly, conservation of energy is
(e)
t

 

p
~
+ ~u e +
= (~u
~ ) ~q + ~u ~f

(2.3)

where ~q is the heat flux vector and e = c v (T ) is the total internal energy and c v is the specific
heat coefficient based on a constant volume. It is of note that formally only the conservation
of momentum (2.2) constitutes the Navier-Stokes equations, not mass (2.1) and energy (2.3).
This work will continue to refer to the above set in the colloquial manner as the Navier-Stokes
equations.

CHAPTER 2. AXISYMMETRIC FLOW MODEL

20

The heat flux vector, ~q, is obtained from Fouriers law:


~q = T

(2.4)

where is the thermal conductivity of the fluid. Assuming a Newtonian fluid where the viscous
~
stresses are proportional to the rates of deformation, the molecular stress tensor,
~ , represents the
molecular rate of the transport of momentum and is:


1~
~
~
~ ~

~ = 2 S
I ~u
3

(2.5)

~
~ is the mean strain rate tensor.
where is the molecular viscosity. S
An equation of state is required to determine the pressure. For thermally perfect gases the ideal gas
law is:
p = RT

(2.6)

where R is the universal gas constant and T the temperature. More accurate modelling of the
equation of state to account for fluids at high pressure and low temperature will be discussed in
Section 5.1.2.
Combined, these equations completely describe the behaviour of a compressible, thermally perfect,
non-reacting flow in multi-dimensional domains. Modification to this equation set will follow in
subsequent chapters to address the additional complexities introduced to the SCAIH motor flow
field by combustion, turbulence, chemical kinetics and radiation processes.

Chapter 3

Turbulent Combustion
Flames in combustion processes can be classified as either premixed or non-premixed. Premixed
flames are those where the fuel and oxidizer components have been mixed on a molecular level
before burning. Those which have not been mixed beforehand are referred to as non-premixed
flames and have a much wider breadth of application including gas turbine and diesel engines, oil,
gas, pulverized coal-fired boilers and furnaces, chemical lasers, as well as liquid and hybrid rocket
motors.

Oxidizer

Flame

Fuel

Temperature
Oxidizer

Fuel

Reaction Rate

General structure of a laminar diffusion flame [42].

Figure 3.1 shows the structure of a non-premixed flame where the oxidizer and the fuel diffuse
together in a reaction zone that defines the flame. The rate of reaction is influenced by the rate
of mixing and also the rate of molecular diffusion. This interaction is further complicated by flow
turbulence.
21

CHAPTER 3. TURBULENT COMBUSTION

22

Turbulence
Turbulence is described as an unsteady variation in flow field properties and can be identified in
flows by the irregular formation and structure of eddies or vortices. Most real-world engineering
flows exhibit turbulent characteristics. From extremely small scale flows such as narrowed arteries
due to cholesterol build-up, to mega-scale flows like galaxy formation, turbulence plays a role. In
the field of rocketry, turbulence affects the external flow over the vehicle as well as the internal
flow of the fuel and oxidizer, their mixing, combustion, and exhaust.
The Navier-Stokes equations as discussed in Section 2.5 provide a full flow description but require
accurate resolution of all turbulent scales which today is not computationally possible for most
practical flows. This is because turbulence involves variation in time and length scales on a much
wider range than laminar flows. The difference between the large and small scales of a typical
turbulent flow can be orders of magnitude. To accurately capture the variation in scales a numerical
simulation needs a very fine grid making the problem computationally expensive. Solving the
Navier-Stokes equations for complete resolution of the turbulent scales is termed direct numerical
simulation (DNS) and is too computationally expensive, even with modern supercomputers, for
practical application [43, 44].
Turbulence has been called the last major unsolved problem of classical physics. This is because
an analytical solution for the description of turbulent flow in three-dimensions has not been found.
Many models however have been developed that capture aspects of turbulences irregularity. This
field is called turbulence modelling and itself is a complex yet necessary aspect to modelling practical engineering flows, such as the injector flow fields of hybrid rocket motors.

3.1

Turbulence Modelling

Turbulence modelling seeks to accurately represent the different length scales involved in turbulence while avoiding the computational expense of resolving them completely through DNS. The
length scales in question exist within a continuum; that is there are not discrete groupings based
on eddy size that can separate large- from small-scale turbulence. The largest scales involved are
on an order of magnitude similar to the flow width; or of the size of the object around which the
turbulence is occurring. Alternately the small scales can be many orders of magnitude smaller.
These large scale eddies are energy bearing and largely responsible for the enhanced mixing behaviour and stresses of turbulent flows. The turbulent kinetic energy contained in this vorticity
dominated motion cascades down to the smallest scale and ultimately dissipates to heat through

CHAPTER 3. TURBULENT COMBUSTION

23

the action of molecular viscosity. So decaying turbulence is a conversion of kinetic energy present
in the large scale eddies to heat at the molecular level.
The smallest scales of turbulence are still larger than the molecular length scale or the mean free
path of the flow and are referred to as the Kolmogorov length scales [45]. Defined as

1
4

"

(3.1)

where is the kinematic viscosity and " is the rate at which the large eddies supply energy, the
Kolmogorov scale for a typical turbulent flow could be 106 times smaller than the large scales.
Thus, the rate at which energy is supplied via large eddies must be equivalent to the amount dissipated through heat [46, 47]. Recently, large eddy simulation (LES) methods have shown progress
in reducing overall computation cost by modelling the large-scale eddies that carry most of the turbulent kinetic energy; however, these methods are still relatively new and undeveloped and more
expensive than Reynolds and Favre-averaging methods that will be discussed below and used in
this thesis to represent the influence of the turbulence [48, 49].
To accurately capture turbulence and its effects on the flow field a continuum of scales need to be
resolved. The computational grid is limited by the smallest scale, and must be at least as fine as the
Kolmogorov length. In order to avoid the expensive resources of DNS a statistical method known
as Reynolds-averaging is employed.

3.2

Reynolds-Averaging Procedure

Turbulence deals with random fluctuations of flow properties and so it is appropriate to use a
statistical approach to represent the whole. Osborne Reynolds was a pioneer of turbulence and in
1895 introduced averaging concepts based on mean and fluctuating parts [50]. There are three
different forms of averaging: time, spatial, and ensemble. Time-averaging is the most commonly
used and will be briefly discussed here; for full details refer to the textbook by Wilcox [51].
Time-averaging assumes the flow does not vary with time and therefore can be averaged using
different points in time. Consider a flow quantity, , expressed as (~x , t). The averaging procedure
x , t), and fluctuating, 0 (~x , t), parts:
will express the instantaneous quantity as a sum of mean, (~

x , t) + 0 (~x , t).
(~x , t) = (~

(3.2)

CHAPTER 3. TURBULENT COMBUSTION

24

The time-average of is defined by


T (~x ) = lim

t+T

(~x , t)d t

(3.3)

where T (~x ) is the mean value of our flow variable and T is the total time interval to average.
Time-averaging the continuity Equation (2.1) is

+ (~u) = 0

(3.4)

where
+ ~u0 )
+ 0 )(~u
~u = (
+ 0 ~u
+ ~
~u
u0 + 0 ~u0
=
+ 0 ~u0
~u
=

(3.5)

For an incompressible constant density scenario Equation (3.4) is zero and matches the continuity
Equation (2.1). Application of this procedure to the Navier-Stokes equations (Section 2.5) yields
what are called the Reynolds-Averaged Navier-Stokes equations (RANS).
For compressible flow the RANS equations introduce new terms that can not be solved a priori and
need additional modelling. We have already seen one of these terms, 0 ~u0 , in Equation (3.5), and
Reynolds-averaging the momentum Equation (2.2) creates more; specifically a triple correlation
involving the Reynolds stress tensor. A variant on this style of averaging called Favre-averaging will
herein be discussed and used.

3.3

Favre-Averaging Procedure

The operating regimes of rocket motors involve very high pressures and temperatures and so compressibility is an important factor. Due to the temperature, density and pressure fluctuations involved, this thesis will focus on the Favre-averaged Navier-Stokes equations (FANS). The Favreaveraging procedure is also known as mass-averaging and uses a density-weighted procedure [52]
developed by Favre in 1965.
e and fluctuating, 00 , component with
Again, the Favre-averaged quantities involve a mean, ,
is used to represent Reynolds-averaging and an overslightly different notation. An overbar ()
e to represent Favre-averaging. A Favre-averaged flow quantity, is given by
tilde ()

CHAPTER 3. TURBULENT COMBUSTION

25

f
~ x , t) = (~
~ x , t) +
~ 00 (~x , t)
(~

(3.6)

e, is defined by
And a mass-averaged velocity vector, ~u
e=
~u

lim

t+T

(~x , t)~u(~x , t)dt

(3.7)

is the Reynolds-averaged density and


where
e + ~u00
~u = ~u

(3.8)

is the instantaneous velocity decomposed into a mass-averaged and fluctuating part respectively.
Multiplying Equation (3.8) by density and completing a time-average yields
e = ~u = ~
u + 0 ~u0
~u

(3.9)

To arrive at the Favre-averaged continuity equation, substitute Equation (3.9) into Equation (3.5)
to get


e =0
~u
+

(3.10)

Here we see that Equation (3.10) has the same form as the laminar version (Equation 2.1) and is a
convenient simplification.

3.3.1

The Favre-Averaged Navier-Stokes Equations

Applying the Favre-averaging procedure to the Navier-Stokes equations in Section 2.5 leads to the
FANS set suitable for a compressible, thermally perfect flow field. The time-averaged forms of the
conservation of mass, momentum, and energy, respectively, can then be expressed using vector
notation as


e =0
~u
+

t





~
e +
e~u
e + P~
~
~ + ~f
~I =
~u
~u

~ +
  t 
h
 i

P



~
e e
e + Dk k ~q + ~q t + ~u
e ~f ,
~
~ ~u
e +
~u
e
e+
=
~ +

(3.11)
(3.12)
(3.13)

CHAPTER 3. TURBULENT COMBUSTION

26

where:
Reynolds-averaged density

e Favre-averaged velocity
~u
~I identity tensor
~
~
~ turbulent Reynolds stress tensor

P mean static pressure


~ viscous Reynolds stress tensor

~
e
e Favre-averaged total energy

Dk

k specific turbulent kinetic energy


~q t

turbulent heat flux vector

turbulent energy diffusion coefficient

~q molecular heat flux vector


~f body force per unit volume

The equations above provide a description of the time-averaged flow field; however, the Favreaveraging process has introduced a number of additional terms to the equations related to the
unspecified turbulent fluctuations which must be handled for closure of the system.

3.4

k Turbulence Model

~
~ in terms of known time-averaged
Forming a relationship for the turbulent Reynolds stress tensor, ,
flow quantities is the basis for a range of by now almost classical turbulence models. Using an
analogy to the way that momentum transfer by molecular motion can be described using molecular
viscosity, Boussinesq postulated that turbulent eddies can be described by an eddy viscosity [53].
The result is known as the Boussinesq approximation and relates the Reynolds stress tensor to the
mean flow strain rate as follows:
~
~ = ~u00 ~u00

1~
2~
~
e) ~
~ ~

= 2 t (S
I ~u
I k
3
3

(3.14)

where t is a so-called turbulent eddy viscosity.


The simplest turbulence models are called algebraic models which use a relationship between eddy
viscosity, t , and the length scales of the flow, l mi x , to compute the Reynolds stress tensor. More
recently, more complicated models based on the turbulent kinetic energy are used which can take
advantage of modern computing resources. These are termed either one- or two-equation models
named for the number of additional transport equations needed. The two-equation k model by
Wilcox [51] is used in this work and will be detailed here.
The specific turbulent kinetic energy, k, is a new quantity defined as the kinetic energy per unit

CHAPTER 3. TURBULENT COMBUSTION

27

volume of the turbulent fluctuations such that


=
k

1
2

~u00 ~u00 .

(3.15)

The transport equation for k is derived by multiplying the instantaneous momentum Equation (2.2)
by ~u00 and applying the time-averaging procedure to the result. Following some algebraic manipulations, one can arrive at the following transport equation describing the time evolution of the total
turbulent kinetic energy given by

~




ek =
e + + t k k
~ ~u
+
~u

where the turbulent eddy viscosity is t =

k
.

(3.16)

The diffusion of turbulent kinetic energy, Dk , given

in Equation (3.13) can now be defined as Dk = + t where is a closure coefficient shown


below.
The specific dissipation rate, , can be thought of as the ratio of turbulent dissipation to turbulent
kinetic energy and is the second transport equation of the k turbulence model. The postulated
form for the transport equation describing the time evolution of the specific dissipation rate is

~




e =
e + + t
~ ~u
~u

2.

+
t
k

(3.17)

This model has a number of closure coefficients given by


=

13
25

, = f , = f , = =

with
=

9
125

, =

f =

and

1 + 70
1 + 80

9
100

1
2

1
1 + 680k2
=

1 + 400k2

k 0 ,
k > 0 ,



(

~
~
~
~
~
~
1

)
:
S


~
~ .
=
, k = 3 k

3
( )

~
~
~ are the vorticity and strain rate tensors, respectively. In tensor notation, these
~ and S
The tensors
are defined as

CHAPTER 3. TURBULENT COMBUSTION

i j =

1
2

Ui
xj

28

Uj
xi

, Si j =

Ui
xj

Uj

xi

(3.18)

The near wall treatment of turbulence is handled by the standard wall function which is used to
directly specify the value of within a set distance from the wall [51].

3.5

Conservation of Species

For simplicity of understanding, the flow up to this point has been treated as composed of a single
non-reactive species. Since combustion involves the chemical transformation of two or more reactants into products, a multiple species reactive mixture must be considered. Specific composition
of the reactants and chemical processes involved in this work will be addressed in Sections (3.6)
and (5.2). Species conservation requires that an additional transport equation be added.
The time evolution of N total species mass fraction, Yn , (n = 1, 2, . . . , N ) is


e = J~n + J~t +
Yen +
Yen ~u

n
t

(3.19)

n is the mean rate of change of species mass fraction (discussed further in Section 3.6), J~n
where
is the molecular diffusive flux for species n and J~t n is the turbulent diffusive flux for species n.
Chemical reactions involving n = 1, 2, . . . , N species also lead to a new definition for the equation of
state where the total mixture pressure, P, is taken to be the sum of the partial pressures according
to Daltons law for a perfect mixture
P=

N
X

e
nRn T
Y

(3.20)

n=1

e is the Favre-averaged mixture temperature. The


where R n is the gas constant for species n and T
individual species enthalpy, hn , also has to be considered in the definition of the Favre-averaged
specific total energy for the mixture. This quantity is defined as
e
e=

e~u
e
~u
2

N
X
n=1

Yen hn

+k

(3.21)

where k is the specific turbulent kinetic energy that needs a separate transport equation (Section 3.4).
Molecular heat transfer is represented similarly to that given in Equation (2.4) with the modification

CHAPTER 3. TURBULENT COMBUSTION

29

of a balance of individual species enthalpy and diffusive flux so that

~q = T

N
X

hn J~n

(3.22)

n=1

where
J~n =

Sc

Yn

(3.23)

and is the mixture thermal conductivity. The Schmidt number, Sc, is a dimensionless parameter
representing the ratio of momentum diffusivity to mass diffusivity:
Sc =

(3.24)

Dn

where Dn is the species diffusion coefficient.


Additionally, there is a turbulent contribution to the heat transfer, ~q t , which can be modelled using
a generalization of Reynolds analogy between momentum and heat transfer so that

~q t = t T

N
X

hn J~t n

(3.25)

n=1

where
J~t n =

t
Sc t

Yn

(3.26)

and t is the mixture turbulent thermal conductivity. Here Sc t is the turbulent Schmidt number, set
to a constant 1 and t =

t cp
P rt

, where P r t is the turbulent Prandtl number which is set to a constant

of 0.9.

3.5.1

Thermodynamic Closure

Thermodynamic relationships and transport coefficients are required for closure of the system of
equations governing the reactive mixture described above. The compressible, reactive, gaseous
mixture is assumed to be thermally perfect, i.e., a mixture of gases in which the specific heats are
only functions of temperature [54].
Thermodynamic and molecular transport properties of each gaseous species are prescribed using
the empirical database compiled by Gordon and McBride [55, 56], which provides curve fits for the
species enthalpy, hn ; specific heat, c p n ; entropy; viscosity, n ; and thermal conductivity, n , all as
functions of temperature, T . For example, the enthalpy and viscosity for a particular species are

CHAPTER 3. TURBULENT COMBUSTION

hydrogen
GRI-Mech 3.0
jetfuel

30

Mechanism

#Species

#Reactions

H2 oxidation
CH4 oxidation
(C10 H22 )n oxidation

8
53
193

40
325
1085

Detailed chemistry mechanisms and the increasing complexity associated with intermediate reaction steps and species [13, 58].

given by
hn = R n T a1,n T 2 + a2,n T 1 ln T + a3,n +
a6,n
4

T3 +

a7,n
5

a4,n
2

T+

a5,n
3

T 4 + b1 T 1 + hof ,

ln n = An ln T +

(3.27)

Bn
T

Cn
T2

T 2+

+ Dn ,

(3.28)

where ak,n , An , Bn , Cn , and Dn are the coefficients for the curve fits and hof is the heat of formation.
n

The Gordon-McBride data set contains curve fits for over 2000 substances, including 50 reference
elements.

3.6

Chemical Kinetics

Chemical kinetics is the study of chemical processes and their rates to better understand how chemical reactions evolve and intermediates and products are formed. Combustion processes involve a
large number of intermediate species and reaction steps that all must be accounted for to accurately
model flames and combustion emissions. Examples of the sizes of chemical kinetic mechanisms
involved in detailed chemical kinetic descriptions of various fuels are given in Table 3.1. The GRIMech 3.0 mechanism [57] is one of the most widely used for methane combustion but has a high
computational cost limiting it to simple flames and small domains without being able to incorporate
other elements that may be used in fuels, such as aluminum or chlorine.
Turbulent reacting flows increase the complexity due to the different scales involved. Figure 3.2
shows the orders of magnitude of chemical and physical time scales. The large differences in
scales here show the ranges that must be treated to properly resolve the flow on a computational
grid. Comparing these scales can lead to a general understanding of the structure of turbulent
flames, which are used to derive models for turbulent combustion. To start, turbulent flow can
be characterized by a turbulent Reynolds number, Re t , which is the ratio of turbulent transport to

CHAPTER 3. TURBULENT COMBUSTION

31

viscous forces:

Physical time scales

Chemical time scales

Slow time-scales
e.g. NO formation

1s

102 s

Intermediate
time scales

Time scales of flow,


transport, turbulence

104 s

106 s
Fast time scales
steady state
partial equilibrium

108 s

Time scales in a turbulent combusting flow [59].

Re t =

u0 l t

(3.29)

where u0 is velocity rms, l t is the turbulent integral length scale and is the kinematic viscosity of
the flow.
The Damkhler number relates the ratio of turbulent time scales, t , to chemical time scales, c ,

Da =

t
c

(3.30)

The Damkhler number is a measure of the influence of flow mixing. As the Damkhler number
increases, Da  1, the chemical time is much shorter than the turbulent time and corresponds to a
thin reaction zone, or a flamelet. Here it can be assumed that chemical reactions occur infinitely fast;
as soon as the reactants mix together products are created. The lower limit, Da 0, corresponds to
a frozen chemistry situation where the reactions are very slow compared to the fluid transport. This
is also known as a perfectly or well-stirred reactor. When the Damkhler number is of order unity
there is the closest balance between turbulent and chemistry interactions and the mixed-is-burnt
assumption can not be used; i.e. finite-rate chemistry needs to be incorporated.
The Damkhler number and Reynolds numbers can also be related to each other such that

CHAPTER 3. TURBULENT COMBUSTION

32

p
Da = Re t Da

(3.31)

for constant Damkhler numbers, Da , where is a proportionality constant [42]. Most practical
combustion processes correspond to medium or high Damkhler numbers and that is the situation
here for rocket combustion and injector flow fields of interest.

3.6.1

Eddy Dissipation Model

The eddy dissipation model (EDM) for chemical kinetics builds on the eddy break-up model [60],
which is intended for turbulent premixed combustion. This model assumes a mixed-is-burnt relationship between the reactants and products related to high Damkhler number flows, meaning
that chemical reactions occur infinitely fast and once the fuel and oxidizer mix the reactants proceed fully to completion. Thus, the reactions are intimately dependent on the rate of turbulent
mixing of the fuel and oxidizer eddies, or, on the rate of dissipation of the eddies.
The EDM proposes that for a single step, irreversible chemical reaction between a fuel and oxidizer:
F + s O (1 + s) O

Where s is the mass stoichiometric coefficient for the products, P. The relationship between the
fuel and oxidizer quantities is fluctuating and is affected by their local concentrations. In a limiting
F , according to the minimum of
manner, the EDM determines the burning rate of the fuel,
F =

"
k

YeF ,

YeO
s

YeP

(1 + s)

(3.32)

where " and k are turbulent dissipation and kinetic energy respectively, s is the stoichiometric
oxidizer requirement, and are adjustable parameters. In this work = 4.0 and = 0.09. YeF ,
YeO , and YeP are the mixture concentrations for fuel, oxidizer and products respectively. The reaction
rate is limited by the deficient species and the turbulence mixing time.
In relation to the eddy break-up model, the EDM relates the dissipation of eddies to the mean
concentration of intermittent quantities instead of requiring knowledge of the concentration fluctuations which can be difficult to determine [61]. The EDM outlined here is generally used for a
simple one-step irreversible chemical reaction such as methane combustion with air:

CH4 + 2 O2 2 H2 O + CO2

CHAPTER 3. TURBULENT COMBUSTION

33

where N2 is inert. The EDM model will not provide the detail that a more representative mechanism
such as the GRI-Mech 3.0 would, however, due to its ease of implementation and low computational
cost it is often found in many commercial CFD software tools.
The solid fuel grain burned in the staged combustion aft-injected hybrid rocket is a mixture of
fuel, oxidizer and binding components. As such, its combustion involves more species than just
methane and oxygen, and a more complex reaction mechanism. A simple extension to multi-species
chemistry using a fixed composition is developed in Section 5.2.1 for treating the combustion of
the gas generator propellant gases via an EDM approach. As is also discussed in Chapter 5, the
combustion of the solid grain to produce the propellant gases will be modelled via a simple chemical
equilibrium analysis.

Chapter 4

Radiative Heat Transfer


Radiative heat transfer is an energy transfer process whereby all matter with a temperature greater
than absolute zero emit and absorb electromagnetic waves through a medium. The emission is
characterized by wavelength and intensity and for the purpose of engineering heat transfer only
ultraviolet, visible, and infrared wavelengths (107 m103 m) are of interest. Emission intensity is
strongly correlated to the temperature of the emitting body and can dominate other forms of heat
transfer at sufficiently high temperatures [62].
The heat transfer terms in Equation (3.13) are those for molecular heat flux, ~q (3.22), and turbulent
heat flux, q~t (3.25). These equations are based on Fouriers Law for heat conduction and species
molecular diffusion and the heat flux is taken to be linearly proportional to temperature
~q = T .

(4.1)

Radiative heat transfer adds to this heat flux and scales more quickly with temperature and is
proportional to the fourth (or higher) power
~qrad T 4

(4.2)

making it an influential component of high-temperature combustion devices including rocket motors. In combustion devices heat transfer to the chamber surroundings can be a limiting factor in
terms of durability. In a rocket combustion chamber radiative heat transfer represents a transfer of
heat from high-temperature flames to relatively cool surroundings, thus impacting chamber design.
That radiation plays an important role in heat transfer is well known and has been reviewed by
Viskanta and Meng [63] and de Ris [64]. Prediction of flame structure, species concentration

34

CHAPTER 4. RADIATIVE HEAT TRANSFER

35
z
s

Cylindrical coordinate system used for radiative heat transfer.

and soot formation all depend on accurate radiation modelling. More recently Byun has shown that
radiative heat flux from soot has a large influence on wall temperatures in a LOX/kerosene rocket
motor [65]. For modelling the radiative transfer in a participating medium (one that involves
absorption, emission and scattering), the solution to the radiation transfer equation is required.

4.1

Radiation Transfer Equation

The radiation transfer equation (RTE) represents a conservation of radiative energy applied to
a monochromatic beam of light where it gains energy through emission, loses energy through
absorption and redistributes energy through scattering [62]. In a direction defined by the unit
vector
s, the RTE is
1 I
c t

+
s I = Ib I s I +

s
4

I (
s0 ) (
s0 ,
s) d0

(4.3)

where c is the speed of light in a vacuum, is the wavenumber, I is the spectral intensity, Ib is the
blackbody radiative intensity, is the absorption coefficient, s is the scattering coefficient, and
is the scattering phase function. The quantities here all vary with location in space, wavenumber,
and time. Additionally the intensity and phase function also depend on direction.
The first term of Equation (4.3) can be neglected if it is assumed that the time and length scales
involved are small compared to the speed of light so that a steady state is reached. In radial and
axisymmetric dimensions with angular coordinates as shown in Figure 4.1 the RTE becomes
(r I )
r

1 (I )
r

I
z

= I + Ib +

s
4

Z
4

I (
s0 ) (
s0 ,
s) d0

(4.4)

CHAPTER 4. RADIATIVE HEAT TRANSFER

36

with the polar and the azimuthal angle, and , and are the direction cosines. An extinction
coefficient, , representing total attenuation by absorption, , and scattering, s , is
= + s .

(4.5)

The divergence of radiative heat flux is the quantity of interest, so rearranging the RTE and integrating over all solid angles gives
~q =

4Ib

!
I d

= 4Ib G

(4.6)

I d .

(4.7)

where the total incident radiation, G , is


G =

Here ~q is the divergence of the radiative heat flux at a specific wavenumber, . For the total
radiative heat flux, ~qrad , Equation (4.6) must be integrated over the entire spectrum. This is
represented as a source term in the conservation of energy Equation (3.13) for the reactive gaseous
mixture.
It is worth noting that Equation (4.6) no longer contains the scattering coefficient because the net
loss of radiative energy is only dependent on the emission minus the absorption and scattering
only alters the direction of a photon. See the text by Modest [62] for more detail regarding the
derivation of the radiation transfer equation.
Equation (4.3) is a multi-dimensional hyperbolic integro-differential equation with variations in
space, direction, wavelength and intensity. As such, analytical exact solutions only exist for simple
cases and numerical approaches are required for practical combustion applications.
Simple solutions to the RTE involve making approximations about the overall behaviour of the
medium. The optically thin and optically thick cases assume the medium to have no attenuating
effect or behave like a black body, respectively. Physically the smallest hydrodynamic and chemical
scales can be considered optically thin while the largest turbulence scales behave more as optically thick. Other approximations not used here include the cold medium approximation, two-flux
approximation and moment methods [62].

CHAPTER 4. RADIATIVE HEAT TRANSFER

4.2

37

Optically Thin Approximation

The optically thin approximation is simple and easy to implement and recently has achieved reasonably accurate results in the study of turbulenceradiation interaction [66, 67] and light- to nonsooting flames [68,69]. As the radiation medium becomes very thin such that  1 where is the
optical depth there is no reduction in intensity so every point in the medium has the same incident
radiation and attenuation rate. This means ~qrad can be evaluated directly without solving the
RTE.
Beginning with the divergence of the heat flux vector, ~q , in Equation (4.6) and integrating over
the entire wavelength spectrum gives
~qrad = 4

Ib Ib, d

(4.8)

where Ib, is the spectral blackbody intensity evaluated at the domain boundary. For a gray media
( = = constant) this may be simplified to

4
~qrad = 4 T 4 T

(4.9)

where is the Stephan-Boltzmann constant and T is the ambient temperature.


For non-gray media the analysis includes the Plank-mean absorption coefficient, P ,

4
~qrad 4P T 4 T

(4.10)

where P can be represented using the narrow band approximation which states that the absorption
coefficient over an entire spectral range may be approximated by smoothed values averaged over
a narrow spectral range. This is derived from the fact that a gas absorption coefficient varies much
more widely across the spectrum than the blackbody intensity [62]. Summing over the j th band,
P is
P =

Nb 
X
I
j=1

b0
S j
T 4

where S j is the effective line strength for band j.


(4.11)
j

CHAPTER 4. RADIATIVE HEAT TRANSFER

4.3

38

Spectral Absorption

The spectral absorption coefficient varies wildly with wavenumber requiring a very high resolution
hundreds of thousands of wavenumbersand then the radiation transfer equation must be integrated over the entire spectrum. This is a tiresome process known as line-by-line calculation and
requires knowledge of every single spectral line. The HITEMP database which is a high temperature
version of the popular HITRAN database contains over 1 million spectral lines for each of CO2 and
H2 O alone [70]. Models have arisen to approximate and simplify this process and can be classified
as narrow band, wide band, and global models.

4.3.1

Absorption Models

The narrow band model comes about by replacing the actual absorption coefficient by values that
have been averaged and encompass only a narrow spectral range. By capturing the appropriate
range this method can be nearly as accurate as line-by-line calculations [71]. Wide band models
can be derived by integrating narrow band models across an entire band. Global models calculate
total radiative heat flux directly since this is often the only quantity of interest.

4.3.2

Statistical Narrow Band Correlated-k Model

Two narrow-band derivatives are the statistical narrow band (SNB) and correlated-k (CK) methods.
A hybrid approach of these two methods is used here and will be briefly described. The statistical
narrow band correlated-k (SNBCK) method splits the wavenumber spectrum into intervals, ,
where the Planck function has small variance and solves the RTE for an average intensity over each
interval [72].
The SNB portion assumes that the spectral lines are randomly distributed in terms of both strength
and position across a narrow band. The CK distribution defines a function to represent the narrow
interval, , in terms of k such that the absorption coefficient is between k and k+dk. f (k) is
computed from the absorption coefficient spectrum as a weighted sum of the number of points
where = k [62].
Beginning with the CK distribution function, f (k), and integrating

g(k) =

f (k0 )dk0

(4.12)

CHAPTER 4. RADIATIVE HEAT TRANSFER

39

where g(k) is the cumulative distribution function of the absorption coefficient representing the
probability that k. This can be thought of as a pseudo wavenumber between 0 and 1. The
integrated RTE (4.3) over each band becomes

dg =


k (g) Ib I dg

(4.13)

where k (g) is the inverse function of g(k) and the subscript refers to the specific interval for
which a quantity is defined. Using numerical quadrature, at each quadrature point, g i , the RTE is
I,i
s

= k (g i ) Ib I,i

(4.14)

where I,i is the band intensity at point i. The total intensities are obtained as follows:

I j =
I =

Nq
X
i=1
Nb
X

w i I j,i
(4.15)

I j

j=1

where Nq and Nb are the number of quadrature points and bands, respectively, w i are the quadrature weights associated with g i , and j is the narrow band index.
Here the hybrid approach varies from the CK method by the method of calculating the distribution
function, f (k). Using the Malkmus model, the analytical expression for f (k) is

f (k) = L 1 { }
=

1
2

3/2

1/2

(BS)


exp

B
4


2

S
k

k
S


(4.16)

where L 1 is the inverse Laplace transform, is the narrow-band averaged transmissivity, B is


the effective line half-width, and S the effective line strength. Derived analytically by Lacis and
Oinas [72] the cumulative distribution function is

CHAPTER 4. RADIATIVE HEAT TRANSFER

g(k) =

1
2


erf

40

p  1  a
p 
a
p b k + erf p + b k exp (B)
2
k
k

where erf (x) is the error function, a =

1p
BS
2

and b =

1
2

(4.17)

B/S.

For each quadrature point in Equation (4.14) the corresponding absorption coefficient is evaluated
for each point in the domain. Lui et al. have improved the efficiency of the base SNBCK method
whereby four Gauss-Legendre quadrature points yield sufficient accuracy for detailed flame calculations [73] and is the method implemented by Charest [74].

4.4

Soot Radiation

As will be seen in Chapter 5, the gas-generator gasses produced from the primary combustion stage
of the SCAIH rocket motor contain a large fraction of solid carbon particles. As much as a third of
the generator gas by mass is composed of solid carbon. The presence and combustion of this high
temperature carbon in the secondary stage will contribute to radiative heat loss with peak emission
at wavelengths in the infrared region of the spectrum [13].
The chemistry and physics of soot formation and agglomeration is complex and particle size has an
effect. For radiative analysis particles are assumed to be spherical and this yields accurate results
when considering a statistical averaging of many particles.
Electromagnetic radiation travelling through a medium and interacting with particles can be scattered by diffraction, reflection or refraction and can also be absorbed. Lord Rayleigh was the first
to propose the interaction of single spheres with electromagnetic waves in 1871 and gave simple
solutions for particles that were much smaller than the wavelength of radiation. Radiation scattering in this regime is known as Rayleigh Scattering. Extensions to larger particles with respect to the
wavelength was undertaken by Gustav Mie and led to what is known as Mie Theory. For particles
that are larger than Mie theory accommodates, geometric optics are employed. For practical combustion application, soot particles can be treated using Rayleigh scattering due to relatively small
diameters (< 50nm) and radiation wavelengths in the infrared region ( 3 m) [62].
The combination of radiative scattering and absorption is called extinction (Equation 4.5) which
occurs in a cloud of particles. Typically soot particles are very small and the smaller these particles
become scattering tends toward zero. Assuming then that extinction occurs only through absorption
and that the complex index of refraction does not vary across the spectrum, one can write the

CHAPTER 4. RADIATIVE HEAT TRANSFER

41

spectral absorption coefficient as


S = C

(4.18)

where C is a constant set to 5.5 [75]. This holds for a cloud of spherical particles of varying
size. Considering both optically thin and thick media it has been shown that S does not vary
significantly and so Felske and Tien [76] average the absorption coefficient as
S =

3.72C T
C2

(4.19)

where C2 = 1.4388 is the second Planck function constant. Additionally, for a mixture of gas and
soot there are combined radiative effects from the each of the components such that
mix = S + gas .

(4.20)

It is important to point out that Equation (4.19) only applies to small soot particles and that aggregates which exceed the Rayleigh scattering size will increase the extinction coefficient.

4.5

Solution to the Radiation Transfer Equation

The two common methods for solving the RTE are the discrete ordinates method (DOM) and the
finite volume method (FVM). The DOM is based on discretizing the direction of the radiative intensity and is a finite differencing of the directional dependence of the RTE. The ordinates are the
chosen direction with which weights are individually applied and the RTE is solved in these directions with integrals for each quantity (i.e. radiative source term, radiative heat flux, etc.) being
evaluated by numerical quadrature [62].
The FVM is a variation of the DOM by exactly integrating solid angle volumes and therefore uses
a finite volume approach for both angular and spatial discretization. It is advantageous because it
ensures conservation of radiative energy whereas the DOM suffers from radiative energy loss due
to the approximate nature of numerical quadrature. Both the DOM and FVM can be extended to
arbitrary levels of accuracy and order.
Other methods in use include the discrete transfer method (DTM) first developed by Lockwood
and Shah [77] and Monte Carlo methods which are statistically based and often very computationally expensive [78]. For sooting flames which most closely resemble the case of a rocket
gas-generator involving large quantities of unburned carbon, research has mainly used the DTM or
DOM methods [75, 79] whereas lightly to non-sooting flames have used the optically thin approxi-

CHAPTER 4. RADIATIVE HEAT TRANSFER

42

mation [69, 80].


The DOM was used in this work because it provides a reasonable balance between accuracy and
computational efficiency. Both the DOM solution method and the optically thin approximation
using the SNBCK model applied in this work were previously implemented by Charest et al. [74,81].
Further details of the solution method for the RTE used here are found in Section 6.4.

Chapter 5

Modelling of Gas-Generator Propellant


Gas Formation and Chemical Kinetics
and LOX Expansion Process
Determining the inflow boundary conditions is an important part of accurately modelling the subsequent combustion processes. Dealing with a highly irregular and impure fuel as well as a cryogenically stored oxidizer adds to the challenge of determining these conditions. The injectors designed
and manufactured by CTI are shear coaxial injectors with LOX traveling through the central tube
and generator-gasses traveling through the annual opening. Figure 2.1 shows the geometry of a
single injector. As previously stated, the injector LOX and generator gasses are injected at very high
pressures into a high temperature combustion chamber. Injection systems of this type are complicated by flash evaporation, very rapid mixing and a general lack of understanding of high pressure
combustion processes. This chapter will detail the operating points of the rocket motor including
pressure and critical point considerations, gas-generator composition and combustion, and liquid
oxygen expansion. Solid carbon particulate is a significant and integral part of the fuel gases and
its modelling is also discussed.

43

CHAPTER 5. GAS-GENERATOR PROPELLANT, CHEMICAL KINETICS,

5.1

5.1.1

AND

LOX EXPANSION

44

SCAIH Operating Points

Pressure

The operating pressures of the SCAIH rocket are on the low end compared to other rocket motors.
For example, the space shuttle main engines use turbopumps to inject hydrogen and oxygen at 50.9
MPa with pressure drop in the combustion chamber to 46.6 MPa. One recent hybrid rocket motor
study has a similar chamber pressure of 50 atm (5.07 MPa) [31]. Lower pressure motors also exist,
the RL-10 is a rocket engine used in the upper stages of the Delta and Saturn rockets and has a
chamber operating pressure of 2.7 MPa [12].
Cesaronis SCAIH rocket has expected pressures of 725 psi (5 MPa) for the solid fuel grain (gas
generator) combustion and the liquid oxygen feed line, and 580 psi (4 MPa) for the combustion
chamber. This data has been collected during preliminary test firing conducted at CTI. Figure 5.1
shows the pressure trace for a test firing indicating both inlet line pressure (for the LOX and fuel
gases) and chamber pressure (2nd stage combustion).

Pressure trace over time. Tank is the constant pressure maintained in the LOX tank. Line is the
LOX feed line to the combustion chamber. GG Pressure is the pressure from the gas generator
combustion. Post CC is the pressure in the main combustion chamber. Test conducted at CTI [9].

CHAPTER 5. GAS-GENERATOR PROPELLANT, CHEMICAL KINETICS,

AND

LOX EXPANSION

45

Pressure Variation

As can be seen in Figure 5.1, the pressure for combustion processes varies over time; this includes
the ignition process as pressure builds and the effects of reactant consumption where eventually
pressure will recede. Solid fuel grain consumption will also contribute to declining pressure as the
volume of the combustion chamber increases. To offset declining pressure and maintain somewhat
stable combustion, liquid rockets will use a pressurant such as helium or nitrogen to maintain pressure in the fuel tank or turbopumps to feed the combustion chamber. Maintaining stable combustion
in solid rockets is tricky and concerns fuel regression which has been discussed in Section 1.1.2.
It is important to note that the simulations in this thesis employ constant pressure boundary conditions. All test cases have inlet pressures, both fuel and LOX, of 5 MPa expanded to an outlet
pressure of 4 MPa. These pressures represent the expected or average operating pressures that the
SCAIH rocket has been designed for. The effects of pressure change or fluctuation have not been
considered herein.
The operating pressures of the proposed CTI hybrid rocket system are near the critical points of
many of the constituent species and are responsible for the flow being either sub-, trans-, or supercritical. A brief discussion of the associated consequences of this high pressure now follows.

5.1.2

The Critical Point

The operational regimes of rocket engines can typically span the critical point ranges of the species
involved. Substances near and above the critical point need to be handled differently to accommodate strange behavior. Table 5.1 lists critical point properties of the gases found in the gas-generator.
Near the critical point density and thermal conductivity exhibit large gradients. At the critical point
surface tension goes towards zero and the heat capacity goes towards infinity. Above the critical
point, at supercritical regimes, the surface tension vanishes and the atomization and mixing of fuel
happens through turbulent and diffusive forces. In this regime, fluids behave as dense gases and
increases in pressure cannot liquidate them. Conversely, below the critical point the mixing of the
fuel is driven by surface tension. LOX injection is particularly important because mixing process
modelling will be affected and at an oxidizer-to-fuel ratio (O/F) of 1.4 it accounts for 58% of the
injected mass into the chamber. Liquid oxygen expansion is discussed at the end of this chapter.

CHAPTER 5. GAS-GENERATOR PROPELLANT, CHEMICAL KINETICS,

AND

LOX EXPANSION

Species

Mass Fraction

Tc , K

Pc , MPa

c , kg/m3

O2

1.000

154.59

5.043

438.34

CO
CH4
HCl
H2 O
CO2
N2
H2

0.203
0.126
0.124
0.086
0.055
0.047
0.044

132.86
190.56
324.7
647.14
304.13
126.21
32.97

3.494
4.599
83.1
220.6
73.75
3.398
1.293

430.92
162.7
449.9
321.7
468.19
311.26
31.01

46

Thermodynamic fluid properties at the critical point listed by the mass fraction present in the
combustion chamber at the injector. The values in bold are near or below the chamber operating
conditions. Note: O2 is injected separate to the fuel gases. Source: [14].

5.2

Gas-Generator Modelling

Generation of a gas through the combustion of a solid fuel grain occurs in the initial stage of the
SCAIH rocket combustion cycle. In this stage, the solid grain undergoes incomplete combustion at
cooler than adiabatic flame temperatures to yield a fuel-rich propellant gas. This fuel-rich combustion process is much akin to fuel pyrolysis.
The composition of the solid fuel grain is given in Table 5.2. The product gases from combustion of
these solid fuel components has been computed here with the Chemical Equilibrium and Applications (CEA) software. Developed by NASA Lewis Research Center, CEA can obtain chemical equilibrium compositions for assigned thermodynamic states, such as temperature and volume [55].
Using CEAs theoretical rocket performance capabilities the solid fuel grain is treated independently
as a rocket motor with product gases expanded in equilibrium through a nozzle. The calculation assumes an infinite area combustion chamber, adiabatic combustion, and isentropic expansion
through the nozzle. Flow conditions and species mass fractions resulting from product gasses being
expanded from 5 MPa to 4 MPa are in Table 5.3. The reference pressures have been determined by
CTIs in house experimental testing apparatus and pressure traces indicate solid fuel burn times of
nearly 60 seconds at a pressures close to 5 MPa [17].
The lack of oxidizer in the fuel-rich solid grain and presence of a high mass-fraction of solid carbon
particulate keeps the temperature relatively low while adding to the gas-generator density. This
calculated temperature and velocity are used as boundary condition inputs for the gas-generator
gases. The cross sectional area of the injector is fixed and so mass flow rate is determined with the
flow velocity input. Density values calculated by CEA in Table 5.3 are for the gas phase mixture

CHAPTER 5. GAS-GENERATOR PROPELLANT, CHEMICAL KINETICS,

Component

AND

LOX EXPANSION

Mass Fraction

Function

0.40
0.25
0.28
0.07

Oxidizer
Fuel
Fuel/Binder
Binder

Ammonium perchlorate (AP)


Polyethylene
Hydroxyl-terminated polybutadiene (HTPB)
Dioctyl adipate (DOA)

Solid fuel grain gas generator components as detailed by CTI. These fractions represent an approximation of the overall content as exact mixture fractions are difficult to determine when dealing
with large and complex chain polymers. HTPB, (OH)2 (C4 H6 O0.1 )n , is a common solid rocket fuel
(see Table 1.3).

Gas-generator
p = 5 MPa
u
M
T
mix
total
Species
C (solid)
CO
CH4
HCl
H2 O
CO2
N2
H2
NH3
C 2 H6
C 2 H4
HCN
CH3 Cl

0
0
1199.65 K
10.538 kg/m3
15.255 kg/m3
Mass Fraction
0.309
0.214
0.127
0.124
0.082
0.051
0.047
0.044
< 0.0005
< 0.0001
< 0.0001
< 0.0001
< 0.0001

Combustion chamber
p = 4 MPa
u
M
T
mix
total
Species
C (solid)
CO
CH4
HCl
H2 O
CO2
N2
H2
NH3
C 2 H6
C 2 H4
HCN
CH3 Cl

460.1 m/s
0.630
1172.38 K
8.6635 kg/m3
12.6261 kg/m3
Mass Fraction
0.314
0.203
0.126
0.124
0.086
0.055
0.047
0.044
< 0.0005
< 0.0001
< 0.0001
< 0.0001
< 0.0001

Generator gas composition under expansion from p = 5 MPa in the gas-generator combustion stage
to p = 4 MPa in the second stage combustion chamber. Density of the gas phase mixture is mix ,
and density of the gas and solid phase combined is total .

47

CHAPTER 5. GAS-GENERATOR PROPELLANT, CHEMICAL KINETICS,

solid
gas

AND

LOX EXPANSION

C(s) + O2

CO2

CH4 + 2 O2
2 CO + O2
2 H2 + O2
4 HCl + O2

CO2 + 2 H2 O
2 CO2
2 H2 O
2 H2 O + 2 Cl2

48

Simplified multi-species one-step reaction mechanisms to model combustion of the gas-generator


gases with oxygen (N2 is inert).

only and need to be adjusted to include the solid carbon mass fraction. This value is not used as an
input parameter because density evolves according to temperature and pressure. Simulation results
indicate predicted density of the gas-generator mixture in the inlet region is around 13.1 kg/m3 .
This slight disparity with CEA (12.6261 kg/m3 ) can be explained by the infinite-area-combustor
assumption used by the CEA software; among other things it assumes instantaneous and complete
combustion with no geometric constraints.

5.2.1

Simplified Chemical Kinetic Scheme for Propellant Gases

Using the CEA composition for the solid fuel gas-generator shown in Table 5.3, an extension to
the single one-step mechanism for methane discuss previously in Chapter 3 can be introduced by
considering a set of one-step mechanisms for each fuel component. For simplicity the trace products
(< 0.0005 mass fraction) will be ignored; only CO, CH4 , HCl, H2 O, CO2 , H2 , N2 and solid carbon
will be considered in the present work. Reduced one-step combustion mechanisms are shown
in Table 5.4 where the main fuel components (CO, CH4 , HCl and H2 ) participate in single-step
reactions with oxygen. Nitrogen, N2 is treated as an inert species and does not react. C(s) is soot
or solid carbon particles that have the thermodynamic properties of graphite. The combustion of
C(s) is omitted in this work and its inclusion in the model is discussed next.
As mentioned previously, chemical reaction rates for the mechanism are determined by the eddy
dissipation model for turbulent combustion (Equation 3.32). Keep in mind that while this will
account for more species, the EDM model tends to over-simplify the chemical structure of flames
and is not capable of providing detailed information regarding turbulent flame composition and
structure. This may be addressed in future studies.

CHAPTER 5. GAS-GENERATOR PROPELLANT, CHEMICAL KINETICS,

5.2.2

AND

LOX EXPANSION

49

Treatment of Soot and/or Solid Carbon Particles

Insufficient pressures and temperatures for complete combustion often lead to the formation of
nanoparticles called soot. This is common in applications such as gas turbine combustors, diesel
engines, and solid rocket motors. The presence of soot will affect local temperature concentrations
and radiative transport that in turn affects reaction rates and local species concentrations.
The breakdown of fuel and recombination into soot occurs through a series of chemical reactions;
the intermediate aspects of which are largely misunderstood. Within diffusion flames soot is formed
in a time frame of around ten milliseconds [82]. This is the time it takes for a fuel molecule to break
down into smaller hydrocarbon species and begin to form soot particles on the order of a nanometer
in size. These smallest soot particles consist of cyclic hydrocarbon structures with considerable
amounts hydrogen and oxygen. Soot formation and oxidation is not mutually exclusive and will
continue recursively as the formed soot is broken down in the presence of excess oxygen and
heat [83].
Thus far soot has not been described in much detail and all combustion and fluid processes are
only valid for gas-phase interactions. Section 4.4 mentioned soot radiation and its importance due
to the large mass fraction of carbon particles that are injected into the combustion chamber. With
a mass fraction of 31.4% the carbon particles will dramatically alter the density of the gaseous
mixture and influence other flow properties. While it is clear that soot particles can react and burn,
the eddy dissipation model for determining reaction rates is based on gaseous combustion without
accommodation for reactions between a solid and a gas phase.
In the present modelling, the solid carbon particulates are treated as an inert constituent having
a fixed composition, in this case that of carbon in its graphite state ( = 2.16 g/cm3 ). The thermodynamic properties of the gaseous-solid mixture are set to match that of the ideal gas mixture
with the density and energy adjusted to include the solid particles. The density of the mixture of N
gases, mix , is calculated from the idea gas law as
mix =

P
Rmix T

(5.1)

where Rmix is the gas constant of the mixture defined by


Rmix =

N
X

ci R i

(5.2)

i=1

where ci is the individual mass fraction and R i is the individual gas constant. When carbon is

CHAPTER 5. GAS-GENERATOR PROPELLANT, CHEMICAL KINETICS,

AND

LOX EXPANSION

50

involved, the total density of the gaseous and solid mixture, total , can be expressed as
total =

mix
1 cs

(5.3)

where cs is the mass fraction of the solid carbon. The gaseous species mass fractions must then sum
to unity, including the carbon mass fraction, as
N
X

ci + cs = 1 .

(5.4)

i=1

The initial mixture entering the injector is defined to have a set mass fraction of carbon. At this
point, the solid carbon does not participate in any reactions and behaves inert; similar to nitrogen.
A detailed soot model needs to incorporate a gas-phase kinetic model for the formation, a sootchemistry sub model for gas-particle conversion, and an aerosol dynamics model for inter-particle
collision and heat and mass transfer. Stewart has modelled soot formation in non-premixed kerosene
flames [84], and recently Byun has applied this model to a liquid kerosene/LOX rocket engine [65].
The modelling of soot formation and oxidation is out of the scope of this thesis. For a comprehensive overview of these subjects see the thesis by Charest [74] and references therein.

5.3

Liquid Oxygen Expansion

As mentioned previously in Section 5.1.2, many of the species in the fuel gas mixture will be
near their respective critical point properties. Liquid oxygen is no exception. LOX must be stored
below 90.2 K (at 1 atm) and cannot be liquefied above its critical temperature of 154.59 K. The
computational challenge of representing LOX comes from the large gradients incurred from its
expansion into a combustion environment and the mixing properties needed to model reactions.
An equation of state (EOS) is needed to model the thermodynamic behavior of the liquids and
gasses in this regime via an equilibrium steady-state expansion process.

5.3.1

Peng-Robinson Equation of State

The Peng-Robinson EOS [85] is used in this study to represent the thermodynamic behaviour of
the liquid oxygen stream and the phase transition from a liquid to a gas. It has been selected for
its superior handling of liquid densities as well as species critical properties and has the following
form

CHAPTER 5. GAS-GENERATOR PROPELLANT, CHEMICAL KINETICS,

p=

RT
vb

AND

LOX EXPANSION

a(T )

51

(5.5)

v(v + b) + b(v b)

in which

a(T ) = a(Tc ),

a(Tc ) = a0

R2 Tc2
Pc

b = b0

RTc
Pc

(5.6)

and


p
1/2 = 1 + 1 T /Tc ,

= 0.37464 + 1.54226 0.269922

a0 = 0.457235528921,

b0 = 0.0777960739039

(5.7)
(5.8)

The Peng-Robinson EOS is used to predict flow conditions of the LOX stream via an isentropic
conservation-of-energy approach. Using the known mass-flow rate from CTI and LOX reservoir
conditions of 5 MPa and 90 K, expansion of the LOX is handled by calculating the amount of energy
required to heat and vapourize the LOX injection stream up to the saturated vapour line. This
amount of energy is then be subtracted from the enthalpy of the fuel gas for the purpose of multispecies ideal gas simulations. This also yields an estimate for the injection velocity. The resulting
LOX state at the entrance to the combustion chamber is in the table below.

260.2

m/s

148.86

207.02

kg/m3

There are limitations to this simple estimate: firstly from the isentropic assumption that has neglected friction loses which will slow the flow of both fuel and LOX, and secondly from limitations
due to the Peng-Robinson EOS which tends to be inaccurate for liquid phases overestimating density. Results presented in Chapter 7 predict that in the inlet region the density of the LOX is around
130 kg/m3 . This difference is because the solution method is using the ideal gas equation of state
which, for oxygen near its critical point, notably under-predicts the value. Addressing the real fluid
nature of cryogenic injection is something to be considered in future work and briefly discussed in
Section 8.2.

Chapter 6

Numerical Solution Method


The numerical solution method here refers to the computational algorithmic process that is used
to solve the equations of fluid motion at points in time. The Navier-Stokes equations governing
compressible fluid flow (Equations 2.12.3) as well as the RANS and FANS variants are non-linear
partial differential equations (PDEs) in time. Transforming these PDEs into a set of algebraic equations via spatial and temporal discretization schemes allows for the computation of solutions using
various iterative methods that can be implemented in a computer using a programming language.
This requires representing the variables at some set of discrete locations in both space and time.
Segmenting the spatial domain into smaller portions for individual processing is called the spatial
discretization of the problem of which there are two commonly used types: finite-difference methods and finite-volume methods. Finite-difference methods define the computational nodes whereas
finite-volume methods (FVM) define the control volume boundaries. Finite-difference methods involve solving the differential form of the conservation equations, (Equations 2.12.3), are simple
and effective especially on structured grids and may be generalized to arbitrary orders of accuracy [86]. A disadvantage is that they need special attention to ensure global conservation and
to work with complex geometry. See the text by Ferziger [87] for descriptions of other numerical
methods for CFD, including the finite-element method. This chapter will discuss the finite-volume
approach which is the method used herein.
Once a discretization method has been selected there are a few other components to a complete
numerical solution method. Namely a grid and coordinate reference frame, algebraic solver, and
convergence criteria must all be chosen. Often these choices are problem dependent, and in this
work, techniques suited for the case of a SCAIH rocket motor are considered.

52

CHAPTER 6. NUMERICAL SOLUTION METHOD

6.1

53

The Finite Volume Method

The FVM discretizes the domain by dividing it into smaller, contiguous control volumes. An advantage of the FVM is that it is globally conservative. That is, conservation of the flow properties
are maintained in each individual control volume, and thus, the scheme is automatically globally
conservative at the discrete level. A second advantage of the FVM discretization is that it can be
applied to more general unstructured meshes. The FVM scheme that will now be discussed in detail has been developed numerically by a number of authors and is suitable for solving a system of
partial-differential equations governing two-dimensional axisymmetric laminar and turbulent compressible flows for reactive thermally perfect gaseous mixtures using a fully coupled finite-volume
formulation on body-fitted multi-block quadrilateral mesh [8893].

6.2

Two-Dimensional Axisymmetric Form of Navier-Stokes Equations

The Favre-averaged form of the Navier-Stokes equations for a compressible reactive gaseous mixture, including solid carbon particulates as described in Chapter 3, can be re-expressed using vector
notation as

U
t

+ ~F = S .

(6.1)

The flux dyad, ~F, contains the inviscid and viscous fluxes in both the radial and z directions such
that
~F = (FI FV , GI GV )

(6.2)

and the source term, S, is composed of an axisymmetric inviscid, SaI , axisymmetric viscous, SaV ,
turbulent, St , and finite-rate chemistry portion, Sc ,
S=

1
r

(SaI SaV + St + Sc ) .

(6.3)

The dyad and vector definitions for a two-dimensional axisymmetric coordinate system where r
and z are the axial and radial components, respectively, are as follows:
U=

, e
vr , e
vz , e
e, k, , Ye1 , . . . , YeN

iT

(6.4)

CHAPTER 6. NUMERICAL SOLUTION METHOD

e
vr

vr2 + p
e

vz
e
vr e

(e
e + p)e
vr

e
vr k
FI =

e
vr

e
vr Ye1

..

e
vr YeN

54

r r + r r

rz + rz

q r q t r + ( + t ) k + WF
x

k
( + t ) r
FV =

( + t ) r

J r1 J t1r

..

N
J r J tNr

(6.5)

where WF = vr ( r r + r r ) + vz ( rz + rz ),

e
vz

e
vz e
vz

e
vz2 + p

(e
e + p)e
vz

e
vz k
GI =

e
vz

e
vz Ye1

..

e
vz YeN

zr + zr

zz + zz

qz q tz + ( + t ) k + WG
z

k
( + t ) z
GV =

( + t ) z

Jz1 J tz1

..

N
Jz J tzN

(6.6)

where WG = vr (z r + z r ) + vz (zz + zz ),

e
vr

e
vr2

e
vr e
vz

(e
e + p)e
vr

e
vr k
SaI =

e
vr

e
vr Ye1

..

e
vr YeN

( r r + r r ) ( + )

( rz + rz )

q r q t r + ( + t ) k + WS
r

k
(
+

)
SaV =
t
r

(
+

)
t

J r1 J t1r

..

N
J r J tNr

(6.7)

CHAPTER 6. NUMERICAL SOLUTION METHOD

55

where WS = vr ( r r + r r ) + vz ( rz + rz ),

St = P k

P 2
k

..

0
with
P = r r

e
vr
r

+ rz

Sc = 0

..
.

e
vz

e
vr

e
vr
z

e
vz

+ zz

(6.8)

(6.9)

The reader is directed to the Ph.D. theses by Jha [94] and Gao [95] for more detailed and threedimensional development of equation formulation for a reactive mixture of compressible gases.

6.3

Integral Form of Navier-Stokes Equations and FVM Formulation

The proposed FVM scheme solves the preceeding conservation equations in integral form applied
to each computational cell. Applying the divergence theorem to the vector form of the system of
governing equations given in Equation (6.1), the following integral form can be obtained for a
two-dimensional control area, A:
d
dt

Z
A(t)

UdA +

~n ~F dC =
C(t)

Z
SdA,

(6.10)

A(t)

where U is the solution vector of conserved variables, ~F is the flux dyad, and S is the vector of
source terms. Here, C is the closed contour around the cell and ~n is a unit normal vector pointing
outward. In this form it can be seen that a flow property can be expressed as a balance between
the net flux and the source. If one then considers the application of Equation (6.10) to a given
computational cell, the average value of U and S inside the cell area, A, are defined as
U

Z
UdA,

A
A

(6.11)

CHAPTER 6. NUMERICAL SOLUTION METHOD

56

Z
SdA.

(6.12)

Substituting (6.11) and (6.12) into (6.10) yields


dU
dt

1
A

~F ~n d` = S(U) ,

(6.13)

where d` is an element of the closed contour containing the cell of interest. Since a structured,
multi-block body-fitted grid is used, a semi-discrete form of (6.13) is adopted by seeing that the
change in conserved variable, U, is a sum of the fluxes leaving the control cell such that
d
dt

Ui, j =

Nf
1 X
(~F ~nl)i,j,k + Si,j,k
Ai, j k=1

(6.14)

where li, j is the length of the kth face of cell (i, j) and ~nk is a normal vector to that side. Figure 6.1
shows a k t h cell representation on a structured grid for illustrative purposes; the approach is the
same regardless of grid choice.

Two-dimensional cell on a structured grid.

The general FVM procedure to solve (6.14) involves three steps summarized as follows.
1. Cell Reconstruction: Using the known average value, U, construct a piece-wise approximation of U within the cell which agrees with the overall average, U. This step is a form
of interpolation and referred to as cell reconstruction turning an average value, U, into a
function of the spatial coordinates, U(~x ).
2. Flux evaluation: Find the flux, F(U) at the cell boundary and compare to the adjacent boundary. Because the reconstruction is independent for each cell, a boundary flux will have two

CHAPTER 6. NUMERICAL SOLUTION METHOD

57

unique values resulting in a discontinuity between cells (Figure 6.2). Myriad methods for
resolution of the flux differences along the cell boundary have been developed.

Left and right solution states separated by a flux discontinuity at x = 0.

3. Time Evolution: Advance the solution in time, known as time-marching, to obtain new
values for U.

These steps will now be expanded upon, beginning with cell reconstruction. A one-dimensional
visualization of the FVM is shown in Figure 6.3.

One-dimensional representation of computational cell and finite-volume procedure.

CHAPTER 6. NUMERICAL SOLUTION METHOD

6.3.1

58

Cell Reconstruction

Cell Reconstruction begins with the average value across the cell domain, U, and uses it to define
the spatial distribution of the solution within the cell. This converts the average value into a
function of the spatial coordinates, U(~x ). Information from neighbouring cells is used to calculate
the reconstruction and simply using the left and right cell-centers to determine the boundary flux
gives a first-order approximation. For better accuracy, more information is required and can come
from, for example, the next-to-nearest cell neighbours which uses a larger stencil size. A leastsquares approach is used to determine the gradients at the cell interface while minimizing the sum
of errors taken over the whole stencil.
It has been shown by Godunov that there are no schemes higher than first order accurate that
preserve monotonicity [96]. The monotonic behaviour of the scheme is important for both shock
capturing and stability and so a considerable amount of research has gone into higher-than-first
order reconstructions. Non-linear reconstruction methods were first introduced by Boris and Book
with their flux-corrected transport method in 1973 [97]. More sophistication led to the total variation diminishing schemes by Harten [98] and recently another class of schemes called essential
non-oscillatory has showed promise [99].
A combination of 1st and 2nd order methods can be used such that 2nd order methods are used
in smooth regions and 1st order methods in shock or discontinuous regions. Keeping a scheme
monotonic is enforced through the use of slope-limiters [100,101]. The purpose of a slope limiter is
to prevent local minima and maxima by ensuring the solution state at the cell boundary is between
the minimum and maximum average cell value across the neighbouring cells of interest. This will
ensure monotonicity while still allowing the scheme to capture discontinuities. Well known limiters
include those by Barth-Jesperson [102] and the one used here by Venkatakrisnan [103].
For this work, the linear reconstruction is performed using a piecewise limited least-squares method
with the Venkatakrisnan slope limiter. This limiter improves upon work done by Van Albada [104]
by producing better convergence. Beginning with a cell (i, j), at the cell interface (i+ 21 , j), the flux
has the form




~F(i, j, m) ~n(i, j, m) = ~F R WL , WR , ~n(i, j, m) ,

(6.15)

where ~n is the unit normal vector of the cell interface, R is the solution to the Riemann problem,
and WL and WR are the left and right primitive solution vectors from the piece-wise limited linear

CHAPTER 6. NUMERICAL SOLUTION METHOD

59

reconstruction procedure given by


WL = Wi, j + i, j Wi, j d~x L ,
WR = Wi+1, j + i+1, j Wi+1, j d~x R

(6.16)

where is the slope-limiter. The quantities d~x L = ~x ~x i, j , d~x R = ~x ~x i+1, j are the cell interface
locations. Wi, j, k and Wi+1, j, k are cell-averaged primitive solution vectors.
Venkatakrishnans slope limiter, , is given by

Wmax Wi, j

W W

 k i, j 
W W
i, j =
WminW i, j

k
i, j

for

Wk Wi, j > 0

for

Wk Wi, j < 0 ,

(6.17)

otherwise

where ( y) is a smooth function given by


( y) =

y2 + 2 y
y2 + y + 2

(6.18)

and Wmax = max(Wi, j , Wneighbours ), Wmin = min(Wi, j , Wneighbours ), and Wk is the unlimited reconstructed solution value at the kth flux quadrature point.

6.3.2

Flux Evaluation

Mentioned earlier was the advantage that FVMs are globally conservative. This is due to the property that flux leaving one cell must enter the adjacent cell. Evaluation of the flux at the cell boundary is a crucial component of the FVM. This localized problem that occurs at the boundary between
two cells (2D) or volumes (3D) is known as a Riemann problem.
Shown graphically in figure 6.4 the Riemann problem is a special form of a one-dimensional initial
value problem whereby a discontinuity is separated by two initial states. These two states represent
the fluxes occurring at the boundary between adjacent cells.
The first method for solution to the Riemann problem came in 1959 when Godunov proposed a
scheme for use with non-linear hyperbolic conservation systems that is both able to resolve discontinuities and monotonic [96]. Monotonicity refers to a solution being strictly increasing or strictly
decreasing in time whereas non-monotonic solutions have oscillatory behaviour which leads to
instability. Godunovs method uses an exact solution to the Riemann problem to resolve discontinuities at the cell boundaries. This is ideal, however, it comes with added computational expense.

CHAPTER 6. NUMERICAL SOLUTION METHOD

60

cell interface

~
F

UL

UR

i+ 12

i+1

The Riemann problem showing a left and right state separated by a flux at the cell interface.

Modern exact Riemann solvers such as one by Gottlieb and Groth [105] are many factors more
efficient than Godunovs first, however, often an approximation is sufficient for a finite-volume application since only the cell flux is required. Recall Equation (6.2) which shows the flux is composed
of a viscous and an inviscid component.

Inviscid Flux Evaluation

The inviscid flux is hyperbolic in nature and evaluation is handled by approximate Riemann solvers
which are computationally attractive especially when handling problems with added complexity
such as: advanced equations of state, chemical kinetics and turbulence. These add transport
equations to the system thereby increasing expense. Approximate, non-iterative methods can provide the necessary informationthe flux at the cell interface. For these reasons turbulent combustion simulations regularly use approximate Riemann solvers; most commonly those by Roe in
1981 [106], Osher in 1982 [107], Harten, Lax, and van Leer (HLL) in 1983 [108], Einfeldt (HLLE)
in 1988 [109], and Toro et. al (HLLC) in 1994 [110].
The approximate Riemann solver used in this work is the advection upstream splitting method
(AUSM) that was introduced because of its simplicity and ability to handle a wide range of Mach
numbers. In its third revision, the AUSM+ -up scheme is valid at all speed regimes and in a
Mach-number-independent fashion [111113]. AUSM+ -up differs from other approximate Riemann solver by splitting the flux into two physically distinct parts; a pressure flux and a convective
flux. See the paper by Liou for the detailed implementation [113].

CHAPTER 6. NUMERICAL SOLUTION METHOD

61

Viscous Flux Evaluation


The viscous flux component of the system is elliptic in nature and therefore should not be upwinded.
It requires a different approach because it depends on the solution state as well as the solution
gradient across the cell boundary,
FV = FV (Wi+ 1 , j, k , Wi+ 1 , j, k ).
2

(6.19)

The vector of primitive solution variables, Wi+ 1 , j, k , is averaged using the left and right recon2

structed solution states, WL and WR described above. Evaluation of the gradient, Wi+ 1 , j, k is
2

done with a least-squares approach. The results here use a Green-Gauss integration procedure over
a diamond reconstruction path [95, 114].

6.3.3

Time Marching

Once discretized in space, the original set of PDEs (Equations 3.113.13) has been transformed into
a set of coupled set of non-linear ordinary differential equations (ODEs). Next, the solution needs
to be discretized in time and advanced forward. These methods are called time-marching methods
and seek to solve either a time-accurate case where the flow is unsteady, or a time-independent
case where the flow is steady. Alternatively a steady-state can be reached by integrating out the
transient, or unsteady, portion of the equations leaving the solution sufficiently close to a steadystate. Explicit methods popular in CFD include those of Euler, Adams-Bashforth, and Runge-Kutta
and implicit methods ones are the Implicit Euler, Trapezoidal, and Adams-Moulton. These methods
have various advantages and disadvantages depending on the application. Lomax et al. [115] as
well as Hirsch [116, 117] have detailed texts regarding time-marching methods.
The computational results presented here are mostly for steady-state forms which have employed
a multi-stage optimal smoothing time-marching algorithm developed by van Leer et al. [118]. As
the name suggests, between 2 and 6 stages have been developed in the scheme that give optimal
damping or smoothing of high-frequencies for a given spatial-differencing operator. This scheme is
known to perform better than Runge-Kutta methods and works well with multi-grid topographies
(although multi-grid implementation was not considered in this work). An M-stage time integration
of (6.14) from time t n to t n+1 is

U0 = Uni,j,k

i,j,k
m stage:
Um
= U0i,j,k m t n Ri,j,k (Um1 ) m = 1 M ,
i,j,k

n+1
Ui,j,k = UM
i,j,k

CHAPTER 6. NUMERICAL SOLUTION METHOD

62

where t n is the time step size and m are the multi-stage coefficients. See the paper by van Leer
for the m values.
In a few cases, time-accurate results were required to asses solution stability. For these unsteady
flow field simulations, a standard explicit second-order accurate Runge-Kutta time marching scheme
was used.

6.4

Radiation Transfer Equation Evaluation

The inclusion of radiative heat transfer is handled by the optically thin approximation and the
discrete ordinates method. For the optically thin approximation, the additional source term, ~qrad ,
is evaluated along with the other source terms as part of S in Equation (6.10). This has minimal
impact on the overall computational performance.

6.4.1

Discrete Ordinates Method

Using the DOM for the RTE evaluation is more computationally expensive due to the unknowns
associated with non-gray radiation. Because of this the RTE is decoupled from the flow equations
and solved sequentially in a loosely-coupled fashion at each time-step [74].
The DOM approximation to Equation (4.14) is
ml (r I,iml )
r

1 (ml I,iml )
r

+ ml

I,iml
z

= k (g i ) I b I,iml

(6.20)

where m and l are the polar and azimuthal direction indices for M different polar directions and
L different azimuthal directions, respectively. This is a set of ODEs that is solved using a spacemarching procedure by Carlson and Lathrop [119].

6.5

Parallel Architectures

Domain decomposition involves breaking down the computational domain into smaller sub-domains.
In CFD this involves taking a mesh and cutting it into smaller sub-meshes, or blocks. In this manner
individual blocks can be distributed to different processors in a parallel architecture. Important to
this process is the communication between blocks to ensure that computational information can be
passed back and forth so to act as a single domain.

CHAPTER 6. NUMERICAL SOLUTION METHOD

63

The multi-block quadrilateral mesh used in this thesis lends itself naturally to domain decomposition and enables efficient and scalable implementations of the solution algorithm for reactive
gaseous mixture conservation equations on distributed-memory multi-processor architectures [88
90, 93].
A parallel implementation of the block-based adaptive mesh refinement scheme has been developed
using the C++ programming language and the message passing interface (MPI) library [120]. The
domain decomposition procedure in place is a highly scalable parallel algorithm that has been
applied to myriad regimes including laminar combusting flows [90], turbulent combusting flows
in two and three space dimensions [91, 94], micro-scale flows [121], sooting laminar diffusion
flames [74, 92], and compressible flows with a high-order scheme [99]. The effects of adaptive
mesh refinement and parallel scaling performance were not assessed here.
All computations were performed on a high performance parallel cluster consisting of 3857 nodes
with 16 GB RAM per node and a high-speed interconnect at SciNet HPC Consortium [122]. The
nodes each have 8 Intel Xeon E5540 cores (2.53 GHz) and are inter-connected via a non-blocking
switch with four 4x-DDR InfiniBand links.

Chapter 7

Numerical Simulation Results


The mathematical modelling for turbulent combustion and proposed numerical solution method
outlined in the previous chapters are used here to provide predictions of the SCAIH rocket motor
flowfield. All of the simulations are presented for an axisymmetric two-dimensional multi-block
body fitted mesh. The influence of mesh resolution will be investigated first, followed by characteristics of the flow field and the flame anchoring mechanism. Lastly, the effects from chamber
wall-temperature, solid carbon particulate, reacting hydrogen chloride, and radiative heat transfer
will be shown in sequence.

7.1

Baseline Input Parameters

The baseline configuration considered here was determined in accordance with parameters from
CTI and inlet flow conditions and as calculated in Chapter 5. The simulation results presented here
have input conditions as listed in Table 7.1. The user specified variables are inlet velocity, temperature, and pressure of the fuel and LOX. Additionally, the combustion chamber back pressure is set
to a constant 5 MPa and the boundary condition is subsonic outflow at the right-hand side (exit) to
the chamber. The dimensionless quantities in Table 7.1: Mach, Prandtl, and Reynolds numbers, all
depend on density which varies as the streams reach the combustion chamber, especially the LOX
which undergoes rapid expansion and warming. These are average values in the region and not
absolute.

64

CHAPTER 7. NUMERICAL SIMULATION RESULTS

user specified

for reference

u
T
P

area
Mach
Prandtl
Reynolds

m/s
K
MPa
kg/s
kg/m3
m2

65

Fuel Gas

LOX

460
1175
4

150
150
4

0.198
13.1
3.292 382 5 105
0.63
0.464
2.38 105

0.228
130
7.917 304 4 106
0.89
0.724
6.14 106

Simulation quantities at the entrance to the combustion chamber for LOX and fuel gases. Area is
injector cross-sectional area.

7.2

Mesh Resolution

To determine the mesh resolution required to afford sufficient acuracy for the FANS-based simulation (i.e., required mesh resolution such that the numerical discretization errors of the second-order
scheme are sufficiently reduced), computational results for different meshes were tested. Figure 7.1
shows views of the six levels of mesh resolution tested of increasing cell density from 14,928 cells
up to 1,363,968 cells. This corresponds to a minimum cell width in the injector post region on the
order of 40 m for Levels 5 and 6. These mesh comparison cases included 31.4% solid carbon by
mass, had adiabatic boundary conditions and excluded radiative heat transfer. Each case was run
until a steady state converged solution was reached with the exception of Figures 7.1e and 7.1f.
The simulations on these two finest meshes exhibited flow instabilities, including fluctuations in
pressure and temperature, and would not converge to a steady state.
This fluctuating behaviour is illustrated in Figure 7.2, which compares results for the six meshes
in terms of temperature, density, and mass fraction of oxygen along the injector and combustion
chamber centerline. The cases at Levels 14 reached a steady state. Levels 5 and 6 have noticeably
different behaviour with fluctuations in temperature, density and mass fraction. While certainly
not fully grid converged, it would seem that the centerline profiles are converging as the mesh is
refined.
The peak temperature, seen in Figure 7.2a, is near the exit of the combustion chamber for the
lowest resolution meshes and continues to increase through the chamber exit. The middle two
resolutions, Levels 3 and 4, show a temperature peak at 0.16 m (2700 K) and 0.1 m (2600 K)
respectively. This shows the diffusion flame heating up as the oxidizer mixes with the products and
cooling down as the fuel is consumed. Similar behaviour is seen with density, Figure 7.2b, and O2

CHAPTER 7. NUMERICAL SIMULATION RESULTS

66

mass fraction, Figure 7.2c. As grid resolution increases, Levels 14 show a smooth decreases in
O2 mass fraction. At the lower resolutions, 1 and 2, not all of the O2 is consumed before the flow
reaches the end of the chamber. The flames for the mid level (3 & 4) resolutions run out of oxygen
at 0.16 m and 0.11 m, respectively, which correlates to the maximum temperature. Both levels 5
and 6 show flames with depleted O2 at x = 0.05 m and fluctuations in the flame region.
Based on simulations done at the high resolution (Levels 5 and 6 in Figure 7.1), there was no
evidence of convergence of the numerical predictions to a steady-state solution, even after simulation times as long as 3.5 ms. The same case on a coarser mesh (Level 4 in Figure 7.1) shows a
convergence history seen in Figure 7.3 where the energy reaches a steady state after 0.4 ms. This
behaviour indicates that the SCAIH flow conditions may actually be statistically unsteady with an
oscillatory nature. Certainly this is suggested by the FANS predictions.
FANS solutions on coarse grids have too much smoothing of the turbulent quantities as well as
insufficient fine scale resolution to demonstrate fluctuations in the flow field. Solutions on mesh
Levels 14 converge to a steady state and exhibit good convergence behaviour (Figure 7.3). Residual convergence of 4 or 5 orders of magnitude was obtained for solutions on coarse grids that
generally exhibit smooth flow behavior. On the finer grid, Levels 5 and 6, convergence fell quickly
to only 1 order of magnitude and fluctuations appeared. Depending on the nature of the flow, continued grid refinement will begin to exhibit fluctuations in the flow field that have not been damped
by the turbulence model. Lian concludes that rocket motor flows of this nature will not converge
by RANS codes [123]. Other authors have shown unsteady flow behaviour using DNS [38, 124],
LES [37, 125], and unsteady RANS [126] methods for similar classes of problems.

7.3

Flow Field Characteristics

The difference between a steady-state solution on a coarse mesh (Level 4) and a quasi-steady
solution on a fine mesh (Level 5) is seen in Figures 7.47.6 which show combustion chamber
contour plots for temperature, pressure, and CH4 mass fraction. Each plot contains a comparison
between a steady-state solution on the top and a quasi-steady flow in the middle. The bottom of the
plots shows an instantaneous time-accurate snapshot which closely resembles the quasi-unsteady
behaviour for the fine mesh.
The temperature field is indicative of the overall behaviour because it concerns mixing, reactions,
and heat release. The flame in Figure 7.4 is seen developing at the interface between the fuel gas
and the LOX and spreads out slightly in the radial direction and more in the axial direction, nearly
to the last quarter of the chamber. The largest axial width of the flame is representative of the

CHAPTER 7. NUMERICAL SIMULATION RESULTS

67

(a) Level 1: 14,928 cells

(b) Level 2: 48,864 cells

(c) Level 3: 173,760 cells

(d) Level 4: 377,280 cells

(e) Level 5: 658,560 cells

(f) Level 6: 1,363,968 cells

The near-injector region of the grid showing the different levels of cell resolution. The block
outlines are not shown.

CHAPTER 7. NUMERICAL SIMULATION RESULTS

68

(a) temperature

(b) density

(c) mass fraction O2


Mesh resolution effects on temperature, density, and O2 mass fraction as taken along the injector
and combustion chamber centerline. x = 0 corresponds to the injector faceplate. Other parameters:
31.4% solid carbon by mass, adiabatic boundary conditions, no radiative heat transfer.

CHAPTER 7. NUMERICAL SIMULATION RESULTS

69

Energy convergence history for a Level 4 mesh compared with a Level 5 mesh up to t=1.0 ms.

flame temperature at around 2800 K. As will be seen in the following section, the flame anchors
upstream directly on the injector post. The quasi-unsteady result on the finer grid has two notable
differences; the first is the overall flame temperature and the second is the flame width in the radial
direction. The temperature of the unsteady case is lower in the axial direction (see Figure 7.2a)
although has similar peaks in hot pockets. The unsteady flame structures will tend to broaden
the flame and carry more unburnt fuel away from the centerline. Some of this behaviour can be
attributed to 2-D results where large-scale structures cannot exist near and across the centerline.
3-D unsteady solutions in which large-scale structures exist across the axis show both complete
combustion in this region and a narrower flame breadth [123].
The mass fraction of methane consumption is in Figure 7.5 and shows that the quasi-unsteady solution has far more unburnt methane downstream near the centerline than the steady-state solution.
This corresponds to oxygen consumption which is can be seen in Figure 7.2c for Level 4 (steady)
compared to Levels 5 and 6 (quasi-unsteady) that show nearly all the oxygen is depleted by the
time it reaches a third to a half way downstream.

CHAPTER 7. NUMERICAL SIMULATION RESULTS

Contour plots showing the difference between a steady-state, converged flowfield on a coarse mesh (Level 4 at top), a quasi-unsteady flowfield on
a fine mesh (Level 5 in middle), and an unsteady time accurate flowfield (Level 5 at bottom) for temperature.

70

CHAPTER 7. NUMERICAL SIMULATION RESULTS

Contour plots showing the difference between a steady-state, converged flowfield on a coarse mesh (Level 4 at top), a quasi-unsteady flowfield on
a fine mesh (Level 5 in middle), and an unsteady time accurate flowfield (Level 5 at bottom) for CH4 mass fraction fuel inlet mass fraction of
methane is 0.126

71

CHAPTER 7. NUMERICAL SIMULATION RESULTS

Contour plots showing the difference between a steady-state, converged flowfield on a coarse mesh (Level 4 at top), a quasi-unsteady flowfield
on a fine mesh (Level 5 in middle), and an unsteady time accurate flowfield (Level 5 at bottom) for pressure chamber back pressure is set to a
constant of 4.0 MPa

72

CHAPTER 7. NUMERICAL SIMULATION RESULTS

73

Lastly, pressure fluctuations are seen in Figure 7.6 and are dominant along the centerline as well as
the large recirculation region near the upstream chamber wall. As no perturbations are introduced
at the inflow level, the fluctuations must arise from somewhere. Pressure oscillations can come
about due to combustion chamber acoustics and a generally accepted model for sustaining these
oscillations is from the evaporation and chemical mixing processes interacting with and supporting
each other [126]. Further study of combustion instability is important to flows of this type which
operate near the critical point of the fuel and oxidizer. Unfortunately, such work is beyond the
scope of this thesis and is therefore not discussed here.

7.4

Flame Anchoring Mechanism

The flame anchoring mechanism is of importance when considering heat transfer in the region
close to the injector post tip. As mentioned in Section 2.2, in a steady state the flame will anchor in
the vicinity of the injector face. This anchoring occurs because a stagnation region will recirculate
fuel-rich gas to maintain a reaction zone for ignition.
To investigate the effects of the LOX stream velocity on the flame anchoring mechanism, results
for three different velocities are shown in Figure 7.7. These cases are for a level 4 mesh with
31.4% solid carbon and no radiative heat transfer effects. Boundary conditions on the LOX post
are set to adiabatic. Figure 7.7 shows a strong recirculation region with a reverse flow that is
feeding fuel-rich gas into the LOX post. There are 3 different LOX flow velocities compared here:
u = 75m/s, u = 150m/s, and u = 260m/s. The higher LOX velocity, u = 260m/s, (Figure 7.7c)
is the calculated condition using the Peng-Robinson EOS expansion and this shows the flame anchored very close to the LOX post, yet a finite distance away which helps to avoid excessive heating
directly on the injector. When the flame rests directly on the injector the heat can cause ablation, as
is the case in Cesaronis experimental runs. This is shown to be the case when the LOX is injected at
a lower velocity. Figure 7.7b, u = 150m/s, represents the correction using the experimental mass
flow rate provided by CTI and the flame anchoring mechanism is much larger with high temperatures > 2700K directly on the LOX post. The lowest velocity, u = 75m/s in Figure 7.7a is for
reference and similarly shows the recirculation zone increase in size with more of the LOX post exposed to temperatures > 2700K. The LOX stream at u = 260m/s has more momentum preventing
the diffusion of oxygen into the reactant zone nearest the post resulting in cooler temperatures.
Oefelein [38] has shown similar flame anchoring results for a LOX/H2 flame, albeit under notably
different flow conditions; H2 was injected at 150 K and 125 m/s and LOX at 100 K and 30 m/s.

CHAPTER 7. NUMERICAL SIMULATION RESULTS

(a) LOX inlet velocity u = 75m/s

(b) LOX inlet velocity u = 150m/s

(c) LOX inlet velocity u = 260m/s


Flame anchoring mechanism with showing recirculation regions of fuel-rich gases. Stream traces
show the recirculation zones. In all cases the flame is anchored directly on or very near to the lox
post with the hottest post in 7.7a and the coolest post in 7.7c.

74

CHAPTER 7. NUMERICAL SIMULATION RESULTS

75

(a) wall temperature = 300 K

(b) wall temperature = 600 K

(c) wall temperature = 1200 K


Isothermal boundary conditions on the combustion chamber wall and resulting temperature flow
field. Mesh resolution is level 4, species composition excludes solid carbon.

7.5

Chamber Wall Temperature

Wall temperature was also investigated here to asses any effect on the temperature profile of the
flame in the combustion chamber. For a Level 4 mesh resolution, the chamber and injector walls
were set to constant temperature boundary conditions; the LOX post boundary condition was set
to adiabatic. Figure 7.8 shows the temperature profile for three cases: Twall = 300 K, Twall = 600
K, and Twal l = 1200 K. The temperature distribution shows little change between the different wall
conditions. Maximum flame temperature is 3983 K for the Twall = 1200 K case and 3993 K for the
Twal l = 300 K case. As this has little effect on the flame field, the wall boundary conditions were
set to adiabatic for all the results in this thesis. Radiative heat transfer effects are not included, nor
is solid carbon particulate.
In addition, it was stated in Chapter 1 that CTI is mainly concerned with excessive temperatures in
the near-injector region that are causing ablation of the injector face. Conditions at the combustion
chamber wall may still affect heat transfer; however, they have not been emphasized at this time.

CHAPTER 7. NUMERICAL SIMULATION RESULTS

7.6

76

Effects of Solid Carbon

As noted previously, the gas-generator fuel stream contains as much 31.4% by mass of solid carbon
particulate determined by a CEA simulation (Table 5.3). This carbon travels into the combustion
chamber and affects the overall flow properties reducing flame temperature, offsetting reactant
compositions, and participating in further combustion and radiative heat transfer.
In the current simulations, the soot particles are set to match the temperature of the gaseous mixture and do not participate in any reactions. Therefore, even in regions where the temperature
is high enough to support additional carbon combustion the soot will not burn. Incorporating a
mechanism for soot was not investigated and this is a limitation of the present simulations (see
Section 5.2.2).
For comparison purposes and to understand the influence of the solid carbon on the LOX injector
flow field, Figure 7.9 shows a simulation without soot compared to one with 31.4% soot, all on a
Level 5 mesh. The high mass fraction of carbon present in Figure 7.9b dramatically reduces the
flame temperature with peaks around 2900 K compared to Figure 7.9a that has peaks of 3400 K.
Figure 7.9c also shows the wall temperature of the LOX post. The carbon free fuel stream has a
maximum temperature of 4000 K compared to the carbon rich stream that has a peak of 3200 K.
These high temperatures directly on the post are from fuel rich recirculation regions in the wake
of the post. The flame anchoring mechanism has been discussed in Section 7.4. The simulation
results of Figure 7.9 do not include the effects of radiative heat transfer.

CHAPTER 7. NUMERICAL SIMULATION RESULTS

(a) No solid carbon in the fuel stream.

(b) 31.4% solid carbon in the fuel stream.

(c) LOX injector post temperature profile.


Effects of carbon particulate in the fuel stream showing the temperature profile on the LOX injector
post wall (a), and the near-field temperature contours for 0% soot (b), and 31.4% soot (c).

77

CHAPTER 7. NUMERICAL SIMULATION RESULTS

reacting

inert

78

CH4 + 2 O2
2 CO + O2
2 H2 + O2

CO2 + 2 H2 O
2 CO2
2 H2 O

HCl
C(s)
N2

A simplified multi-species reaction mechanism with hydrogen chloride being treated as inert.

7.7

Effects of Chlorine

The gas-generator contains a significant amount of chlorine in the form of hydrogen chloride, HCl
(12.4% by mass), and trace amounts of methyl chloride, CH3 Cl. Table 5.4 shows the simplified
multi-species reaction mechanism which includes the decomposition of HCl into water vapour and
chlorine gas. Chlorine has the highest electron affinity of the elements and will contribute to the
heat release and flame temperature in the combustion chamber. In order to assess the importance of
the HCl combustion on the predicted results and the influence of the simple one-step representation
of the chlorine chemistry, computational results with the one-step treatment for chlorine chemistry
are compared in this section to results in which HCl is assumed to not react. Figure 7.10 compares
the effect of hydrogen chloride participating in this single-step reaction versus being treated as
inert. The resulting mechanism is for CH4 , CO, and H2 , and shown in Table 7.2.

CHAPTER 7. NUMERICAL SIMULATION RESULTS

(a) Temperature flow field for non-reacting chlorine.

(b) Temperature flow field for reacting chlorine.

Effects of reacting chlorine on the flow field.

79

(c) Flame temperature of reactive hydrogen chloride with the corresponding mass fraction of chlorine gas produced to the temperature when inert. All as a
function of distance downstream from the injector face (x = 0).

CHAPTER 7. NUMERICAL SIMULATION RESULTS

80

The flame temperature with the reacting HCl mechanism is only slightly higher overall with a peak
of 3000 K compared to the inert HCl which peaks at 2950 K. The reacting mechanism also produces
Cl2 which is expected to further interact with other species in the field, such as H2 O, however this
is not incorporated in this simplified mechanism. This case includes solid carbon particulate which
has a cooling effect on the flow as discussed in Section 7.6, and excludes radiative heat transfer. It
should be noted that the results in Figure 7.10) are the only ones with inactive chlorine, all other
cases in this thesis have HCl reacting according to Table 5.4.

7.8

Effects of Radiative Heat Transfer

Radiative heat transfer effects on the flow field are compared in Figure 7.11 for the optically thin
model, 7.11b, and the radiation transfer equation, 7.11c, against the absence of radiation. These
models have been discussed in Chapter 4. The presence of large amounts of carbon particulate ( 13
mass fraction) affects the temperature field by contributing to radiative heat loss. Precise effects
of radiation on the temperature field are difficult to determine because the peak temperatures
fluctuate over time by as much as 100 K when considering the instantaneous solutions. A general
trend can be seen however which is that without radiation modelled the peak flame temperature
is around 2900 K which is 100 K higher than the optically thin solution. Similarly, the optically
thin solution is around 2800 K which is around 100 K higher than the radiation transfer equation
(T 2700 K). This is expected because the radiation transfer equation accounts for re-absorption
whereas with the optically thin model the intensity leaving a surface travels through the domain
without any attenuation. Assuming the media to be optically thin without any attenuation will
raise the temperature of the chamber by limiting the radiative heat loss when compared to the
RTE. Additionally, the area outside of the flame in the chamber is warmest for the RTE case due to
radiative heat transfer from the flame to the surroundings.

CHAPTER 7. NUMERICAL SIMULATION RESULTS

(a) No radiation modelled peak flame temperature is around 2900 K

(b) Optically thin model peak flame temperature is around 2800 K

Effects of radiative heat transfer on the temperature field for the optically thin model (b), and the ratiation transfer equation (c), compared to no
radiation.

81

(c) Radiation transfer equation peak flame temperature is around 2700 K

Chapter 8

Conclusions and Future Research


8.1

Conclusions

The intent of this research was to propose and develop a comprehensive numerical model of a
staged combustion aft-injected hybrid rocket motor and obtain first results using this model for
a baseline LOX injector flow-field configuration. The advantages of a SCAIH motor configuration
over a traditional hybrid rocket include better mixing efficiency by using a warm gas-generator, reduced payload delivery cost, and relatively high performance. The flow conditions in the SCAIH are
complex and involve multiple species, solid carbon transport, and operating conditions in critical
point regimes, and have not been modelled to date using modern CFD tools. Particular significance is placed on the investigation of the flame anchoring mechanism and temperature field in the
high-speed, high-temperature, reacting flow with solid carbon particulate and radiative heat transfer effects. The influence of carbon soot being transported through the chamber cools the flame
temperature significantly and the effects of radiation further this cooling. The flame anchoring
mechanism is shown to rest on the injector post due to recirculation regions while at temperatures
greater than 2700 K. Although a multi-phase flow model was not considered, the LOX is shown to
be nearly consumed within a third of the chamber distance from the injector face at the mass flow
rate design point. Overall, the flow is shown to be unsteady in nature due to fluctuations which
cannot be dampened by the turbulence model in the FANS solver.
82

CHAPTER 8. CONCLUSIONS

8.2

AND

FUTURE RESEARCH

83

Areas of Future Research

This simulation of hybrid rocket combustion presents many challenges and there is considerable
room for future research due to the different specialties involved. The 2007 text by Chiaverini
and Kuo [10] outlines much of the present work on traditional hybrid rocket configurations with
considerably less emphasis on non-traditional configurations such as the staged-combustion aft
injected hybrid rocket presented here. Solid fuel pyrolysis and regression have driven much of the
research in recent decades and determined the fundamental physical processes involved for many
fuels including sophisticated mixtures like HTPB. However, often detailed chemical kinetics and
multi-phase physics are simplified or left out of studies completely.
Within the SCAIH rocket motor simulation framework developed as part of this thesis research, the
prominent areas for improved modelling include: chemical kinetics, and multi-phase spray injection, coupled with representative treatment of soot particles. Each of these are now discussed
in turn.

CHEMICAL KINETICS
Improvements to the chemical kinetics described in Sections 3.6 and 5.2.1 can include a PCM-FPI
implementation. The PCM-FPI scheme incorporates the effects of detailed chemistry on turbulent
flames and has recently shown promising developments. FPI is Flame-Prolongation of the Intrinsic
low dimensional manifold (ILDM) approach [127]. The ILDM analyses the eigenstructure of the
local source terms to identify slow chemical processes and generates a look-up table that can be
used to evaluate chemical kinetics. Vervisch et al. [128] and Domingo et al. [129] have developed
a presumed conditional moment (PCM) modelling approach used to incorporate the influence of
turbulent fluctuations on the chemical kinetics.
A PCM-FPI implementation has been developed and implemented in the groups CFFC code that
can model turbulent diffusion flames with detailed chemistry using adaptive mesh refinement in a
parallelized environment [94]. This model has not yet been used to study the cases presented here.

MULTIPHASE SPRAY MODELS


The liquid oxygen stream breakup, atomization and droplet formation are present areas of research,
since in addition to rocket motors, multi-phase mixing layer physics is an important aspect to gas
turbine combustors and diesel engines. A general review of numerical modelling issues pertaining
to jets, sprays, shear and mixing layers at relevant pressures and temperatures suggests that fluid

CHAPTER 8. CONCLUSIONS

AND

FUTURE RESEARCH

84

jet disintegration under supercritical conditions is different from subcritical atomization [130]. Recently Oefelein has conducted a direct numerical simulation investigation into the thermophysical
characteristics of a LOX/H2 flame at supercritical pressures which used a multicomponent real-gas
and liquid mixture state package, developed by the author, and found a diffusion dominated mode
of combustion along with exceedingly large thermophysical property gradients [38, 131].

SOLID CARBON/SOOT PARTICULATE COMBUSTION


Both soot formation and oxidation modelling are necessary aspects of hybrid rocket propulsion.
Soot formation in the SCAIH rocket largely occurs in the solid fuel grain (first stage) combustion and
is not the focus of this work. Solid carbon oxidation is of importance to the second stage combustion
and its modelling has been neglected here. Implementation of a solid carbon particulate oxidation
model would improve the accuracy of radiative and heat release processes as well as local species
concentrations.

Presently, hybrid rocket motors are well beyond the concept stage and research efforts will continue
to build as the market for lower cost and smaller scale launch vehicles grows. Continued CFD
research and development in the abovementioned areas will add to numerical simulation validity
as it continues to become more prominent in the development and testing of hybrid rocket motors.

References
[1] Staats, E. B., Cost-Benefit Analysis Used In Support Of The Space Shuttle Program, Tech.
rep., NASA, May 1972.
[2] Mullane, M., Riding Rockets: The Outrageous Tales of a Space Shuttle Astronaut, Scribner,
2007.
[3] Dunbar, B., Kennedy Space Center: Frequently Asked Questions, http://www.nasa.

gov/centers/kennedy/about/information/shuttle_faq.html, February 2008,


[retrieved 15 August 2011].
[4] Pielke, R. and Byerly, R., Shuttle programme lifetime cost, Nature, Vol. 472, No. 38, April
2011.
[5] Flaherty, J. M., Federal Budget, http://www.fin.gc.ca/access/budinfo-eng.asp,
2011, [retrieved 23 August 2011].
[6] Obama, B. H., Budget of the United States Government, http://www.gpo.gov/fdsys/

browse/collectionGPO.action?collectionCode=BUDGET, 2011, [retrieved 23 August 2011].


[7] Canadian Space Agency, CSA: Frequently Asked Questions, http://www.asc-csa.gc.

ca/eng/about/faqs_csa_general.asp#2, 2007, [retrieved 23 August 2011].


[8] Prentice, J., The Canadian Space Agency: 2008-2009 Estimates, http://images.

spaceref.com/news/2009/csa-eng.pdf, 2009, [retrieved 23 August 2011].


[9] Cesaroni Technology Incorporated, Canadian Small Launch Vehicle Overview, Presentation
prepared for UTIAS, January 2009.
[10] Chiaverini, M. J. and Kuo, K. K., editors, Fundamentals of Hybrid Rocket Combustion and
Propulsion, AIAA, 2007.
85

REFERENCES

86

[11] Silvera, I. F. and Cole, J. W., Metallic Hydrogen: The Most Powerful Rocket Fuel Yet to
Exist, Journal of Physics: Conference Series, Vol. 215, 2010.
[12] Hill, P. G. and Peterson, C. R., Mechanics and Thermodynamics of Propulsion, Pearson, 2nd
ed., 1992.
[13] Turns, S. R., An Introduction to Combustion: Concepts and Applications, McGraw-Hill, New
York, 2nd ed., 2000.
[14] Lide, D. R., editor, CRC Handbook of Chemistry and Physics, CRC Press, Florida, 90th ed.,
2010.
[15] Clark, J. D., Ignition! An Informal History of Liquid Rocket Propellants, Rutgers University
Press, 1972.
[16] Altman, D. and Holzman, A., Overview and History of Hybrid Rocket Propulsion, chap. 1,
AIAA, 2007.
[17] Pilon, B. and Louwers, J., Development of Staged Combustion Aft-Injected Hybrid (SCAIH)
Propulsion at Cesaroni Technology Inc, Paper 20106786, AIAA, 2010.
[18] Chiaverini, M. J., Serin, N., Johnson, D. K., Lu, Y., and Kuo, K. K., Regression Rate Behavior
of Hybrid Rocket Solid Fuels, Journal of Propulsion and Power, Vol. 16, No. 1, 2000, pp. 125
132.
[19] Lee, C., Na, Y., and Lee, G., The Enhancement of Regression Rate of Hybrid Rocket Fuel by
Helical Grain Configuration and Swirl Flow, Paper 20053906, AIAA, July 2005.
[20] Yuasa, S., Yamamoto, K., Hachiya, H., Kitagawa, K., , and Oowada, Y., Development of a
Small Sounding Hybrid Rocket with a Swirling-Oxidizer-Type Engine, Paper 20013537,
AIAA, July 2001.
[21] Nagata, H., Aikawa, N., Akiba, R., Kudo, I., Ito, K., and Tanatsugu, N., Combustion Characteristics of Propellants for Dry Towel Hybrid Rocket Motor, IAF Paper 975.2.08, 48th
International Astronautical Congress, October 1997.
[22] Rice, E. E., Gramer, D. J., St. Clair, C. P., and Chiaverini, M. J., Mars ISRU CO/02 Rocket
Engine Development and Testing, 7th NASA International Microgravity Combustion Symposium, June 2003.
[23] Haag, G., Sweeting, M., and Richardson, G., An Alternative Geometry Hybrid Rocket for
Spacecraft Orbit Transfer Manoeuvers, IAF Paper 00W.2.07, 51st International Astronautical Congress, October 2000.

REFERENCES

87

[24] Atlantic Research Corporation, Hybrid Propulsion Technology Program, Reference Publication CR-183952, NASA, January 1990.
[25] Culver, D. W., Comparison of Forward and Aft Injected Hybrid Rocket Boosters, 27th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Sacramento, June 1991.
[26] Lund, G. K., Starrett, W., and Jensen, K., Development and Lab-Scale Testing of a Gas
Generator Hybrid Fuel in Support of the Hydrogen Peroxide Hybrid Upper Stage Program,
37th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Salt Lake City, July
2001.
[27] Markopoulos, P. and Abel, T., Development and Testing of a Hydrogen Peroxide Hybrid
Upper Stage Propulsion System, 37th AIAA/ASME/SAE/ASEE Joint Propulsion Conference
and Exhibit, Salt Lake City, July 2001.
[28] Groth, C. P. T. and Glder, . L., Numerical Analysis of Injection, Mixing, and Combustion Processes in Hybrid Rocket Motors with Liquid Oxygen Injection, Research Proposal
Submitted to Cesaroni Technology Incorporated, October 2009.
[29] Farmer, R., Cheng, G. C., and Metshc, T., Analysis of steady and unsteady hybrid combustion, JANNAF propulsion and joint subcommittee meeting, 1996.
[30] Farmer, R., Cheng, G. C., Jones, S., and Arves, J., A practical CFD model for simulating
hybrid motors, JANNAF 35th combustion, airbreathing propulsion, and propulsion systems
hazards subcommittee joint meeting, 1998.
[31] Chen, Y., Chou, T., Gu, B., Wu, J., Wu, B., Lian, Y., and Yang, L., Multiphysics simulations
of rocket engine combustion, Computers and Fluids, Vol. 45, No. 1, 2011, pp. 2936.
[32] Osherov, A. and Natan, B., The Starting Transient in a Gas-Generator Hybrid Rocket Motor,
Propellants, Explosives, Pyrotechnics, Vol. 25, 2000, pp. 260270.
[33] Karabeyoglu, M., Altman, D., and Cantwell, B., Combustion of Liquefying Hybrid Propellants: Part 1, General Theory, Journal of Plasma Physics, Vol. 18, No. 3, 2002, pp. 610620.
[34] Eilers, S. D. and Whitmore, S. A., Correlation of Hybrid Rocket Propellant Regression Measurements with Enthalpy-Balance Model Predicitons, Journal of Spacecraft and Rockets,
Vol. 45, No. 5, 2008, pp. 10101020.
[35] Kalitan, D. M., Salgues, D., Mouis, A., Lee, S. Y., Pal, S., and Santoro, R. J., Experimental
Liquid Rocket Swirl Coaxial Injector Study Using Non-Intrusive Optical Techniques, Paper
2005-4299, AIAA, July 2005.

REFERENCES

88

[36] Gill, G. S. and Nurick, W. H., Liquid Rocket Engine Injectors, Reference Publication SP8089, NASA, 1976.
[37] Zong, N. and Yang, V., Near-field flow and flame dynamics of LOX/methane shear-coaxial
injector under supercritical conditions, Proceedings of the Combustion Institute, Vol. 31,
2007, pp. 23092317.
[38] Oefelein, J. C., Thermophysical characteristics of shear-coaxial LOXH2 flames at supercritical pressure, Proceedings of the Combustion Institute, Vol. 30, No. 2, 2005, pp. 29292937.
[39] Lux, J. and Haidn, O., Flame Stabilization in High-Pressure Liquid Oxygen/Methane Rocket
Engine Combustion, Journal of Propulsion and Power, 2009, pp. 1523.
[40] Hulka, J. R., Scaling of Performance in Liquid Propellant Rocket Engine Combustion Devices, Paper 2008-5113, AIAA, July 2008.
[41] Versteeg, H. K. and Malalasekera, W., An Introduction to Computational Fluid Dynamics,
Pearson, Glasgow, 2nd ed., 2007.
[42] Veynante, D. and Vervisch, L., Turbulent Combustion Modeling, Progress in Energy and
Combustion Science, Vol. 28, 2002, pp. 193266.
[43] Moin, P. and Mahesh, K., Direct Numerical Simulation: A Tool in Turbulence Research.
Annual Review of Fluid Mechanics, January 1998, pp. 539578.
[44] Jaramillo, J., Prez-Segarra, C., Oliva, A., and Claramunt, K., Analysis of different RANS
models applied to turbulent forced convection, International Journal of Heat and Mass
Transfer, Vol. 50, No. 19-20, 2007, pp. 37493766.
[45] Tennekes, H. and Lumley, J. L., A First Course in Turbulence, The MIT Press, Cambridge,
Massachussets, 1983.
[46] Kolmogorov, A. N., Local structure of turbulence in an incompressible fluid for very large
Reynolds numbers, Comptes rendus (Doklady) de lAcadmie des sciences de lU.R.S.S., Vol. 31,
1941, pp. 301305.
[47] Kolmogorov, A. N., On degeneration of isotropic turbulence in an incompressible viscous
liquid, Comptes rendus (Doklady) de lAcadmie des sciences de lU.R.S.S., Vol. 31, 1941,
pp. 538540.
[48] Pitsch, H., Large-Eddy Simulation of Turbulent Combustion, Annual Review of Fluid Mechanics, Vol. 38, 2006, pp. 453482.

REFERENCES

89

[49] Vasilyev, O. V., Lund, T. S., and Moin, P., A General Class of Commutative Filters for LES in
Complex Geometries, Journal of Computational Physics, Vol. 146, 1998, pp. 82104.
[50] Reynolds, O., On the Dynamical Theory of Incompressible Viscous Fluids and the Determination of the Criterion, Philosophical Transactions of the Royal Society of London, Vol. 186,
1895, pp. 123164.
[51] Wilcox, D., Turbulence Modeling for CFD, DCW Industries Inc., La Caada, California, 2nd
ed., 2002.
[52] Favre, A., Equations des gaz turbulents compressibles, Journal de Mcanique, Vol. 4, No. 4,
December 1965, pp. 391421.
[53] Boussinesq, J., Thorie de lEcoulement Tourbillant, Mmoires prsents par divers savants
lAcadmie des Sciences, Vol. 23, 1877, pp. 4650.
[54] Anderson, J. D., Modern Compressible Flow with Historical Perspective, McGraw-Hill, New
York, 1990.
[55] Gordon, S. and McBride, B. J., Computer Program for Calculation of Complex Chemical
Equilibrium Compositions and Applications I. Analysis, Reference Publication 1311, NASA,
1994.
[56] McBride, B. J. and Gordon, S., Computer Program for Calculation of Complex Chemical
Equilibrium Compositions and Applications II. Users Manual and Program Description, Reference Publication 1311, NASA, 1996.
[57] Smith, G. P., Golden, D. M., Frenklach, M., Moriarty, N. W., Eiteneer, B., Goldenberg, M.,
Bowman, C. T., Hanson, R. K., Song, S., Gardiner, W. C., Lissianski, V. V., and Qin, Z., GRIMech 3.0, http://www.me.berkeley.edu/gri_mech/.
[58] Lindstedt, R. P. and Maurice, L. Q., Detailed chemical-kinetic model for aviation fuels,
Journal of Plasma Physics, Vol. 16, 2000, pp. 187195.
[59] Maas, U. and Pope, S. B., Simplifying Chemical Kinetics: Intrinsic Low-Dimensional Manifolds in Composition Space, Combustion and Flame, 1992, pp. 239264.
[60] Spalding, D. B., Mixing and chemical reaction in steady confined turbulent flames, Thirteenth Symposium (International) on Combustion, The Combustion Institute, 1971, pp. 649
657.

REFERENCES

90

[61] Magnussen, B. and Hjertager, B., On mathematical modeling of turbulent combustion with
special emphasis on soot formation and combustion, Sixteenth Symposium (International)
on Combustion, The Combustion Institute, 1976, pp. 719729.
[62] Modest, M. F., Radiative Heat Transfer, Academic Press, New York, 2nd ed., 2003.
[63] Viskanta, R. and Meng, M. P., Radiation heat transfer in combustion systems, Progress in
Energy and Combustion Science, Vol. 13, 1987, pp. 97160.
[64] de Ris, J., Fire radiationA review, Proceedings of the Combustion Institute, Vol. 17, No. 1,
1979, pp. 10031016.
[65] Byun, D. and Baek, S. W., Numerical investigation of combustion with non-gray thermal
radiation and soot formation effect in a liquid rocket engine, International Journal of Heat
and Mass Transfer, Vol. 50, 2006, pp. 412422.
[66] Coelho, P., Approximate solutions of the filtered radiative transfer equation in large eddy
simulations of turbulent reactive flows, Combustion and Flame, Vol. 156, No. 5, 2009,
pp. 10991110.
[67] Kabashnikov, V. and Kmit, G., Influence of turbulent fluctuations on thermal radiation,
Journal of Applied Spectroscopy, Vol. 31, No. 2, 1979, pp. 9637.
[68] Dworkin, S. B., Smooke, M. D., and Giovangigli, V., The impact of detailed multicomponent
transport and thermal diffusion effects on soot formation in ethylene/air flames, Proceedings of the Combustion Institute, Vol. 32, No. 1, 2009, pp. 11651172.
[69] Smooke, M. D., Long, M. B., Connelly, B. C., Colket, M. B., and Hall, R. J., Soot formation
in laminar diffusion flames, Combustion and Flame, Vol. 143, 2005, pp. 613628.
[70] Rothman, L. S. et al., The HITRAN 2008 molecular spectroscopic database, Journal of
Quantitative Spectroscopy and Radiative Transfer, Vol. 110, 2008, pp. 533572.
[71] Soufiani, A. and Taine, J., High temperature gas radiative property parameters of statistical narrow-band model for H2 O, CO2 and CO, and correlated-K model for H2 O and CO2 ,
International Journal of Heat and Mass Transfer, Vol. 40, No. 4, 1997, pp. 987991.
[72] Lacis, A. A. and Oinas, V., A description of the correlated k distribution method for modeling
nongray gaseous absorption, thermal emission, and multiple scattering in vertically inhomogeneous atmospheres, Journal of Geophysical Research, Vol. 96, No. D5, 1991, pp. 9027
9063.

REFERENCES

91

[73] Liu, F., Smallwood, G. J., and Glder, . L., Application of the statistical narrow-band
correlated-k method to low-resolution spectral intensity and radiative heat transfer calculations - effects of the quadrature scheme, International Journal of Heat and Mass Transfer,
Vol. 43, No. 17, 2000, pp. 31193135.
[74] Charest, M. R. J., Numerical Modelling of Sooting Laminar Diffusion Flames at Elevated Pressures and Microgravity, Ph.D. thesis, University of Toronto, 2011.
[75] Liu, F., Guo, H., Smallwood, G. J., and Glder, . L., Effects of gas and soot radiation on soot
formation in a coflow laminar ethylene diffusion flame, Journal of Quantitative Spectroscopy
& Radiative Transfer, Vol. 73, 2002, pp. 409421.
[76] Felske, J. D. and Tien, C. L., Use of the Milne-Eddington absorption coefficient for radiative
heat transfer in combustion systems, Journal of Heat Transfer, Vol. 99, No. 3, 1977, pp. 458
465.
[77] Lockwood, F. C. and Shah, N. G., A new radiation solution method for incorporation in
general combustion prediction procedures, Proceedings of the Combustion Institute, 1981,
pp. 14051414.
[78] Howell, J. R., The Monte Carlo method in radiative heat transfer, Journal of Thermophysics
and Heat Transfer, Vol. 120, No. 3, 1998, pp. 547560.
[79] Liu, F., Guo, H., Smallwood, G. J., and Glder, . L., Numerical modelling of soot formation and oxidation in laminar coflow non-smoking and smoking ethylene diffusion flames,
Combustion Theory and Modelling, Vol. 7, 2003, pp. 301315.
[80] Dworkin, S. B., Connelly, B. C., Schaffer, A. M., Bennett, B. A. V., Long, M. B., Smooke, M. D.,
Puccio, M. P., McAndrews, B., and Miller, J. H., Computational and experimental study of a
forced, time-dependent, methane-air coflow diffusion flame, Proceedings of the Combustion
Institute, Vol. 31, No. 1, 2007, pp. 971978.
[81] Charest, M. R. J., Groth, C. P. T., and Glder, . L., Solution of the equation of radiative
transfer using a NewtonKrylov approach and adaptive mesh refinement, Journal of Computational Physics, Vol. 231, 2012, pp. 30233040.
[82] Thomas, A., Carbon formation in flames, Combustion and Flame, Vol. 6, No. C, 1962,
pp. 4662.
[83] Haynes, B. S. and Wagner, H. G., Soot formation, Progress in Energy and Combustion Science, Vol. 7, 1981, pp. 229273.

REFERENCES

92

[84] Stewart, C. D., Syed, K. J., and Moss, J. B., Modelling soot formation in non-premixed
kerosene-air flames, Combustion Science and Technology, Vol. 75, 1991, pp. 211226.
[85] Peng, D.-Y. and Robinson, D. B., A New Two-Constant Equation of State, Industrial & Engineering Chemistry Fundamentals, Vol. 15, No. 1, 1976, pp. 5964.
[86] Pulliam, T. H. and Zingg, D. W., Foundations of Computational Fluid Dynamics, SpringerVerlag, Berlin, 2012.
[87] Ferziger, J. H. and Peri`c, M., Computational Methods for Fluid Dynamics, Springer-Verlag,
Berlin, 3rd ed., 2002.
[88] Gao, X. and Groth, C. P. T., A Parallel Adaptive Mesh Refinement Algorithm for Predicting
Turbulent Non-Premixed Combusting Flows, International Journal of Computational Fluid
Dynamics, Vol. 20, No. 5, 2006, pp. 349357.
[89] Gao, X., Northrup, S. A., and Groth, C. P. T., Parallel Solution-Adaptive Method for TwoDimensional Non-Premixed Combusting Flows, Progress in Computational Fluid Dynamics,
Vol. 11, No. 2, 2011, pp. 7695.
[90] Northrup, S. A. and Groth, C. P. T., Solution of Laminar Diffusion Flames Using a Parallel
Adaptive Mesh Refinement Algorithm, Paper 20050547, AIAA, January 2005.
[91] Jha, P. K. and Groth, C. P. T., Chemistry Approaches for Laminar Flames: Evaluation of
Flame-Prolongation of ILDM and Flamelet Methods, Combustion Theory and Modelling,
Vol. 16, No. 1, 2012, pp. 3157.
[92] Charest, M. R. J., Groth, C. P. T., and Glder, . L., A Computational Framework for Predicting Laminar Reactive Flows with Soot Formation, Combustion Theory and Modelling,
Vol. 14, No. 6, 2010, pp. 793825.
[93] Sachdev, J. S., Groth, C. P. T., and Gottlieb, J. J., A Parallel Solution-Adaptive Scheme for
Predicting Multi-Phase Core Flows in Solid Propellant Rocket Motors, International Journal
of Computational Fluid Dynamics, Vol. 19, No. 2, 2005, pp. 159177.
[94] Jha, P. K., Modelling Detailed-Chemistry Effects on Turbulent Diffusion Flames using a Parallel
Solution-Adaptive Scheme, Ph.D. thesis, University of Toronto, 2011.
[95] Gao, X., A Parallel Solution-Adaptive Method for Turbulent Non-Premixed Combusting Flows,
Ph.D. thesis, University of Toronto, 2008.

REFERENCES

93

[96] Godunov, S. K., Finite-Difference Method for Numerical Computations of Discontinuous Solutions of the Equations of Fluid Dynamics, Matematicheskii Sbornik, Vol. 47, 1959, pp. 271
306.
[97] Boris, J. P. and Book, D. L., Flux-Corrected Transport. I. SHASTA, A Fluid Transport Algorithm That Works, Journal of Computational Physics, Vol. 11, 1973, pp. 3869.
[98] Harten, A., High Resolution Schemes for Hyperbolic Conservation Laws, Journal of Computational Physics, Vol. 49, 1983, pp. 357393.
[99] Ivan, L., Development of High-Order CENO Finite-Volume Schemes with Block-Based Adaptive
Mesh Refinement, Ph.D. thesis, University of Toronto, June 2011.
[100] Barth, T. J. and Fredrickson, P. O., Higher Order Solution of the Euler Equations on Unstructured Grids Using Quadratic Reconstruction, Paper 90-0013, AIAA, January 1990.
[101] van Leer, B., Upwind and High-Resolution Methods for Compressible Flow: From Donor
Cell to Residual Distribution Schemes, Communications in Computational Physics, Vol. 1,
No. 2, 2006, pp. 192206.
[102] Barth, T. J., Recent Developments in High Order K-Exact Reconstruction on Unstructured
Meshes, Paper 93-0668, AIAA, January 1993.
[103] Venkatakrishnan, V., On the Accuracy of Limiters and Convergence to Steady State Solutions, Paper 93-0880, AIAA, January 1993.
[104] Albada, G. V., Leer, B. V., and Roberts, W. W., A Comparative Study of Computational Methods in Cosmic Gas Dynamics, Astron. Astrophysics, Vol. 108, 1982, pp. 76.
[105] Gottlieb, J. J. and Groth, C. P. T., Assessment of Riemann Solvers for Unsteady OneDimensional Inviscid Flows of Perfect Gases, Journal of Computational Physics, Vol. 78,
1988, pp. 437458.
[106] Roe, P. L., Approximate Riemann Solvers, Parameter Vectors, and Difference Schemes, Journal of Computational Physics, Vol. 43, 1981, pp. 357372.
[107] Osher, S. and Solomon, F., Upwind difference schemes for hyperbolic systems of conservation laws, Mathematics of Computation, Vol. 38, No. 158, 1982, pp. 339374.
[108] Harten, A., Lax, P. D., and van Leer, B., On Upstream Differencing and Godunov-Type
Schemes for Hyperbolic Conservation Laws, SIAM Review, Vol. 25, No. 1, 1983, pp. 35
61.

REFERENCES

94

[109] Einfeldt, B., On Godunov-Type Methods for Gas Dynamics, SIAM Journal on Numerical
Analysis, Vol. 25, 1988, pp. 294318.
[110] Toro, E. F., Spruce, M., and Speares, W., Restoration of the Contact Surface in the HLLRiemann solver, Shock Waves, Vol. 4, No. 1, 1994, pp. 2534.
[111] Liou, M.-S. and Steffen, C. J., A New Flux Splitting Scheme, Journal of Computational
Physics, Vol. 107, 1993, pp. 2339.
[112] Liou, M.-S., A Sequel to AUSM: AUSM+ , Journal of Computational Physics, Vol. 129, 1996,
pp. 364382.
[113] Liou, M.-S., A Sequel to AUSM, Part II: AUSM+ -up for all speeds, Journal of Computational
Physics, Vol. 214, 2006, pp. 137170.
[114] Mathur, S. R. and Murthy, J. Y., A Pressure-Based Method for Unstructured Meshes, Numerical Heat Transfer, Vol. 31, 1997, pp. 191215.
[115] Lomax, H., Pulliam, T. H., and Zingg, D., Fundamentals of Computational Fluid Dynamics,
Springer-Verlag, Berlin, 2001.
[116] Hirsch, C., Numerical Computation of Internal and External Flows, Volume 1, Fundamentals of
Numerical Discretization, John Wiley & Sons, Toronto, 1989.
[117] Hirsch, C., Numerical Computation of Internal and External Flows, Volume 2, Computational
Methods for Inviscid and Viscous Flows, John Wiley & Sons, Toronto, 1990.
[118] van Leer, B., Tai, C. H., and Powell, K. G., Design of Optimally-Smoothing Multi-Stage
Schemes for the Euler Equations, Paper 89-1933-CP, AIAA, June 1989.
[119] Carlson, B. G. and Lathrop, K. D., Transport theory - The method of discrete ordinates,
Computing Methods in Reactor Physics, edited by H. Greenspan, C. N. Kelber, and D. Okrent,
Gordon and Breach, London, 1968, pp. 171266.
[120] Gao, X. and Groth, C. P. T., Parallel Adaptive Mesh Refinement Scheme for ThreeDimensional Turbulent Non-Premixed Combustion, Paper 2008-1017, AIAA, January 2008.
[121] McDonald, J. G. and Groth, C. P. T., Numerical modeling of micron-scale flows using the
Gaussian moment closure, Paper 2005-5035, AIAA, June 2005.
[122] Loken, C., Gruner, D., Groer, L., Peltier, R., Bunn, N., Craig, M., Henriques, T., Dempsey, J.,
Yu, C.-H., Chen, J., Dursi, L. J., Chong, J., Northrup, S., Pinto, J., Knecht, N., and Zon, R. V.,
SciNet: Lessons Learned from Building a Power-efficient Top-20 System and Data Centre,
Journal of Physics: Conference Series, Vol. 256, No. 1, 2010, pp. 012026.

REFERENCES

95

[123] Lian, C. and Merkle, C. L., Contrast between steady and time-averaged unsteady combustion simulations, Computers and Fluids, Vol. 44, 2011, pp. 328338.
[124] Oefelein, J. C., Mixing and Combustion of Cryogenic Oxygen-Hydrogen Shear-Coaxial Jet
Flames at Supercrtiical Pressure, Combustion Science and Technology, Vol. 178, No. 13,
2006, pp. 229252.
[125] Masquelet, M. and Menon, S., Large-Eddy Simulation of Flame-Turbulence Interactions in
a Shear Coaxial Injector, Journal of Propulsion and Power, Vol. 26, No. 5, 2010.
[126] Zhang, H. Q., Ga, Y. J., Wang, B., and Wang, X. L., Analysis of combustion instability via
constant volume combustion in a LOX/RP-1 bipropellant liquid rocket engine, Science China
Technological Sciences, Vol. 55, 2012, pp. 10661077.
[127] Gicquel, O., Darabiha, N., and Thvenin, D., Laminar premixed hyrogen/air counterflow
flame simulations using flame prolongation of ILDM with differential diffusion, Proceedings
of the Combustion Institute, Vol. 28, 2000, pp. 19011908.
[128] Vervisch, L., Hauguel, R., Domingo, P., and Rullaud, M., Three facets of turbulent combustion modelling: DNS of premixed V-flame, LES of lifted nonpremixed flame and RANS of
jet-flame, Journal of Turbulence, Vol. 4, 2004, pp. 136.
[129] Domingo, P., Vervisch, L., and Sandra Payet, R. H., DNS of a premixed turbulent V flame and
LES of a ducted flame using a FSD-PDF subgrid scale closure with FPI-tabulated chemistry,
Combustion and Flame, 2005, pp. 566586.
[130] Bellan, J., Supercritical (and subcritical) fluid behavior and modeling: drops, streams, shear
and mixing layers, jets and sprays, Progress in Energy and Combustion Science, Vol. 26, 2000,
pp. 329366.
[131] Oefelein, J. C., General package for evaluation of multicomponent real-gas and liquid mixture states at all pressures, Tech. rep., 2004.

Anda mungkin juga menyukai