You are on page 1of 246

FULLY COUPLED MODEL FOR HIGH-TEMPERATURE ABLATION AND

A REACTIVE-RIEMANN SOLVER FOR ITS SOLUTION

A Dissertation

Presented to

The Graduate Faculty of The University of Akron

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

Nathan Mullenix

May, 2010
FULLY COUPLED MODEL FOR HIGH-TEMPERATURE ABLATION AND

A REACTIVE-RIEMANN SOLVER FOR ITS SOLUTION

Nathan Mullenix

Dissertation

Approved: Accepted:

____________________________ ______________________________
Advisor Department Chair
Dr. Alex Povitsky Dr. Celal Batur

____________________________ ______________________________
Committee Member Dean of the College
Dr. Minel Braun Dr. George Haritos

____________________________ ______________________________
Committee Member Dean of the Graduate School
Dr. Scott Sawyer Dr. George R. Newkome

____________________________ ______________________________
Committee Member Date
Dr. S. I. Hariharan

______________________________
Committee Member
Dr. Gerald W. Young

ii
ABSTRACT

Ablation is a process of rapid material removal from a solid surface by

chemical reactions, sublimation and other erosive processes. Ablation absorbs large

quantities of heat, which makes it a desirable process for Thermal Protection Systems

(TPS) used on aerospace vehicles that encounter severe thermal environments, and has

been in use since the beginning of the Space Age. The ablation process consists of

several coupled sub-processes including gas dynamics, heat, and ablative mechanisms at

the surface.

Experimental techniques can reproduce some but not all of the parameters of the

flight environment, and for this reason numerical modeling has been undertaken to

provide greater understanding. The past state of the art has involved modeling only some

of the sub-processes with simplifications of the others, which tended to incorrectly

predict the thickness of TPS. The computational resources necessary to model all of the

sub-processes in a coupled manner have become available.

A tightly coupled mathematical model for the non-charring ablation problem is

developed within this dissertation. The Reynolds Transport Theorem (RTT) is used to

derive a set of governing equations that takes into account the movement of the ablating

surface and the resulting mass transfer.

A set of numerical methods has been developed and/or modified from existing

forms. For example, finite volume discretization schemes are modified to allow for

iii
arbitrary composition (solid or gas) within a fixed body-fitted computational grid, and a

new method called a reactive-Riemann solver is derived to solve the mass transfer across

the ablating surface and its movement. Finally, serial and parallel algorithms are

developed for the numerical methods.

Individual components of the model and numerical methods are validated against

standard test cases. The full solver is used for ablation problems across a range of free-

stream and surface conditions, and these results are shown to agree with experimental

data. The effect of surface defects on the local ablation rate and process are shown for

the first time: a localized defect can greatly enhance the local ablation rate and create a

region of sublimation dominated ablation even if the rest of the surface is ablating

primarily via oxidation.

iv
DEDICATION

The author dedicates this dissertation to the three important women in his life:

To his wife Irene who has stood by his side throughout the ups and downs of this

research program;

To his mother Rebecca who raised him through life’s triumphs and tribulations;

To his grandmother Jacqueline who taught him everything he knew until he reached

kindergarten.

v
ACKNOWLEDGEMENTS

The author would like to acknowledge the fellowship provided by the Dayton

Area Graduate Studies Institute (DAGSI) and the Air Force Research Laboratory (AFRL)

via their joint AFRL/DAGSI Ohio Student-Faculty Research Fellowship program without

which this dissertation may not have come about, and the sponsors of the research topic

Dr. Datta Gaitonde and Dr. Jean-Luc Cambier. The author would also like to

acknowledge the computational time provided by the Ohio Supercomputer Center (OSC).

vi
TABLE OF CONTENTS

Page

LIST OF TABLES ............................................................................................................. xi

LIST OF FIGURES .......................................................................................................... xii

CHAPTER

I. INTRODUCTION ........................................................................................................... 1

1.1 Literature Review ..................................................................................................... 3

1.1.1 An Overview of the Historical Applications of Ablation in Aerospace ........... 3

1.1.2 Experimental Ablation Studies ......................................................................... 7

1.1.3 Ablation Models ............................................................................................. 15

1.1.4 Numerical Modeling of Ablation .................................................................... 18

1.1.5 Conclusions from the Literature Review ........................................................ 28

1.2 Statement of Problem and Research Objectives ..................................................... 29

II. THEORETICAL MODEL ........................................................................................... 31

2.1 Derivation of Conservation Equations .................................................................... 31

2.1.1 Definition of the Non-Charring Ablator Problem........................................... 31

2.1.2 Development of the Reynolds’ Transport Theorem for the Non-Charring


Ablation Problem ............................................................................................ 32

2.1.3 System of Conservation Equations ................................................................. 35

2.1.4 Equation for Volume Change of Solid ........................................................... 37

vii
2.1.5 Transformation of Surface Integrals ............................................................... 38

2.2 Conservation Equations in General Orthogonal Coordinates ................................. 40

2.2.1 Global and Species Continuity Equations ...................................................... 42

2.2.2 Gas Momentum Equation ............................................................................... 44

2.2.3 Energy Equations ............................................................................................ 48

2.2.4 Solid Volume Equation................................................................................... 49

2.3 Governing Equations in Two-Dimensional Vector Form....................................... 50

2.4 Auxiliary Equations ................................................................................................ 55

2.4.1 Describing the Composition of a Mixture ...................................................... 56

2.4.2 Thermodynamic Relations .............................................................................. 58

2.4.3 Diffusion Velocity .......................................................................................... 59

2.4.4 Transport Properties ........................................................................................ 60

2.4.5 Chemical Reactions in the Gas Phase ............................................................. 62

2.4.6 Surface Chemical Reactions ........................................................................... 64

2.4.7 Surface Energy Balance .................................................................................. 71

2.4.8 Initial and Boundary Conditions ..................................................................... 72

2.5 Summary ................................................................................................................. 73

III. NUMERICAL METHODS ........................................................................................ 74

3.1 Discretization of Equations ..................................................................................... 74

3.1.1 Definition of Finite Volumes .......................................................................... 74

3.1.2 Discrete Form of Governing Equations .......................................................... 78

3.1.3 Time Discretization......................................................................................... 81

3.1.4 Spatial Discretization of Flux Terms .............................................................. 83

viii
3.1.5 Scalar Derivative Discretizations.................................................................... 88

3.2 Riemann Solvers for Inviscid Fluxes ...................................................................... 89

3.2.1 Exact Riemann Solver .................................................................................... 91

3.2.2 Steger-Warming Flux Splitting: An Approximate Riemann Solver ............... 97

3.2.3 MUSCL-Hancock Scheme for Second-Order Fluxes ................................... 103

3.2.4 Exact Reactive-Riemann Solver ................................................................... 105

3.3 Calculation of Remaining Fluxes ......................................................................... 110

3.3.1 Viscous/Heat Transfer Flux .......................................................................... 110

3.3.2 Mass Diffusion Fluxes .................................................................................. 111

3.3.3 Solid and Surface Fluxes ................................................................................... 112

3.4 Calculation of Source Terms ................................................................................ 113

3.4.1 Geometrical Source Terms ........................................................................... 113

3.4.2 Dissipation Terms ......................................................................................... 114

3.4.3 Gas Phase Chemical Reactions ..................................................................... 114

3.5 Time Step Calculation .......................................................................................... 117

3.6 Numerical Boundary Conditions .......................................................................... 118

3.7 Algorithms ............................................................................................................ 120

3.7.1 Serial Algorithm ........................................................................................... 121

3.7.2 Parallel Algorithm......................................................................................... 122

3.8 Summary ............................................................................................................... 125

IV. RESULTS AND DISCUSSION ............................................................................... 128

4.1 Validation.............................................................................................................. 128

4.1.1 One-Dimensional Inviscid Gas Dynamics .................................................... 129

ix
4.1.2 Supersonic Flow over a Wedge .................................................................... 134

4.1.3 Flat Plate Boundary Layer ............................................................................ 136

4.1.4 Driven Cavity................................................................................................ 137

4.1.5 Solid Heat Transfer ....................................................................................... 140

4.2 Preliminary results obtained using a simplified model ......................................... 141

4.3 Test Cases ............................................................................................................. 144

4.3.1 Test Case 1: Sea level flight with defect-free surface .................................. 145

4.3.2 Test Case 2: Sea level flight with defect at TPS surface .............................. 161

4.3.3 Test case 3: Arc-jet conditions in oxidation and nitridation


dominant regime ........................................................................................... 173

4.3.4 Test case 4: arc-jet conditions in sublimation dominant regime .................. 177

4.3.5 Comparison to experimental data ................................................................. 184

4.4 Summary ............................................................................................................... 186

V. CONCLUSIONS ........................................................................................................ 188

REFERENCES ............................................................................................................... 194

APPENDICES ................................................................................................................ 202

APPENDIX A. LIST OF SYMBOLS ......................................................................... 203

APPENDIX B. CONSERVATION EQUATION DERIVATIONS ........................... 208

APPENDIX C. MATERIAL PROPERTIES ............................................................... 216

APPENDIX D. NUMERICAL SCHEMES ................................................................. 225

x
LIST OF TABLES

Table Page

1.1: Constants for Maahs’ mass flux curve fits for various graphites. ............................. 16

2.1: Unknowns in Each System of Equations. .................................................................. 56

4.1: Initial conditions for the Sod Problem. .................................................................... 129

4.2: Freestream conditions for the varying test cases. .................................................... 145

4.3: Geometry of the defect. ........................................................................................... 162

4.4: Data for peak mass fluxes within the defect. ........................................................... 171

4.5: Stagnation point ablative mass fluxes ...................................................................... 185

C.1: Molecular Weights and Enthalpies of Formation ................................................... 216

C.2: Coefficients for Specific Heat Polynomial Curve Fits ............................................ 218

C.3: Reaction Number Designations............................................................................... 220

C.4: Forward Reaction Rate Coefficients. ...................................................................... 220

C.5: Equilibrium Constant Coefficients .......................................................................... 221

C.6: Values of the Collision Integrals ............................................................................. 222

xi
LIST OF FIGURES

Figure Page

1.1: Schematics of charring and non-charring ablators:


(a) charring ablator (b) non-charring ablator. .................................................................. 2

1.2: Schematic of a Typical Arcjet Test Facility. ............................................................... 9

1.3: Scala and Gilbert’s temperature ranges for graphite ablation processes. ................. 10

1.4: Mass flux vs. stagnation point temperature for different types of graphite. ............. 11

1.5: Mass flux vs. Temperature for ATJ graphite. ........................................................... 12

1.6: Shape profiles from a typical PANT run. ................................................................. 13

2.1: Control Volumes for the Non-Charring Ablation Problem. ...................................... 34

2.2: Body-fitted coordinate system for a typical TPS ablation problem.......................... 40

2.3: Comparison of Ytrehus and Hertz-Knudsen Carbon Sublimation Models at 4000K


and 4250K. ..................................................................................................................... 71

3.1: Schematics of a typical TPS ablation problem. (a) Arbitrary ablation problem
(b) Arbitrary grid and associated control volume .......................................................... 76

3.2: Schematic of control volumes associated with point (i, j). ........................................ 78

3.3: Possible location and extent of solid.......................................................................... 83

3.4: Zones for flux calculations. (a) F (b) G. .................................................................... 84

3.5: Riemann Problems. (a) Standard for inviscid gas dynamics


(b) Reactive for non-charring ablation........................................................................... 90

3.6: Comparison of near-surface temperature for 1st and 2nd Order Riemann solvers. .. 105
3.7 Boundary conditions and mesh for a typical problem. ............................................. 119
xii
3.8: Domain decomposition for 8 processors. ................................................................ 123

3.9: Parallel Communication Schemes. (a) Time Step


(b) Sub-domain Boundary Data. .................................................................................. 124

3.10: Speedup and efficiency vs. number of processors. ................................................ 125

4.1: Results of Grid Refinement for Exact Riemann Solver. .......................................... 130

4.2: Results of Grid Refinement for the Steger-Warming Riemann Solver. .................. 131

4.3: Comparison between Riemann solvers. ................................................................... 132

4.4: Comparison between 1st and 2nd order Riemann solvers. ........................................ 133

4.5: Pressure contours for Mach 3 flow over a 30° wedge. ............................................ 135

4.6: Comparison of Riemann solvers for Mach 3 flow over a 30° wedge. ..................... 135

4.7: Velocity Profiles for Flat Plate Boundary Layer. (a) 0.25 m (b) 0.5 m
(c) 0.75 m. .................................................................................................................... 137

4.8: Driven cavity results for Re=5000. (a) Horizontal velocity profiles along
vertical line through geometrical center (b) Velocity vectors. .................................... 139

4.9: Numerical vs. analytical solution for conduction with a constant heat flux in
a semi-infinite solid. .................................................................................................... 141

4.10: Temperature contours of the converged flowfield in the simplified model. ......... 142

4.11: Concentration of C3 at various times within the simplified preliminary model.


(a) 1 ms (b) 5 ms (c) 10 ms. ......................................................................................... 144

4.12: Grid for test cases 1 and 2. ..................................................................................... 146

4.13: Predicted vs. actual shock shape for M∞ = 9.5. ..................................................... 148

4.14: Comparison of temperature between chemically reacting and non-reacting


flow at M∞= 9.5. .......................................................................................................... 148

4.15: Comparison of pressure between chemically reacting and non-reacting


flow at M∞= 9.5. .......................................................................................................... 150

4.16: Comparison of density between chemically reacting and non-reacting


flow at M∞= 9.5. .......................................................................................................... 150

xiii
4.17: Stagnation line profiles for test case 1 with chemically reacting flow without
ablation......................................................................................................................... 151

4.18: Profiles of mass concentration for test case 1 along the stagnation line with
chemically reacting flow and without ablation. ........................................................... 152

4.19: Contours of temperature and O2 concentration for test case 1 with chemically
reacting flow without ablation. .................................................................................... 153

4.20: Contours of temperature and O concentration for test case 1 with chemically
reacting flow without ablation. .................................................................................... 154

4.21: Tangential velocity profiles along lines normal to the surface for test case 1
with chemically reacting flow without ablation. ......................................................... 154

4.22: Temperature profiles along lines normal to the surface for test case 1
with chemically reacting flow without ablation. ......................................................... 155

4.23: O concentration profiles along lines normal to the surface for test case 1
with chemically reacting flow without ablation. ......................................................... 155

4.24: Solid Temperature. (a) Contours. (b) Surface Profile. ........................................... 157

4.25: Ablative mass fluxes for test case 1. (a) Sublimation (b) Oxidation
(c) Nitridation (d) Comparison to overall. ................................................................... 158

4.26: Surface position in local non-dimensional coordinates for test case 1. ................. 160

4.27: Contours of ablation products for test case 1. (a) CO (b) CN (c) C3. .................... 160

4.28: Mass concentration profiles along the stagnation line for test case 1
with ablation. ............................................................................................................... 161

4.29: Original and defect surfaces. ................................................................................ 162

4.30: Temperature contours in the defect pre-ablation. (a) Overall


(b) Upstream Top Corner (c) Downstream Top Corner .............................................. 164

4.31: Velocity vectors in defect region. (a) Overall (b) Upstream top corner
(c) Center region near original surface. ....................................................................... 165

4.32: Concentration of atomic oxygen in the defect region. ........................................... 166

xiv
4.33: Comparison between surface temperatures for test case 1 and 2
prior to ablation. ........................................................................................................... 167

4.34: Comparison between ablative mass fluxes for original and defect surfaces.
(a) Overall (b) Defect region ....................................................................................... 168

4.35: Peak mass fluxes within the defect. (a) Top upstream corner
(b) Downstream side (c) Bottom. ................................................................................ 172

4.36: Grid for test cases 3 and 4. ..................................................................................... 174

4.37: Temperature contour comparison between test case 1 and 3................................. 174

4.38: Profiles of species mass concentration and temperature along the


stagnation line for test case 3. ...................................................................................... 175

4.39: Convergence of ablative mass flux for test case 3. ............................................... 176

4.40: Ablative mass fluxes for test case 3. ...................................................................... 177

4.41: Profiles of species mass concentration and temperature along the


stagnation line for test case 4 using Hertz-Knudsen. ................................................... 178

4.42: Ablative mass fluxes for test case 4 with Hertz-Knudsen sublimation. ................ 180

4.43: Ablative mass fluxes for test case 4 with Ytrheus sublimation. ............................ 181

4.44: Comparison between Hertz-Knudsen and Ytrehus models (a) mass fluxes
(b) injection velocities. ................................................................................................ 181

4.45: Comparison of wall conditions for ablative and non-ablative test case 4.
(a) Shear stress (b) Velocity profiles (c) Heat flux (d) Temperature Profiles. ............ 183

4.46: Comparison between computational results and the experiments of


Lundell and Dickey. ..................................................................................................... 185

xv
CHAPTER I

INTRODUCTION

Space vehicles and high speed aircraft are exposed to a severe heating

environment. An important aspect of their design is the choice of a thermal protection

system (TPS), which serves to keep the underlying structure and payload at an acceptably

low temperature. Several distinct types of TPS have been considered and/or

implemented, which either store this thermal energy, removal of it through some process,

or some combination of the too. One of the earliest and most common forms of TPS is

an ablative heat shield.

Ablation is the process of removing material from a surface by vaporization,

chemical reaction or other erosive process. In the context of hypersonic and/or high

temperature gas dynamics, this removal of material from the surface also absorbs heat

which otherwise would increase the temperature of the underlying structure of the

aerospace vehicle. Ablators generally fall into two categories: charring and non-

charring.

In a charring ablator, the virgin TPS material undergoes a chemical reaction,

which transforms the solid into a char, and releases a gas, which flows through pores

within the char and interacts with the boundary layer, cooling it; this process is know as

pyrolysis. The char, which serves as an insulator, undergoes chemical reactions and/or

1
sublimation at its surface. Figure 1.1.a shows a schematic of the ablation process for a

charring ablator.

In a non-charring ablator, the TPS material does not undergo pyrolysis. Instead

the only reactions that occur are on the surface of the material. Again material can be lost

due to chemical reactions with the surrounding gas, or sublimation of the solid into gas.

Figure 1.1.b shows a schematic of the ablation process for a non-charring ablator.

Heat flux Heat flux


from surrounding Released gas from surrounding Released gas
gas Decrease
cools / disrupts the gas Decrease
cools / disrupts the
in heat
flux
boundary layer in heat
flux
boundary layer

Heat Absorbed/Released by Chemical Reactions /Sublimation Heat Absorbed/Released by Chemical Reactions /Sublimation

Char and pyrolysis gas


temperature rises absorbing heat
Temperature of TPS rises to absorb
Virgin TPS chars and releases gas remainder of heat flux
absorbing heat
TPS
Temperature of virgin
TPS TPS rises to absorb
remainder of heat flux

(a) (b)
Figure 1.1: Schematics of charring and non-charring ablators: (a) charring ablator
(b) non-charring ablator.

Although each type of ablator has its own specific phenomena that must be

considered, there are general phenomena that must be considered regardless of the type of

ablator. Gnoffo1 provides an overview of some of these phenomena:

The combination of convective and radiative heating states are often sufficient to
cause the vehicle’s TPS to degrade through pyrolysis or ablation. The ablation tracks
a heating pulse that generally peaks some time prior to peak dynamic pressure.
Ablation gases react with the atmospheric gases processed by the bow shock. They
may also serve to partially absorb radiation directed toward the vehicle surface.
Furthermore, in the case of massive ablation the vehicle has aerodynamics that are
controlled by a pre-ablative shape early in the trajectory, active shape change during
the heating pulse, and a post ablative shape later in the trajectory. Atmospheric
reconstruction (simulating the flowfield surrounding the vehicle) in these
2
circumstances requires the temporal evolution of the shape change and associated
aerodynamics to be well understood.

This introductory chapter will be divided into several sections and subsections.

Section 1.1 provides a review of the available literature on ablation. Section 1.2

identifies the problem that will be explored by this dissertation and outlines the approach

to its solution.

1.1 Literature Review

A review of the available literature on the topic of ablation has been undertaken.

The literature can be organized into several loose groupings: a historical overview of the

use of ablation, experimental studies, and analytical and computational studies.

1.1.1 An Overview of the Historical Applications of Ablation in

Aerospace

In this subsection, several of the programs that have relied on ablation for TPS

will be highlighted. For a more detailed overview, Heppenheimer2 provides a historical

context for ablation research as it relates to the development of high speed aircraft and

space craft.

In the 1950’s, research was being conducted into the best way to protect weapons

payloads being carried by ballistic missiles. Several different methods of thermal

protection were considered, including a heat sink where a thick metal covering absorbed

the heat flux, a hot structure method which used shingles of insulating material that

radiated away heat, and transpiration cooling technique in which an onboard supply of

3
water would boil from the heat transfer, and be forced through the porous surface by the

resulting pressure, thickening and cooling the boundary layer. All three of these

techniques had difficulties ranging from weight (heat sink), to hotspots (transpiration).

An alternative method was suggested where fiber-reinforced plastics would char, and

eventually vaporize. This technique, deemed ablation, would capture heat in the increase

of the temperature of the virgin and charred material, and also in the vaporization of the

material from the surface, with the vaporized material serving to reduce the heat flux

through interaction with the boundary layer. In addition to fiber-reinforced plastics,

pyrolitic graphite (formed from chemical vapor disposition and having a low thermal

conductivity) was also being investigated. The original test methods involved placing

subscale models in the exhaust of specially designed rocket motors, but methods

involving arcjets and shocktubes were developed.3

Before proceeding it is interesting to point out that Sutton3 developed an informal

requirement that the TPS should be able to withstand a wrench being dropped on it.

Given the demise of the Columbia orbiter, this statement proved to be very prophetic

over 40 years later.

In the late 1950’s and early 1960’s NASA’s research focused on the development

of TPS for Mercury, Gemini, and Apollo (see for example 4, 5, 6 respectively) . The first

two Mercury flights used a beryllium heat sink for TPS, but the switch was made to

fiberglass reinforced resins. The increasing size and the higher speed reentry for the

Apollo capsule led to different ablative materials being used for the TPS. 3

Also in the late 1950’s and early 1960’s, the Air Force was designing and flight

testing the X-15 rocket plane. The X-15 was designed for suborbital flight at a maximum

4
speed of Mach 7, and used a nickel alloy as its skin which provided a heatsink TPS 2, 7. A

follow on project, dubbed X-15-2, was to fly faster and required the use of an ablative

TPS 8.

At the same time that NASA was developing the Gemini spacecraft, the Air Force

was designing the Dyna-Soar, also known as the X-20. This program consisted of a

small winged vehicle similar to the space shuttle that was to be launched on a Titan III

booster rocket. The TPS consisted of a reusable system very similar in concept to that

which would be used on the later space shuttle. This program was canceled in 1963 due

to political reasons, and no flight tests were ever conducted. 2, 9

Even with no actual space-flight program of its own, the Air Force continued to

conduct tests of lifting body reentry vehicles. In the early 1960’s a program known as

ASSET was conducted to test the material behavior of advanced TPS materials, but still

used a small amount of ablative Teflon on its leading edges to provide thermal protection

during the boost phase 10. In 1967 and 1968 a follow on program called PRIME was

conducted, which used a simple ablative TPS, because it was more concerned with

aerodynamic behavior of the lifting body design; one of its conclusions was that it was

possible to use a charring-ablator TPS covering without degradation to the aerodynamic

contours of the vehicle if properly designed11.

Although an ablative TPS was originally considered for the space shuttle (see for

example 12, 13) its inability to be reused was one of the factors that led to the non-ablative

tiles that re-radiated heat and insulated the structure being chosen. Even with this

discontinuation of use for manned vehicles, ablative TPS were still being considered for

unmanned missions14.

5
In the time frame of the 1980’s many aerospace vehicles were being designed or

put into service15. Most of the focus was on the beginnings of flight operations for the

Space Shuttle. One of the interesting programs was the development of so-called

aeroassisted orbital transfer vehicle, which was proposed to interface with other vehicles,

and help them change orbits by dipping into the upper reaches of the atmosphere. One of

the problems involved was that the TPS would be exposed to vacuum for long periods of

time, and have to survive numerous thermal cycles without begin serviced. Also under

development was the proposed National Aerospace Plane16, which had several possible

missions, such as being a follow on vehicle for the space shuttle, and providing high-

speed suborbital point-to-point transport. The latter mission presented challenges

because the TPS would be exposed to long duration flight at high speed through thicker

portions of the lower atmosphere, which made for a much more severe heating

environment. Ablative TPS materials were considered for these vehicles, but at the time

reusable non-ablative heatshields were in favor.

The 1990’s saw the development to various stages of numerous aerospace

vehicles that required TPS for reentry and/or high speed flight.15 There were several

proposed space shuttle replacement/follow on programs, such as the X-33 17, X-34

(which had an ablatively cooled rocket nozzle18), DC-X 19, and OSP (orbital spaceplane).

Also under development was an emergency crew return vehicle for the space station.

While none of these manned vehicles ever flew, and only summarily considered ablative

TPS in the design process, several unmanned probes were being developed with ablative

TPS such as the Mars Exploration Rovers20, and the Stardust Probe 21, 22.

6
In recent time, interest has grown within the Air Force and NASA in hypersonic

propulsion, particularly scramjets. In the environment necessary for supersonic

combustion, there are thermal protection issues within the motor (i.e. flameholders), and

for the airframe (especially the inlet region). Some test programs that have been

conducted include the X-43A program, Hyshot, and the Hyper-X program, which have

all been aimed at gaining flight data on scramjets. Future vehicles that use scramjet

propulsion will fly in denser areas of the atmosphere where TPS conditions will be more

extreme. 2

Finally, the current decade has seen a reemergence of ablative TPS technology for

manned flight with the development of the Constellation Program23. The reentry

trajectories from destinations beyond low-earth orbit require more thermal protection

than the current state of-the art non-ablative heat-shields can handle.

1.1.2 Experimental Ablation Studies

Throughout the fifty-plus year history of using ablative TPS, experimental studies

have been conducted to gain understanding of the ablation phenomena, and to verify the

predictions of models. Other studies have been conducted to explore the effects of

ablation on other aspects of hypersonic flight, such as transition to turbulence, or on the

aerodynamic characteristics of the vehicle. An overview of the various experimental

techniques and studies will be provided below.

As noted by Sutton 3, the original method for investigating ablation was to place a

sample of material within a rocket exhaust plume. This provided a high temperature, but

7
did not provide the velocity or chemical composition that would be encountered in the

real flight environment.

In 1959, two test flights were conducted using a system called RVX-1 3, which

tested silica and glass reinforced resin ablators. The entry vehicle was launched on a

ballistic trajectory.

Rashis and O’Hare 24 presented the results of flight study of a Teflon heat shield

on a higher-speed reentry vehicle. This consisted of an instrumented upper stage of a

Scout launch vehicle which fired after the vehicle had begun its descent to increase the

reentry speed. This test attempted to get real time ablation rate data, but the ionized air

interfered with the operation of the ablation sensor.

According to Walberg and Sullivan14, during the late 1950’s many other

experimental techniques were investigated, including Oxyacetylene torches, combustion

heated wind tunnels, plasma torches, and the arc-heated wind tunnel, which by 1962 was

established as the most capable experimental facility. These arc-jets are still only able to

replicate some aspects of the reentry flight conditions, namely enthalpy and chemistry.

Wick25 provides a detailed explanation of the test procedure for using arc jets to

determine ablation characteristics. Figure 1.2 contains a schematic of the arc jet facility.

A gas is heated by an electric arc and then discharged through a converging-diverging

nozzle, providing a hot supersonic jet. The test model is placed in this jet, as well as

various other instruments, including calorimeters to determine the heat transfer rate.

8
Figure 1.2: Schematic of a Typical Arcjet Test Facility. 25

For the particular aims of this study, it will be useful to narrow the scope of a

review of experimental ablation studies to those done on graphite, which is a non-

charring ablator. Graphite ablates in air by oxidation (chemically reacting with either

molecular or atomic oxygen), nitridation (chemically reacting with atomic nitrogen), and

sublimation (phase change to gaseous carbon of varying species). Scala and Gilbert 26

present a model of graphite ablation and define several temperature ranges across which

different ablation processes dominate, which are pictured in Figure 1.3. They propose

that at low temperatures, the oxidation rate is controlled by a reaction rate equation which

is a function of temperature. At moderate temperatures, this rate becomes limited by the

amount of oxygen that diffuses to the surface. At higher temperatures (depending on the

pressure), the sublimation becomes more important until eventually the mass flux is too

large to allow oxygen or nitrogen to diffuse to the surface. Although not pictured

nitridation becomes important in a range overlapping the diffusion controlled oxidation

and sublimation regimes.


9
Figure 1.3: Scala and Gilbert’s temperature ranges for graphite ablation processes. 26

Metzger, Engel, and Diaconis27 provide details of graphite ablation tests done in

arcjet facilities. Several difficulties were noted, namely adequately reaching flight

conditions (high enough total enthalpy, but subsonic flow), measurement methods to get

rates, and non-continuum flow. The temperatures fall within a range of ~1500 to 2500 K,

which is below the range where sublimation becomes important.

Maahs conducted a study28 in 1970 on graphite ablation that compared four

different types of graphite, and compared there response to temperatures between 1600

and 3450K (the latter of which is in the range where sublimation begins to be important).

Mach numbers of the freestream flow ranged from 2 to 9. An important result from this

study is that oxidation of carbon is not diffusion limited within the range of considered

conditions, which contradicted earlier theoretical work.

Maahs conducted another experimental study29 in 1972, in which 45 different

commercially available graphites were used. Figure 1.4 shows that when exposed to the

same test conditions (M∞ = 2.0, h0 = 2.17 MJ/kg, p0 = 5.6 atm), identically sized and
10
shaped models reach different temperatures and ablate at varying rates. The study

attempted to correlate 10 independent material properties (several different densities, size

of pores, particle size, purity, and thermal conductivity) with ablation performance in an

attempt to determine which graphite was suitable for a given application. One major

observation of the study was that graphites with grain sizes above 0.22mm ablated with

very rough surfaces. Once every graphite with grains larger than this were eliminated,

the ablation performance was correlated to density, purity (ash content), and pore size,

with density being the most important to limiting recession rates.

0.05

0.045
time averaged mass flux (g/(cm s))
2

0.04

0.035

0.03

0.025

0.02

0.015

0.01

0.005

0
2000 2050 2100 2150 2200 2250 2300 2350
Temperature (K)

Figure 1.4: Mass flux vs. stagnation point temperature for different types of graphite.
Data from 29

Lundell and Dickey30, performed a study of the ablation of a specific type of

graphite (ATJ), in a temperature range of 2570 to 4030K. These data cover oxidation,

nitridation and sublimation temperature ranges. Figure 1.5 shows the time averaged mass

11
flux rate as a function of surface temperature, for specimens of varying size and shape

and test conditions. They attempted to correlate surface temperature, effective nose

radius and pressure with ablation rates, and then to compare the results with several

available thermochemical ablation theories. They found that the available theories either

over- or under- predicted the mass flux at a given temperature, especially in the

sublimation range over 3500K.

0.25

1.52 cm dia blunt


time averaged mass flux (g/(cm2s))

1.52 cm dia hemisphere


0.2
3.04 cm dia blunt
3.04 cm dia hemisphere

0.15

0.1

0.05

0
2500 3000 3500 4000
Temperature (K)

Figure 1.5: Mass flux vs. Temperature for ATJ graphite.


Data from 30

In the early 1970’s an intensive research program called the Passive Nosetip

Technology Program (PANT) was being conducted by the Acurex Corporation on behalf

of the Air Force. This was an attempt to explore shape change, transition to turbulence,

and the effects of surface roughness on ablation characteristics. Figure 1.6 shows the

shape change of a sample with the associated change in the bow shock from a typical test.
12
The final goal of the program was to develop computational tools for reentry vehicle

nosetip design. 31

Figure 1.6: Shape profiles from a typical PANT run. 31

A more recent study was performed by Chen et al.32, in conjunction with an

attempt to verify the predictions of a computational solver. They provide data for a Poco

graphite model that is a 10° half angle spherically-tipped cone with a nose radius of 1.905

cm. Two test series were conducted, with one being in the sublimation temperature

range, and the other in a temperature range where only oxidation would occur.

In addition to the above graphite studies, an interesting study on the effect of

differing atmospheres on Teflon ablation was conducted by Marvin and Pope 33. They
13
used air, carbon dioxide, molecular hydrogen, molecular nitrogen, and argon in varying

combinations for arcjet testing. The experiments showed that the presence of argon in

either a pure carbon dioxide, or mixed carbon dioxide and molecular nitrogen atmosphere

increased the heat transfer to the surface, resulting in a higher ablation rate.

In addition to experimental studies that were concerned with determining ablation

properties of specific materials, studies have been conducted in order to determine the

effect of ablation on other aspects of a flight vehicles performance.

Wilkins and Chapman34 conducted a study of the effects of ablation on transition

to turbulence. They used small scale models launched on a ballistics range. It should be

noted that these were plastic models, and that they were sharp tipped as opposed to most

reentry vehicles which are blunt tipped. They show that the transition Reynolds number

decreases with ablation as the cone angle increases; ablation effects the transition more

on larger cone-angle bodies, which could be associated with lower Mach number at the

boundary layer edge.

Kruse35 conducted an experimental study on the effects of ablation on

aerodynamics. Identical models made of an ablating plastic or a non-ablating metal, were

fired at hypersonic speed at a ballistics range. He found that ablation greatly increased

the dynamic instability, reduced drag by reducing skin friction (interrupting boundary

layer), and induced roll.

Another experimental study by Kruse 36, studies the effect of ablation on an

axisymmetric body entering an atmosphere at an angle of attack. In some cases, this will

lead to one region of the otherwise symmetric vehicle being exposed to higher

temperatures and thus higher ablation rates. The experiment consisted of an originally

14
axisymmetric model being machined to simulate an asymmetric ablated surface, which

was then fired on a ballistics range. The experiment concluded that the model would

have a deviation from its original flight attitude, but that this deviation tended to be stable

and would not lead to catastrophic instability.

In general, experimental data provides a good understanding of ablation

phenomena. The most common experimental test method is to place a subscale model in

a heated gas jet (arcjet), which can only reproduce some aspects of the actual flight

environment. The experimental data usually consists of an averaged recession or mass

loss rate at the stagnation point only (Chen et al.32 provide data at 45° from the original

stagnation point), as well as an averaged heat transfer rate from a similar non-ablating

model, with pressures and temperatures at various points occasionally being provided. It

should also be noted that these data are usually a time average over an extended run at a

given test condition, as opposed to the constantly changing environment encountered

along a real flight trajectory.

1.1.3 Ablation Models

Throughout the course of ablation research, there have been three main types of

models that have been used to predict ablation rates, including the Q* model, empirical

models based on curve fits for particular materials, and a thermochemical ablation model.

The Q* model was first proposed by Sutton 3. In this model, a heat of ablation

(Q*) is determined for a material through some experimental correlation, and usually

there is also an associated ablation temperature determined. Material is removed from

the surface once it has reached the necessary temperature, and the necessary energy has

15
been added. This model is still in use, and was one of the methods considered in the

author’s master’s thesis 37. This model can be used to give a quick estimate of the

amount of ablated material, but does not lend itself to a detailed analysis. Laub 38, states

that this idea came out of Teflon ablation, and the fact that it occurs only over a narrow

temperature range. He also states that it is important to realize that the Q* is a correlation

from a steady state ablation experiment, and should not be used for transient ablation

analysis.

Another technique that has been employed is to take arcjet experimental data, and

find a correlating function for it. For example, Maahs 28 provides a function for graphite:

B
  Ae
m T C
pstag ,

where A, B, C are constants listed in Table 1.1, T is the surface temperature, and pstag is

the stagnation point pressure. The coefficients vary with the specific graphite selected,

and have an average residual percentage in the range of 12-40%. It should be noted that

these curve fits are only correlated to the mass loss at the stagnation point of a

hemisphere, and would not be valid for a different shape, or position.

Table 1.1: Constants for Maahs’ mass flux curve fits for various graphites. 28
Material A B C
LMSC glass like Carbon 0.227 -7045 0.458
Vitreous Carbon 0.176 -6360 0.453
Pyrolitic Graphite 0.292 -8505 0.231
ATJ slow 0.059 -3050 0.411
ATJ fast 0.217 -6745 0.761

The third choice of ablation model is thermochemical ablation. In this model, the

actual chemical reactions that occur in ablation and their associated chemical reaction

rate are considered. The specific chemical reactions that needed to be considered differ

16
by the ablator material, and also by the amount of accuracy desired. One of the first

models published for graphite was by Scala and Gilbert 26, which included equilibrium

reactions for rate controlled oxidation, diffusion controlled oxidation, nitridation, and

sublimation of C3. Park39, and Park, Jaffe and Patridge40 present parameters necessary

for the solution of non-equilibrium chemical reaction-rates, and transport properties, and

include the species that need to be considered in calculating ablation of a charring ablator

of the carbon-phenolic type. Zhluktov and Abe41 present a chemical kinetic model for

carbon ablation that involves the intermediate reactions of absorption of gas phase

particles to the solid, and then desorption of the resulting molecule. Havstad and

Ferencz42 provide a review of some of the available surface ablation models of graphite,

and compare the results. Milos and Rasky43 reviewed the models available for predicting

the surface thermochemistry involved with TPS modeling, and noted that the models

were in advance state of development, but they were not being implemented by the

developers of computational simulations of ablation.

The above thermochemical ablation models generally incorporate a sublimation

rate that is a function of a difference between the equilibrium and actual pressure or

concentration (depending on the model) of the sublimation product present, a

thermodynamic speed, and a constant rate coefficient. A different type of rate equation

exists that is based on modeling the behavior of the subliming material over the Knudsen

layer, which has the advantage of incorporating a limit on the injection velocity of

subliming material intrinsically. Ytrehus 44, 45 provided the original development of this

model, and recently Pekker, Keidar and Cambier46 have attempted to improve the model

17
to include thermal conductivity across the layer. A comparison between the different

sublimation rates will be attempted as part of this study.

There is some controversy as to whether the surface chemical reactions of

oxidation and nitridation should be considered as equilibrium or non-equilibrium.

According to 40, “In past analyses, gas-surface reactions were assumed to be in

equilibrium. In recent years, it has become apparent … that the flow over an entry vehicle

is not in thermochemical equilibrium.” Non-equilibrium surface chemical reactions were

incorporated in the current study.

1.1.4 Numerical Modeling of Ablation

Experimental modeling of ablation is not sufficient for design purposes.

According to Gnoffo et al.15 , “Aerothermodynamic data fully representative of the flight

environment are not available from ground based experimental facilities. Thus, the high-

fidelity numerical simulation techniques … are employed.” Ablation consists of many

coupled processes, but until recently, the numerical modeling of ablation phenomena

could be generally be divided into two independent camps 43, namely (1) computational

heat transfer in the TPS and (2) computational fluid dynamics (CFD). An overview of

studies within these categories is presented below, as well as studies that contain attempts

at coupling them.

1.1.4.1 Computational Heat Transfer in the TPS

A review of recent studies done on computational heat transfer with ablation is

provided to show the current state of the art. A heat flux from the gas is assumed,

18
provided by a rudimentary calculation, or from an existing CFD simulation. Either a Q*

or thermochemical model is used to calculate the ablation rate. Several methods are used

including finite-difference/volume schemes, finite element analysis, or the Heat Balance

Integral equation method.

Hogan, Blackwell and Cochran 47 studied the surface recession of a solid during

ablation. They strongly couple the ablation and heat transfer portions of the problem, but

do not account for change in gas dynamics, and use a steady-state boundary layer

prediction for the heat transfer to the surface provided by the gas. They move the mesh

by assuming the ablating material acts like a fictitious elastic material. Two different

models are used for simulating ablation namely a thermochemical model and the

Q* model. The Q* ablation model is used to compare the developed code to the Stefan

problem, which is a moving boundary heat transfer problem. The thermochemical model

is applied to a spherical-tipped cone. They note a 15% change in the overall length of

there model and a blunting of the nose, which would change the surface boundary

conditions.

Storti48 created a numerical model to solve ablation problems using a moving

boundary, and treating ablation as Q*. The ablated material is treated as a fictitious 2nd

phase. This had a very simple boundary condition of a steady heat flux.

The extension of the Heat Balance Integral (HBI) method to problems involving

more realistic ablation was described by Potts49. It requires the assumption of chemical

equilibrium and some knowledge or assumption of the incoming heat flux.

Chen and Milos50 developed a new material solver for charring ablators to

incorporate with their loosely coupled ablation solution framework called FIAT. This

19
was an attempt to use an implicit time scheme for ablation, whereas the industry standard

CMA uses explicit time steps. There was a tendency for the overall solution to be very

sensitive to numerical instabilities caused by CMA. This is a finite-volume code that

requires the heat fluxes from the surrounding gases and radiation as inputs.

Shih et al. 51 studied charring ablators that form a melt layer, using a deformable

control volume. They use a Q* model for ablation. This study accounts for the pyrolysis

gas flow through the char zone.

Dec and Braun52 provide another numerical ablation simulation that uses a

Q*model. It is a one-dimensional finite volume code on a uniform grid applied at the

stagnation point. The material considered in the study was a cork-silicon TPS that was

used for the Mars Pathfinder mission.

Amar, Blackwell, and Edwards 53 use a deforming control volume to track surface

recession in a finite-element scheme. They use both the Q* and thermochemical ablation

models. This study was one-dimensional.

Mitchell and Myers54 present a simple one-dimensional ablation solver that tracks

the surface of a material using Q* ablation and a constant surface heat flux.

1.1.4.2 Computational Fluid Dynamics (CFD) Ablation Studies

A review of recent CFD studies with applications to ablation shows varying levels

of complexity being employed including boundary layer solutions, viscous shock layer

(VSL) equations, and the full Navier-Stokes equations. In general, most codes use over-

simplified boundary conditions that do not take into account spatial variations in surface

temperature or heat flux, or realistic surface mass or energy balances 43.

20
Gnoffo 1 presents a review of the gas dynamic issues involved with planetary

entries, and the use of CFD as a design tool. An overview of the use of CFD and other

predictive methods for several recent missions, including Mars Global Surveyor, Mars

Pathfinder, Galileo, and Stardust is presented. In the case of the Galileo probe, the state

of the art computations overestimated surface recession by between 31% and 96% at the

stagnation point, and underestimated it by 18% to 43% in a different location.

Zhluktov and Abe, use VSL calculations to find the ablation performance around

the MUSES-C capsule. The boundary condition for the solid surface uses a simplifying

assumption that the heat flux is quasi-stationary; it is approximately equal to the mass

flux times the specific heat and wall temperature. This tends to over-predict the wall

temperature.

Engel, Farmer and Pike 55 considered equilibrium flow with a Q* ablation model,

and no conduction into the solid. The study was mainly concerned with radiation effects.

The solver was one-dimensional, and a solution was only found along the stagnation line.

Komurasaki and Candler56 determined that strong ablation can cause a transition

to turbulence much earlier than expected using a two-equation (k-) model and the

RANS. They neglect chemical reactions because their influence on transition is still not

completely modeled, which significantly overestimate the peak temperature. They have

an assumed mass flux for the ablation products.

Zhong et al. 57 used DSMC to study the gas surrounding the ablating TPS of the

Stardust mission. They assumed a constant ablation rate of 0.016 kg/m2s, and obtain a

wall temperature only for a boundary condition.

21
Barbante, Degrez and Sarma58, developed a simulation code for the boundary

layer near a non-ablating TPS. They included chemical reactions in the gas, and at the

catalytic surface. The wall temperature and the conditions at the edge of the boundary

layer must be specified.

Ivankov 59 studied radiative heating of the gas near the stagnation region of a

blunt body, and its effect on a Q* model of ablation, without taking into account surface

recession. An injection velocity is specified as a function of wall temperature, radiative

flux to the surface, and material properties; no limit is place on this velocity.

Candler and MacCormack 60, present a mathematical model and numerical solver

for chemically reacting flow that could be used in conjunction with an ablation solver if

given the proper boundary conditions. They present only non-ablating wall boundary

conditions, but they provide comparisons for the chemical flow field around a spherical-

tipped cone.

Several codes that are in use for hypersonic flow simulations are described in 15.

LAURA is a solver that incorporates the Roe scheme for inviscid flux calculations at 2nd

order of approximation away from discontinuities, contains various aerothermodynamics

models, and has an adaptive grid tool that is used near shocks and surfaces. GASP is a

program that contains more aerothermodynamic data, and uses either Roe or Van Leer

schemes to calculate the inviscid fluxes. GIANTS is a laminar code that allow for

thermal-chemical non-equilibrium. Felisa_Hyp is a second order accuracy code that has

its own grid generator.

The codes listed above can be used in an ablation simulation by being provided

proper boundary conditions. A description of the implementation of such a boundary

22
condition into LAURA is provided by Thompson and Gnoffo61, where the boundary

condition that was implemented required the solution of three equations:

p gas   gasvgas
2
 pinj  injvinj
2

m~  v , (1.1.1)
abl inj inj

pinj  pinj inj,Tinj 

which is solved for the unknown values of pinj, inj and vinj when a mass flux and wall

temperature are provided. The first equation represents conservation of momentum

between the conditions at the wall, and the known conditions at some point near the wall,

the second equation is from the definition of mass flux, and the third equation provides

the pressure in terms of the known injection pressure and the unknown injection density

(usually the ideal gas law).

1.1.4.3 Coupled Studies

In addition to studies that only concentrate on one portion of the problem, such as

gas dynamics, conductive heat transfer or the actual ablation reactions, there have been

attempts to look at the entire ablation problem, especially in recent years. In order to

better understand the coupled studies, it is useful to propose a definition of the degree of

coupling as either loose or tight. In a loosely coupled scheme, two separate solvers are

used for CFD and computational heat transfer, with a solution being found in each solver,

and then data is exchanged through boundary conditions, and then the solvers are iterated

again with the new boundary condition held as a constant, with this process continuing

until some convergence criteria is met. In a tightly coupled solver, the boundary

conditions are updated with any changes in either solver. According to Gnoffo 1, “One

23
of the most challenging aspects of TPS design is the proper coupling of the material

response with the external flow solution.”

Conti, et al. 62 assessed the current state of the art:

Current methods used to predict the aerodynamics and heating of reentry vehicles
concentrate attention on some aspects of the problem at the expense of others, partly
due to historical reasons in the development of these methods, and partly to the
limitations of the computing equipment.

As an example of why they thought a coupled solution was needed they stated:

.. the aerodynamic heat flux to the vehicle depends to the first order on its surface
temperature, which is the resultant of a surface energy balance that depends on the
aerodynamic heat flux itself. Thus, the local surface temperature seeks a value that
will satisfy simultaneously the aerodynamic heating and material response processes.
It is therefore desirable to eliminate the artificial separation of these interacting
phenomena by developing a unified model that takes full account of flowfield
features as well as response of the heat shield material, within a single computational
tool.

They note that the in the past that this was looked at as too ambitious given the state of

computer technology. They believed that theirs was the first attempt at a coupled

solution. The gas dynamics portion of their code used the MacCormack flux vector

splitting scheme, and modeled two species, equilibrium air and the ablation product C3,

which were not allowed to react. Turbulent modeling was handled by the Baldwin-

Lomax model with “transition derived empirically from experience”. The solid modeling

consisted of heat conduction in the normal direction, with a surface energy balance used

to provide the relevant boundary conditions. The ablation model used was a B’ lookup

table, with assumed equilibrium reactions. The models were tightly coupled with the

solid and gas being computed on separate overlapping grids. Portions of the grid were

either activated or deactivated depending on the location of the surface.

24
Bhutta and Lewis63 presented a loosely coupled ablation solver. For the gas

dynamics they use a parabolized Navier-Stokes solver and use an existing material

response code for Teflon. They do not account for surface recession.

Milos and Rasky43, presented a study with the objective of developing new

interaction frameworks to couple existing codes. They define Computational Surface

Thermochemistry (CST) as “a multi-disciplinary activity to bridge the gap between

existing CFD and CSM methodologies.” In the case of designing the TPS they make the

case for coupling as:

This problem is of critical importance for optimization of the TPS for future
hypersonic vehicles and spacecraft. TPS are traditionally designed with one and two-
dimensional engineering codes. Occasionally a detailed computational solution is
obtained, but these solutions rarely contain the correct CST boundary conditions. To
compensate for the uncertainties in the analyses, a safety margin of extra TPS
material is added to the final design, and the structural weight must also be increased.
The TPS/structural weight is typically significantly larger than the payload
weight, by up to a factor of four. Therefore a reduction in the TPS weight has a
cascade effect: the structural weight is also reduced, resulting in a direct increase in
payload and scientific capability. Clearly there is a need for more accurate, multi-
dimensional computational tools which can be used to reduce the uncertainties in TPS
analysis and to optimize the TPS distribution around hypersonic vehicles and
spacecraft.

In an attempt to solve this problem they developed a loose coupling methodology, with

transient solid modeling, several steady state CFD solutions conducted at various times to

get an estimate of the heat flux to the surface, and boundary conditions to couple the two

different solutions regimes. They proposed surface mass and energy balances that would

be useful for the coupling procedure.

Olynick, Chen and Tauber 64 describe a TPS prediction methodology used on the

Stardust mission that loosely couples a non-equilibrium Navier-Stokes solver with a

material response code and a radiation prediction code. To their knowledge, this was the

25
first time that a flow calculation involving any coupling was used as a primary tool for

heatshield design. The heatshield material was PICA, which is a charring ablator. They

note that the mass injection at the stagnation point causes a change in the standoff

distance of the bow-shock which changed the heating environment around the vehicle.

The surface ablation reduced the surface heat transfer, temperature and shear, but did not

affect the pressure. Also, their model resulted in a 35% drop in the peak heating rate.

The code did not allow for shape change, but the results were used to predict shape

change and its effects in some separate aerodynamic calculations.

Kuntz, Hassan, and Potter 65 developed an iterative structure for loosely coupling

the ablation problem. The iterative technique was necessary because the use of an

explicit time stepping in the loosely coupled system developed severe numerical

oscillations. The analysis was conducted for the MUSES-C capsule.

Suzuki, Furudate, and Sawada66 present a loosely coupled solution to the

aerodynamic heating environment for the MUSES-C capsule, which used a charring

ablator in its TPS. According to them the motivation behind the development of this

technique was:

Ablation phenomena of heatshields involve complicated physics that make prediction


quite difficult. However, a simpler physical model that is suitable for explaining a
specific problem can fail to give an overall description of the problem. Therefore the
only viable model for accurate prediction of ablation phenomena seems to be to
account for the various aspects of the problem simultaneously…

Their coupling technique involved iterating a solution between a CFD code and a

material response code by exchanging information through the respective boundary

conditions. Although one of their listed goals in the development of the simulation was

to account for shape change, they postponed its inclusion for a later version.

26
Czybyk et al67developed a methodology that consisted of loosely coupling gas

dynamics and solid material response because after reviewing the existing design

philosophy, the authors found that, “a comprehensive, integrated, multi-dimensional

methodology is needed to facilitate the detailed evolution of critical phenomena inherent

to ablating gas/solid interfaces.” A steady state CFD solution is found using CFD++, and

then the surface values are provided to the material response code which used

equilibrium chemistry. This work did not handle surface recession, but a stated goal was

to add a grid displacement tool in the future.

Daimon et al. 68 present a loosely coupled ablation model for a different

application. In their study they are trying to determine the pyrolysis of a solid rocket

motor nozzle and its effect on performance. A material response code provides a

blowing correction factor for the CFD boundary conditions, and the solutions are iterated

until converged. In this study shape change is accounted for with the computational grids

being regenerated.

Ayasoufi et al.69 studied ablation of a charring ablator using a loosely coupled

solver. They propose a boundary condition to be used for the coupling, and present

simple test cases.

Gosse and Candler70 developed a tightly coupled solver to simulate the needed

TPS for a novel launch system. This launch system consists of a airplane mounted

electromagnetic launch-rail that accelerates a carrier vehicle to a high velocity in a short

distance, with the carrier maintaining enough speed through the atmosphere to achieve

orbital height were an additional rocket inserts the payload into orbit. This presents a

severe heating environment for the TPS because the high velocities are achieved in the

27
relatively dense portions of the atmosphere, as compared to the normal reentry case

where the highest velocities occur in the rarefied upper atmosphere. The actual coupling

procedure is not detailed. The surface recession rate is used to move the grid points, and

then the other gird points are adapted as needed. An interesting point is that the solution

is integrated with a simple trajectory solver to find the flight path of the vehicle and

provide input to the solver. The material response code is compared to available

experimental data and is shown to under-predict the recession rate by significant margins

(two to three times).

Prakash and Zhong 71, present another attempt at a loosely coupled ablation

solver. Their solver does not account for shape change, but does include non-equilibrium

chemically reacting flow.

1.1.5 Conclusions from the Literature Review

In the process of the literature review several conclusions can be reached. First,

while experimental studies are useful, the existing test facilities do not capture the entire

flight environment. Second, the modeling of the individual processes involved in

ablation is mature, but detailed coupled modeling of the entire process is needed to

improve the accuracy of the predicted TPS performance. Third, research has focused on

defect-free surfaces, while the presence of even small defects may lead to mission failure.

Finally, with the return to ablating TPS for future missions, the development of better

computational tools is not just an academic exercise.

28
1.2 Statement of Problem and Research Objectives

Based on the conclusions drawn from the literature review, a need for a new

solution methodology for the ablation problem was identified. While attempts have been

made to reuse existing computer codes, they have suffered from either inflexibility in the

boundary conditions that are implemented, or form numerical instabilities that are a

product of the interface structure that is being used. Therefore it was the goal of this

research program to develop a solution methodology starting from first principles for the

TPS ablation problem, with care taken to put it in a form that also can be applicable for

other similar problems such as rocket nozzle sublimation or solid fuel combustion. The

considered materials were limited to non-charring ablators, specifically graphite, due to

the complexity and inconsistencies between theories on the pyrolysis process; however,

the surface reactions between charring and non-charring ablators are similar and future

work could extend the developed method to charring problems. Within the scope of this

project several research objectives were identified and met.

First, a comprehensive model for the gas-dynamics, heat transfer, and ablation

processes was developed starting from first principles. Within this overall model,

provisions were made for the inclusion of various sub-models, such as using the Hertz-

Knudsen or the Ytrehus models to determine the sublimation rate. The details of this

model and its development can be found in chapter 2.

Second, numerical techniques to solve the developed model, capable of handling

the movement of the ablating surface, and accommodating varying levels of surface

detail, such as small defects (cavities), were created or modified from existing forms.

New techniques that were developed include a reactive-Riemann solver for ablation

29
fronts. Existing techniques that required modification include various finite-volume

methods, discretization schemes, and parallel algorithms. These numerical techniques

are described in chapter 3. In particular the critical advantage of body-fitted grid versus

staircase grid and second- versus first-order schemes were outlined.

Next the model was validated against test cases of varying complexity. A

discussion of these results appears in chapter 4. Four ablation test cases were conducted

across a wide range of temperatures, pressures, and ablation conditions. Two of the test

cases were at sea level conditions with one of the cases having a small defect in its

surface from which the data on the cavity dynamics and ablation enhancement were

obtained. The other two test cases were at arc-jet conditions of ground experiments with

one having a surface temperature in the oxidation and nitridation dominated temperature

range, and the other being within the sublimation dominated temperature range (see

Figure 1.3).

30
CHAPTER II

THEORETICAL MODEL

In this chapter the theoretical model for a non-charring ablation problem is

developed. Section 2.1 contains the derivation of the conservation equations. In Section

2.2 these conservation equations are transformed to general orthogonal coordinates.

Section 2.3 provides the governing equations in two-dimensional (axisymmetric) vector

form. Section 2.4 details the auxiliary equations necessary for closing the systems

including thermodynamic relations, transport properties, chemical reactions in the gas,

surface chemical reactions, and the necessary boundary conditions.

2.1 Derivation of Conservation Equations

In this section the governing conservation equations are derived. A definition of

the non-charring ablator problem is provided. The Reynolds Transport Theorem is used

to generate a set of conservation equations for this problem.

2.1.1 Definition of the Non-Charring Ablator Problem

A general concept of the non-charring ablation problem can be given as follows.

Consider an arbitrary body of solid material that is immersed within an arbitrary gas.

Now consider that this body of solid material can lose material and deform (change

31
shape) only through some arbitrary reaction that occur at its surface, consistent with the

definition of a non-charring ablator provided in the Introduction. Material leaves the

surface with an injection velocity vector of Vinj, and the surface recedes with a related

velocity vector, Vsurf. The non-charring ablation problem consists of determining the

flowfield of the gas around the arbitrary body, the heat transfer between the gas and solid,

and the determination of the composition, thermodynamic properties, and associated

velocities of the ablated material that is injected from the surface.

2.1.2 Development of the Reynolds’ Transport Theorem for the Non-

Charring Ablation Problem

In order to develop the proper form of the conservation equations for the non-

charring ablation problem, it is useful to have a relation between the conservation laws,

which are provided for a system, and its equivalent for a control-volume. The Reynolds’

Transport Theorem (RTT) relates the time rate of change in an extensive property, N, of a

system to the time rate of change of its associated intensive (does not depend on mass)

property . The relationship between these properties is given by:

N  m , (2.1.1)

where m is the mass of the system. In order to use the RTT, a set of systems and an

associated set of control volumes must be defined.

At some initial time t0, an overall system SO is defined to encompass the arbitrary

solid body and some amount of the fluid that surrounds it. A second system, Ssolid is

defined to encompass only the solid body portion of SO. A third system, Sgas contains

32
only the gas portion of SO. For any given extensive property N, the amount of the

property in each system can be related by:

NO  N gas  N solid , (2.1.2)

where the subscript refers to the particular system. If the time derivative is taken on both

sides of Eq. 2.1.2, a relation between the time rates of changes for each system is found:

dN O dN gas dN solid
  (2.1.3)
dt dt dt

If the time rate of change is known for any two of the systems, the other can be found for

example the time rate of change within Sgas is:

dN gas dN O dN solid
  . (2.1.4)
dt dt dt

Also at the initial time t0 define an arbitrary fixed (with respect to an inertial

frame of reference) control volume, CvO, such that it contains some arbitrary amount of

SO, and all of Ssolid. The control surface of CvO is denoted by CsO and it has a unit

outward normal denoted nO. Next define an arbitrary deformable control volume Cvsolid,

such that it contains all of Ssolid at t0. The associated control surface, Cssolid, coincides

with the initial surface of Ssolid; it has a unit outward normal denoted by nsolid. The region

of fluid contained within CvO but outside of Cvsolid is denoted by Cvgas. Figure 2.1 shows

the relation between these control volumes.

33
V
O
s
nO

C
Cvgas

nsolid V surf

Cssolid
V inj Cvsolid

Figure 2.1: Control Volumes for the Non-Charring Ablation Problem.

For the overall control volume Cv the RTT is given by72:

dN O 
  dV   V  n O dA (2.1.5)
dt t CvO CsO

where V is the velocity vector at any point along Cs0. For an arbitrary deformable

control volume, such as Cvsolid the RTT is given by 72:

dN solid 
t Cv
 dV   Vr ,solid  n solid dA , (2.1.6)
dt solid Cssolid

where Vr,solid is the velocity vector relative to the moving control surface Cssolid, and it is

defined by:

Vr ,solid  Vinj  Vsurf , (2.1.7)

where Vsurf is the velocity vector of the moving control surface. It should be noted that

Vinj and Vsurf are given in the initial inertial frame of reference, and that within the solid

the injection velocity is zero. Using Eq 2.1.4, the RTT for Cvgas can be defined:

34
dN gas  
dt
 
t CvO
dV   V  nO dA   dV   Vr ,solid  n solid dA . (2.1.8)
CsO
t Cvsolid Cssolid

The volume integral over CvO can be written as the sum of the integrals over Cvgas and

Cvsolid:

 dV   dV   dV ,


CvO Cvgas Cvsolid
(2.1.9)

which simplifies Eq. 2.1.8 to its final form:

dN gas 
dt
  dV  Cs V  nO dA  Cs Vr ,solid  n soliddA .
t Cv
(2.1.10)
gas O solid

In Eq. 2.1.6 and 2.1.10, the surface integral over Cssolid is the flux of material resulting

from the surface reaction. It should be noted that in the above equations the time

derivative must be kept outside of the integral, due to the variability in the volume over

which the integral is taken.

2.1.3 System of Conservation Equations

The equations for conservation of mass, momentum and energy are derived in

Appendix B. The global continuity equation for Cvsolid is given by Eq. B.1.2 :


t  dV   V
Cvsolid Cssolid
surf  n solid dA  0 , (2.1.11a)

or Eq. B.1. for Cvgas:


 dV  Cs V  nO dA  Cs  Vinj  Vsurf  n soliddA  0 .
t Cv
(2.1.11b)
gas O solid

The species continuity equation for Cvgas is given by Eq. B.2.5 :

35

  s dV  Cs  s V  Vs  nO dA
t Cv
  w dV , (2.1.12)
  V  Vsurf  Vs  n solid dA
gas O
s
 s inj Cvgas
Cssolid

where w s is the specific mass production in units of mass per unit volume per unit time,

and the vector Vs represents the diffusion velocity of each individual species caused by

gradients in concentration73. The momentum equation for the gas is given by Eq. B.3.7:


 VdV  Cs VV  nO dA
t Cv
 fdV   p  n dA   τ  n dA 0 0


Cvgas CsO CsO
, (2.1.13a)
 VVinj  Vsurf  n soliddA
gas O

 
 p  n solid dA 
Cssolid
 τ  n solid dA
Cssolid
Cssolid

where f is the vector of body forces, and p is the pressure tensor, and the shear tensor

whose components can be written as74:

 2e11  e22  e33 


2
 11 
3
 22   2e22  e11  e33 
2
3
 33   2e33  e22  e11  .
2 (2.1.13b)
3
 12  e12
 13  e13
 23  e23

The energy equation for Cvsolid is given by Eq. B.4. :


t  e dV T  kT  n solid dA


Cvsolid Cssolid
(2.1.14a)
  eT Vsurf  n solid dA   q
Cvsolid
rad dV   q
Cvsolid
gen dV
Cssolid

and for Cvgas by Eq. B.4. :

36
 kT  n
CsO
O dA   kT  n
Cssolid
solid dA


   e    p V  n
N

t  eT dV   eT V  n O dA 


s 1 CsO
s T s s s O dA
Cvgas CsO
 . (2.1.14b)
 
Cssolid
eT Vsurf  n solid dA   q rad dV   q gen dV
Cvgas Cvgas

  p  V  n
CsO
O dA   τ  V  n
CsO
O dA

2.1.4 Equation for Volume Change of Solid

In the developed system of equations the control volume containing the solid can

deform or change in volume. An equation is also necessary for this property. In the RTT

for Cvsolid (Eq. 2.1.6), define N as the mass times the specific volume, v, and  as the

specific volume v:

d mv solid 
dt

t  vdV   vV
Cvsolid Cssolid
surf  n solid dA , (2.1.15)

Expanding the derivative on the left hand side by the chain rule, and taking into account

the identity that density times specific volume is identically unity yields:

 dm dv  
v
 dt
m  
dt  solid t
Cv
1dV 
Cs

Vsurf  n solid dA ,  (2.1.16)
solid solid

By conservation of mass the first term on the left hand side is identically zero, and

assuming that the density (and therefore specific volume) of a solid remains constant the

second term becomes identically zero, which allows the final form of the solid volume

equation to be given as:


t  1dV   V
Cvsolid Cssolid
surf  n solid dA  0 , (2.1.17)

37
2.1.5 Transformation of Surface Integrals

The conservation equations developed above all contain integrals over control

volumes and control surfaces. It is advantageous to convert the surface integrals into

volume integrals, and this can be accomplished by the divergence theorem:

 ψ  ndA     ψdV ,


Cs Cv
(2.1.18)

where  is an arbitrary vector or tensor. In the current model it is important to note that

if the surface integral is taken over CsO, the resulting volume integral is over CvO, which

can be split using Eq. 2.1.9.

The global continuity equations (2.1.11a-b) become:

 dV     V dV  0 ,



surf (2.1.19a)
t
Cvsolid Cvsolid

and:


t  dV     V dV
Cvgas Cvgas
0 (2.1.19b)
  
Cvsolid
  Vsurf dV 

The species continuity equation (2.1.12) is transformed to:


  s dV
t Cv gas

     s V  Vs dV   w dV .


s (2.1.20b)
Cvgas Cvgas

     V dV
Cvsolid
s surf

38
The momentum equation (2.1.13a) becomes:


 VdV
t Cv  fdV
Cvgas
gas

    VV dV      pdV . (2.1.21)


Cvgas Cvgas

    VV dV


Cvsolid
surf     τdV
Cvgas

The energy equations (2.1.14a-b) can be written as:


 e dV
t    kT dV
T


Cvsolid Cvsolid
(2.1.22a)
    e V    q dV   q
T surf rad gen dV
Cvsolid Cvsolid CvII

and:

   kT dV


 Cvgas
 e dV
     e    p V dV
T N
t 
Cvgas s T s s s

   e V dV


s 1 Cvgas
 T  , (2.1.22b)
Cvgas   q rad dV   q gen dV
     eT Vsurf dV  Cvgas Cvgas

Cvsolid     p  V dV   dV


Cvgas Cvgas

where the dissipation function  has been defined as:

    τ  V  . (2.1.22c)

Finally the solid volume equation can be written as:


1dV     Vinj  Vsurf dV .
t Cv
(2.1.23)
solid Cvsolid

39
2.2 Conservation Equations in General Orthogonal Coordinates

Consider that the system of governing equations presented by Eq. 2.1.19-23 is to

be solved on some arbitrary orthogonal curvilinear coordinate system, where an arbitrary

point P has coordinates of (x1, x2, x3). The derivation of the transformation of equations

written for a Cartesian coordinate system (x, y, z), to this new coordinate system follows

Tannehill et al. 74 Figure 2.2 shows a two-dimensional axisymmetric body-fitted

coordinate system for a typical spherical tipped cone geometry, where x1 is tangential to

any point on the surface, x2 is normal, r is the distance to the axis of symmetry, and K is

the curvature.

Figure 2.2: Body-fitted coordinate system for a typical TPS ablation problem.

The transformation from the new coordinate system to Cartesian coordinates is

given by:

x  x x1 , x2 , x3 
y  y x1 , x2 , x3  . (2.2.1)
z  z  x1 , x2 , x3 

40
The Jacobian of this transformation is:

  x, y , z ) 
, (2.2.2)
x1 , x2 , x3 

and if it is nonzero, the transformations from Cartesian to the new coordinate system:

x1  x1 x, y, z 
x2  x2  x , y , z  , (2.2.3)
x3  x3  x, y, z 

can be defined. The metrics of this transformation, h1, h2, h3, are defined as:

2 2 2
 x   y   z 
h1  2
        
 x1   x1   x1 
2 2 2
 x   y   z 
h2 2          . (2.2.4a)
 x2   x2   x2 
2 2 2
 x   y   z 
h3 2
        
 
 3  3  3
x x x

For the body-fitted coordinate system in Figure 2.2 the metrics are:

h1  1  x2 K x1  h2  1 h3  r x1   x2 cos x1  ,


m
(2.2.4b)

where m is identically zero for two-dimensional and m is identically one for two-

dimensional axisymmetric. The gradient and divergence operators in the new coordinate

system are:

1  1  1 
  i1  i2  i3
h1 x1 h2 x2 h3 x3
, (2.2.5)
1     
A   h2 h3 A1   h1h3 A2   h1h2 A3 
h1h2 h3  x1 x2 x3 

where  is an arbitrary scalar, and A is an arbitrary vector.

41
2.2.1 Global and Species Continuity Equations

Using the transformations in Eq. 2.2.5, the global continuity equation for Cvsolid

(2.1.19a) can be written as:

 
 
h2 h3 u1, surf  
 x1 

t  dV  
1  

h1h2 h3  x2

h1h3 u 2, surf  dV  0 , (2.2.6)
Cvsolid Cvsolid  

   h1h2 u3, surf
 x3


which can be rewritten using the chain rule as:


t  dV   h2 h3  
 u1, surf
Cvsolid

 1  x1 
 
u1, surf   1 

 h1h3  
 h1 x1      u 2, surf
x2 
dV , (2.2.7a)
 
 1  u
 h x 
2, surf 
 dV

Cvsolid
h1h2 h3

 u3, surf h1h2  

Cvsolid  2 2   x3 
  1  u3, surf  

 h3 x3 

where the right hand side can be looked at as a geometrical source term. The same

process can be applied to the global continuity equation (2.1.19b) for Cvgas:

42
  h2 h3  
  u1 
x1 
 dV
t Cv 1 

 h1h3  
gas
   u2 dV
 1   hh h  x2 
 u1   Cvgas 1 2 3
  h1h2  
 h1 x1   u3 
 1   x3 
   u2  dV 
 h2 x2   h2 h3  
 1    u1, surf
Cvgas

 u3   x1 
 h3 x3  1   h1h3  
   u2, surf dV
 1   x2 
 u1,surf   hh h
Cvsolid 1 2 3
  h1h2  
 h1 x1   u3, surf 
1   x3 
   
 h2 x2
u2,surf  dV
Cvsolid
 1  
 u3,surf  . (2.2.7b)
 h3 x3 

Similarly, the species continuity equation (2.1.20) becomes:


  s dV
t Cv  w dV
s
gas Cvgas
 1    u1      u1   h2 h3  
  s       s   
  
 h1 x1    u1, s       u1 , s  x1 
 1    u   1   u2   h1h3  
     s  2   dV  
Cvgas 

h2 x2    u2, s   
Cvgas 1 2 3 
hh h
   s  
  u2, s  x2 
 dV
    u3   h1h2  
  1     u3       s
  
 h3 x3  s   u3, s    
   u3, s  x3 
  
 1    h2 h3  
  su1,surf     s u1, surf
x1 

 h1 x1  1 

 h1h3  
1 
  
 h2 x2
 s u2, surf  dV

 
h1h2 h3 
  s u2, surf
x2 
dV
Cvsolid
 1  
Cvsolid
  h1h2  
  su3,surf  

  u
s 3, surf
x3 
 . (2.2.8)
 h3 x3 

43
2.2.2 Gas Momentum Equation

The momentum equation contains some terms that are not contained in Eq. 2.2.5.

Terms such as   VV  can not use the divergence definition above because VV is a

tensor, not a vector. These terms can be expanded as 74:

  VV   V  V  V  V  . (2.2.9)

The first term on the right hand side of Eq. 2.2.9 can be written in the new coordinate

system as:

 u1 u1 u2 u1 u3 u1 


 h x  h x  h x 
 1 1 2 2 3 3
i
 u1u2 h1 u1u3 h1 u22 h2 u32 h3  1
    
 h1h2 x2 h1h3 x3 h1h2 x1 h1h3 x1 
 u1 u2 u2 u2 u3 u2 
 h x  h x  h x 
V  V    1 1 2 2 3 3
i , (2.2.10)
 u1u2 h2 u2u3 h2 u12 h1 u32 h3  2
    
 h1h2 x1 h2 h3 x3 h1h2 x2 h2 h3 x2 
 u1 u3 u2 u3 u3 u3 
 h x  h x  h x 
 i
1 1 2 2 3 3

 u2u3 h3 u1u3 h3 u22 h2 u12 h1  3


    
 h2 h3 x2 h1h3 x1 h2 h3 x3 h1h3 x3 

and the second term can be written using Eq. 2.2.5. The right hand side of the

momentum equation contains terms of the form   Π , where  is a tensor. These terms

can be expanded as 74:

44
 1   
  h2 h311    h1h312    h1h213  
 h1h2 h3  x1 x2 x3  i
  h  h  h  h 1
 12 1
 13 1
 22 2
 33 3

 h1h2 x2 h1h3 x3 h1h2 x1 h1h3 x1 
 1   
  h2 h312    h1h3 22    h1h2 23  
h h h x x2 x3
Π    1 2 3  1  i . (2.2.11)
  h  h  h  h 2
 12 2
 23 2
 11 1
 33 3

 h1h2 x1 h2 h3 x3 h1h2 x2 h2 h3 x2 
 1   
  h2 h313    h1h3 23    h1h2 33  
h h h x x2 x3
 1 2 3  1  i
  h  h  h  h 3
 23 3
 13 3
 22 2
 11 1

 h2 h3 x2 h1h3 x1 h2 h3 x3 h1h3 x3 

The shear stress tensor  contains strains eij, which also have an expanded form in the

new coordinate system:

1 u1 u 2 h1 u3 h1


e11   
h1 x1 h1h2 x2 h1h3 x3
1 u2 u2 h2 u3 h2
e22   
h2 x2 h1h2 x1 h1h3 x3
1 u3 u 2 h3 u h3
e33    1
h3 x3 h2 h3 x2 h1h3 x1
h   u3  h2   u2  . (2.2.12)
e23  3    
h2 x2  h3  h3 x3  h2 
h   u3  h1   u1 
e13  3    
h1 x1  h3  h3 x3  h1 
h   u1  h2   u2 
e12  1    
h2 x2  h1  h1 x1  h2 

Using Eq. 2.2.9, the momentum equation (2.1.21) can be rewritten as:


t  VdV
Cvgas
 fdV
Cvgas

 V  V  V  V dV      pdV .


Cvgas Cvgas
(2.2.13)

  V  V  V   V dV     τdV


surf surf
Cvsolid Cvgas

45
The momentum equation is a vector equation, and can be split into separate

equations for each direction x1, x2, and x3. Applying the appropriate transformations,

utilizing that the velocity is zero within the solid, and ignoring body forces, Eq. 2.2.13

can be written in the x1 direction as:

  u  h 1   
  2  1  h1h3  
  h1h2  x2 h3 x2  
  
u   u3  h1  1  h1h2   
 1  h h  x h2 x3  
  1 3 3 
  u1  h2 h3  
     dV
t VdV Cvgas  

 h1h2 h3 x1 


Cvgas u 2 h2
2

 1   u 2    
  h1h2 x1 
  1  
 
 h1 x1   p   11   
  u 2 h3 
 1   u u    3
  
      dV  h h x 
1 2
h2 x2    12  
1 3 1
   12  h1 1  h1h3   
Cvgas
 1   u u       
 
  1 3    h1h2  x2 h3 x2  
 h3 x3    13    
  13  h1  1  h1h2  
 
 h1h3  x3 h2 x3 
 
Cvgas  
 11  h2 h3   dV

 h1h2 h3 x1 
 
   22 h2   33 h3 
 h1h2 x1 h1h3 x1  , (2.2.14a)

46
the momentum equation in the x2 direction can be written as:

  u  h 1  h2 h3   
  1  2   

  h1h2  x1 h3 x1  
  
u   u3  h2  1  h1h2   
 2  h h  x h1 x3  
  
   2 3 3  dV
 VdV   u  h h  
   h h h x 
Cvgas 2 1 3
t

Cvgas
  1 2 3 2

 1   u1u 2     2
  2
 
     
u 1 h 1

u 3 h3 
 h1 x1    12     h1h2 x2 h2 h3 x2  , (2.2.14b)
 1   u 2  p    12 h2 1  h2 h3 
  
  

 2
h2 x2    22 
 dV
 
 
h3 x1  
 
Cvgas  h1h2  x1
 1   u u    
   2 3     23  h2 1  h1h2  
 h3 x3    13      
 h2 h3  x3 h1 x3 
 

Cvgas  
 22   h1 h3   dV

 h1h2 h3 x2 
 
   11 h1   33 h3 
 h1h2 x2 h2 h3 x2 

and the momentum equation in x3 can be written as:

  u  h 1  h1h3   
  2  3   
  h2 h3  x2 h1 x2  
  
u   u1  h3  1  h2 h3   
 3  h h  x h2 x1  
    1 3  1
  dV
 VdV   u  h h  
t    h h h x
3 1 2
Cvgas

Cvgas 
  1 2 3 3 
 1   u u    
  1 3
  u 2
 h u 2
h
  2 2
 1 1

 h1 x1    13     h2 h3 x3 h1h3 x3  . (2.2.15)
 1   u u  
  23  h3 1  h1h3   
      dV
2 3
   
 h2 x2    23   
h1 x2  
 h 2 h3  x 2
Cvgas
  2 
 1   u3   
 13  h3 1  h2 h3  


 h3 x3 p   33       
    h1h3  x1 h2 x1 


Cvgas  
  h h   dV
33 1 2 
 h1h2 h3 x3 
 
   22 h2   11 h1 
 h2 h3 x3 h1h3 x3 

47
2.2.3 Energy Equations

The energy equations can also be written in the general orthogonal coordinate

system using the previously defined transformations. The dissipation function, , that

appears can be written as:

 2
   2 e11  e22
2

 e33
2
 e12
2
 e13
2
 e23
2
 2
 e11  e22  e33   ,
2
(2.2.16)
 3 

where the eij’s are defined in Eq. 2.2.12. The energy equation for Cvsolid (Eq. 2.1.22a) can

be transformed to:

 k T  h2 h3  
  
t  eT dV
 h1 x1 x1 
1  k T  h1h3  

Cvsolid

 eT  p u1, surf  
dV
 h1h2 h3  h2 x2 x2 
 1      
  k T  h1h2  
Cvsolid
 h x   k T  
 1 1  h x    h3 x3 x3  , (2.2.17a)
  1 1  
  eT  p u 2, surf    
q rad dV  
q gen dV
 1   k T 
   h2 x2    dV

Cvsolid

Cvsolid
 h2 h3  
Cvsolid   2 2
h x  u1, surf 
 e  p u3, surf    x1 
 1   T 
   eT  p   u  h1h3  
 k T
 h3 x3    
  h1h2 h3 
 2, surf
x2 
 dV
  h3 x3   Cvsolid
  h1h2  
 u3, surf 
 x3 

48
and the energy equation for Cvgas (Eq. 2.1.22b) can be transformed to:

 k T  h2 h3  
 
 h1 x1 x1 
1  k T  h1h3  
 
h1h2 h3  h2 x2 x2 
dV
 
  k T  h1h2  
Cvgas

 h3 x3 x3 

t  e dV T
 
Cvgas
q rad dV  
q gen dV
Cvgas
Cvgas
 k T     h2 h3  

  eT  p u1    u1
x1

 1   h1 x1    
 h x  N    eT  p    h1h3  
 1 1 
  
 s eT s   p s u1s  
 
 h1h2 h3 
 u 2
x2 
 dV
  s 1 
Cvgas
    h1h2  
 k T   u3 
  e  p u 2    x3 
 1   T h2 x2  
  

 N  dV  N
  
 s eT s   p s u1s 
 
h2 h3 
 
 2 2 
 s eT s   p s u 2 s  
h x 
Cvgas
   s 1 x1 
 
  s 1   N 
 h1h3  
 
 k T   1 

 eT  p u3       s eT s   p s u 2 s  dV

h3 x3   hh h  x2 
 1   Cvgas 1 2 3  s 1 
  h x  N   N  h h  
 3 3 
   
 s eT s   p s u3s  


 s 1
 
 s eT s   p s u3s  1 2 
x3 
  s 1 
 1       
 
h h
eT  p u1, surf 
2 3
 u1, surf 
h  x  x1  .(2.2.17b)
 1 1 
 eT  p    h1h3  
 
 1 

 h x eT  p u 2, surf  dV
     u 2, surf  dV
Cvsolid 
h1h2 h3  x2 
2 2

 
Cvsolid
  1  eT  p u3, surf    h1h2  
 h3 x3   u3, surf 
 x3 
  dV
Cvgas

2.2.4 Solid Volume Equation

Applying the above transformations to the solid volume equation (Eq. 2.1.23),

yields:


t  1dV
Cvsolid   h2 h3  
 1 u1, surf   u1, surf 
 x1 
  1   h1h3  
 h1 x1
 1 u 2, surf

  h1h2 h3   u 2, surf
x2 
. (2.2.18)
   h
 2 x2


Cvsolid 
 u3, surf  h1h2  

Cvsolid
 1 u3, surf   x3 
 h 
 3 x3 

49
2.3 Governing Equations in Two-Dimensional Vector Form

The research presented in this dissertation is limited to problems that can be

represented in a general orthogonal two-dimensional or two-dimensional axisymmetric

coordinate system, such as that shown in Figure 2.2. This will cause any terms that

contain velocities or derivatives in the x3 direction to be set equal to zero. In addition to

the geometrical restriction, the solid phase will be assumed to be of constant density, and

to have a constant composition (i.e. no chemical reactions in the bulk of the TPS

material).

Consider that a non-charring ablation problem is given that contains Nsgas species

in the gas phase, and one species in the solid. The gas portion of the problem is governed

by a system of Nsgas +3 equations: Nsgas species continuity equations (Eq. 2.2.28b),

momentum equations in the x1 and x2 directions (Eq. 2.2.14a-b), and an energy equation

(Eq. 2.2.17b). The solid portion of the problem is governed by a system of 2 equations:

an energy equation (Eq. 2.2.17a) and the solid volume equation (2.2.18).

The system of equations for the solid portion of the problem can be written in

vector form as:

  1 Fsolid 1 G solid 
t  U
Cvsolid
solid dV    h
Cvsolid 1 x1

h2 x2 
dV 
 R
Cvsold
solid  S solid dV . (2.3.1a)

Usolid is a vector of unknowns given by:

 e 
U solid   T  . (2.3.1b)
 1 

50
Fsolid contains the fluxes in the x1 direction and is given by:

k T 
 eT u1, surf  
Fsolid     h1 x1  , (2.3.1c)
  u1, surf   0 

where the first set of terms represent fluxes due to ablation, and the second term contains

heat transfer terms. Gsolid contains the fluxes in the x2 direction and is given by:

k T 
 eT u 2, surf  
G solid     h2 x2  . (2.3.1d)
  u 2, surf   0

The term Rsolid contains source terms that arise out of the particular choice of coordinate

system:

   h2 h3   
 u1, surf x 
 eT  p      1 T h2 h3   
1
 h1h2 h3  h1h3      h x x 
  u 2, surf 
x2    
k  1 1 1
 
R solid    h1h2 h3  1 T  h1h3    . (2.3.1e)
   h2 h3      
 u1, surf x     h2 x2 x2  
 1  1    
 
0
 hh h   h 
1 3 
h
 1 2 3  u 2, surf  
  x2  

Finally Ssolid contains any volumetric source terms:

q  q gen 
S solid   rad . (2.3.1f)
 0 

The system of equations for the gas portion of the problem can be written in

vector form as:


t  U gasdV
Cvgas
 1 Fgas 1 G gas 
 R gas  S gas dV
  Cvgas
  h x  h x dV  . (2.3.2a)
Cvgas  1 1 2 2    R surf dV
 1 Fsurf 1 G surf  Cvsolid
  
 h x  h x
Cvsolid  1 1 2 2
dV

51
The vector Ugas contains the conservative unknowns for the gas problem:


U gas  1, gas   s , gas   Nsgas u1 u2 
eT T . (2.3.2b)

Fgas contains the flux terms in the x1 direction:

 1u1,1 
 1u1   0   
     .
.   s u1, s 
  s u1   0   
       . 
Fgas    Ns u1     0     Nsgasu1, Nsgas , (2.3.2c)
   11 
 
gas

 u1  p     12  
2 0
0 
 u1u2   k T   N 
  s eT  s   ps u1s 
e  p u    
 1  h 
   s 1x 


T 1 1
Fgas_ iv F gas _ vh Fgas_ md

which can be split into vectors that contain inviscid fluxes, Fgas_iv, viscous and heat

transfer fluxes, Fgas_vh, and mass diffusion terms, Fgas_md. A similar vector Ggas can be

written for the fluxes in the x2 direction:

 1u2
 0  
  1u2,1 
  
   .   .
 .   0  
  su2     u 
 
s 2, s
    
    
G gas        Nsgasu2, Nsgas . (2.3.2d)
  Ns u2   0 
 gas
    12   
 u1u2    0 
  
 u 2  p     22  0
   k T   
 
2 N


 eT  p u2      e
s T s   p u 
2s 
  h2 x2  s  
s

G gas_ iv 
1

G gas_ vh G gas_ md

There are also vectors that contain the effects of the movement of the solid surface, Fsurf

and Gsurf defined as:

52
 1u1, surf   1u2, surf 
 .   . 
   
  s u1, surf    s u2, surf 
   
     .
Fsurf  , G surf  (2.3.2e)
  Ns u1, surf    Ns u2, surf 
 gas   gas 
 0   0 
 0   0 
   
 eT u1, surf   eT u2, surf 

The geometrical source term Rgas can be split into components for inviscid flow (Rgas_iv),

viscous flow and heat transfer (Rgas_vh), and mass diffusion (Rgas_md), which can be

defined as:

  1   h2 h3   h1h3   
 u1  u2  
 h1h2 h3  x1 x2  
 . 
   s   h2 h3   h1h3   
 u1  u2  
 h1h2 h3  x1 x2  
  
   Nsg a s   h h   h h   
 u1
2 3
 u2 1 3
 
 h1h2 h3  x1 x2  
 

  u  h 1  h1h3    
  2  1   

   h1h2  x2 h3 x2   
 u1   
    u1  h2 h3   

R gas_ iv    h1h2 h3 x1  
  
 
  u 2
 h  
  2 2
 h h  x  
 1 2 1 
   u  h 1  h h    
   1 

2
 2 3   

   h1h2  x1 h3 x1   
  2
u  
      u 2  h1h3   
   h1h2 h3 x2  

  
  u 2
 h 
 1 1
  h1h2 x2  
 , (2.3.2f)
   e T  p     h h
2 3   u2
 h 1 3 
h  
 h1h2 h3  1 x1 
u
   x 2 

53
 0 
 . 
 
 0 
 
  
 0 
  12  h1 1  h1h3   
    
 h1h2  x2 h3 x2  
R gas_ vh     11  h2 h3   22 h2  33 h3   , (2.3.2g)
    
  h1h2 h3 x1 h1h2 x1 h1h3 x1  
  12  h2 1  h2 h3   
    
 h1h2  x1 h3 x1  
   22  h1h3   11 h1  33 h3  
    
  h1h2 h3 x2 h1h2 x2 h2 h3 x2  
  k T   
  h2 h3   k T  h1h3  
  h1 x1 x1 h2 x2 x2  

and

  1    
 u11 h2 h3   u 21  h1h3  
 h1h2 h3  x1 x2  
 . 
  s     
 u11 h2 h3   u 21 h1h3  
 h1h2 h3  x1 x2  
  
   Nsgas     

  h1h2 h3  1N x1 2 3
u h h   u 2 N h1h3   .
R gas_ md x2   (2.3.2h)
 0 
 
 0 
 N 
  
   s eT s   p s u1s  h2 h3   
  1  s 1 x1 
h h h  N 
 
 1 2 3  
   s eT s   p s u 2 s  h1h3  
 s 1 x 
 2

54
The geometrical source term resulting from the surface movement, Rsurf can be written

as:

  1   h2 h3   h1h3   
 u1, surf  u2, surf  
 h1h2 h3  x1 x2  
 . 
  s   h2 h3   h1h3   
 u1, surf  u2, surf  
 h1h2 h3  x1 x2  
  

 h2 h3   h1h3   .
R surf (2.3.2i)
 N 
  1, surf
u  u  
x1 x2 
2 , surf
 h1h2 h3  
 0 
 0 
 
  eT  p  u  
h h   u2,surf h1h3  

 h1h2 h3  1, surf x1 2 3 x2 

Finally the source vector Sgas is defined as:

 w 1 
 . 
 
 w s 
 
  
S gas   . (2.3.2j)
w Ns g a s
 
 0 
 0 
 
q rad  q gen   

2.4 Auxiliary Equations

The system of governing equations for the solid contains two equations with six

unknowns. The system of governing equations for the gas contains Nsgas + 3 equations

with 5Nsgas + 10 unknowns. A listing of these unknowns is found in Table 2.1. (The

unknowns denoted as 1,s,Ns come from the species continuity equations with species

velocities and energies coming from the mass diffusion terms.) In order to close the

55
systems, additional equations are needed. In this dissertation the gas composition will be

a mixture of ten species: O2, N2, O, N, NO, NO+, e-, CO, CN, and C3. The data necessary

for describing this mixture is found in Appendix C.

Table 2.1: Unknowns in Each System of Equations.

2.4.1 Describing the Composition of a Mixture

In order to proceed, it will be useful to define several variables that will be used to

describe the composition of a mixture of N species. The definitions are based on

Anderson75.

First define Ns as the total number of moles of species s, and N as the total

number of moles present in the mixture. The mass of each mole of s can be denoted as

Ms, and the mass per mole for the mixture is M. The specific gas constant, R, for a

mixture can be found by:

R
R , (2.4.1)
M

where R is the universal gas constant, which has a value of 8314 J/(kg * mol * K) in SI

units. The mass of species s, Ms, present can be found by:


56
M s  Ns Ms . (2.4.2)

The molar concentration, Cs, which is the number of moles of species s per unit volume

can be found by:

s
Cs  . (2.4.3)
Ms

The mole fraction, Xs, is defined as the ratio between the number of moles of species s

per mole of the mixture:

NS
Xs  . (2.4.4)
N

The mass fraction, cs, is defined as the mass of species s per mass of mixture:

s
cs  . (2.4.5)

The sum of the mass fractions and the sum of the mole fractions must be equal to one:

N N

c
s 1
s  1, X
s 1
s  1. (2.4.6)

The mass fraction and mole fraction can be related by:

M
X s  cs   . (2.4.7)
 Ms 

The specific gas constant for a mixture R can be obtained by:

N
R   cs Rs , (2.4.8)
s 1

where the individual Rs are found using Eq. 2.4.1, with M replace by Ms. Finally the

molar mass for the mixture M can be found using:

57
1
M N
. (2.4.9)
cs

s 1 Ms

The number of particles of species s per unit volume, ns, is found by:

Ns N A
ns  , (2.4.10)
 1dV
Cv

where NA is Avogadro’s number, which is equal to 6.02 * 1026 particles /(kg * mole).

The mixture pressure can be found by Dalton’s law which gives:

N
p   ps , (2.4.11)
s 1

and the mixture density can be found by:

N
   s . (2.4.12)
s 1

2.4.2 Thermodynamic Relations

A set of equations will now be defined to relate temperature, density, pressure and

energy terms based on thermodynamic principles (see 75, 76). The energy equations for

the solid and the gas both contain the term eT. The density can be found by adding all of

the partial densities up from the species continuity equations (Eq. 2.4.12). The specific

total energy, eT, can be defined as:

u12  u22
eT  e, (2.4.13)
2

where e is the specific internal energy, which can be found in terms of the species

concentration and species specific internal energy, es:

58
N
e   cs es . (2.4.14)
s 1

The individual species specific internal energies can be found by:

es   Cv , s T dT  es0 ,
T
(2.4.15)
Tref

where Cv,s(T) is the specific heat of species s at temperature T, Tref is a reference

temperature, and es0 is the specific energy of formation of species s at Tref. If the only

data given for a species is specific heat at constant pressure, Cp,s(T), and enthalpy of

formation, hs0 , as is often the case, the following relations can be used to obtain the

required data:

es0  hs0  RsTref


. (2.4.16)
Cv , s T   C p , s T   Rs

Curvefits for the function Cp,s(T) and other thermodynamic data (taken from 76, 77and 78)

are given in Appendix C.1.

Finally, the partial pressure of species s, ps, can be found by the ideal gas equation of

state:

ps   s RsT , (2.4.17)

2.4.3 Diffusion Velocity

The system of governing equations for the gas phase contains terms sVs which

represent diffusive mass flux. This mass diffusion can be caused by gradients in

concentration, pressure or temperature, with the last two being generally neglected. The

diffusive mass flux due to gradients in concentration can be written as 75:

59
 s Vs   Dscs , (2.4.18)

where Ds is the mass diffusion coefficient of species s. In a general orthogonal c

coordinate system this can be written (using Eq 2.2.5) as:

 cs
 s u1, s   Ds
h1 x1
. (2.4.19)
 c
 s u 2, s   Ds s
h2 x2

2.4.4 Transport Properties

Values of the transport properties k (thermal conductivity),  (viscosity), and Ds

molecular diffusion coefficient must be determined. The temperatures behind a

hypersonic shock wave are high enough to cause the air to become chemically reactive,

which leads to a mixture of gas that can change significantly in temperature and

composition from point to point; this renders the use of tabular data to look up the

transport properties less effective.

One way to calculate the transport properties for an arbitrary gas mixture is based

on collision integrals76. For a given pair of species, s and r, two collision integrals can be

determined, s1,r,1 , and s2,r, 2  , which can be found in Appendix C.3. Modified versions

of these collision integrals:

1
8 
2Ms Mr
   T  ,
2
 T   
1

1,1
(2.4.20a)
3  RT Ms  Mr 
s ,r s ,r

and

60
1
16  
2Ms Mr
   T ,
2
2s ,r T    
1,1
(2.4.20b)
5  RT Ms  Mr 
s ,r

are used in determining the transport properties.

The thermal conductivity for the gas phase can be found by:

   
 N   
k gas  k B   N k  ,
15 Cs Cs
B  N (2.4.21)
   
  asr Cr  s ,r T    Cr 1s ,r T  
4 s 1 2 s  mol

 r 1   r 1 

where kB is the Boltzmann constant, Cs is the molar concentration defined in Eq. 2.4.3,

the notation s = mol means the summation is only over species that are molecules, and as,r

is defined as:

 ms  m 
1   0.45  2.54 s 
 1 
mr  mr 
as , r 2
, (2.4.22)
 ms 
1  
 mr 

where ms is the mass of one particle of species s. The first term in Eq. 2.4.21 represents

translational energy thermal conductivity, and the second term represents rotational

thermal conductivity. 76

The viscosity of the gas mixture can be written in terms of the modified collision

integrals as 76:

N
msCs
  N
. (2.4.23)
s 1
 C  T 
r 1
r
2
s,r

The mass diffusion coefficient for species, Ds, can be determined from binary

diffusion coefficients, Ds,r, using 75:

61
Ds 
1  X s  , (2.4.24)
N
Xr

r 1 Ds , r
rs

where Xs is the mole fraction of species s defined in Eq. 2.4.7. The binary diffusion

coefficients are found in terms of the modified collision integrals using 76:

k BT
Ds , r  , (2.4.25)
p1s , r T 

where p is the mixture pressure.

For the solid, the thermal conductivity is assumed to be a known function of the

temperature:

ksolid  k T  , (2.4.26)

A curve fit for solid graphite is given in Appendix C.4

2.4.5 Chemical Reactions in the Gas Phase

The species continuity equations contain chemical reaction source terms w s .

Consider that the gas phase problem is defined as having Nsgas species, and that these

species can participate in Nr chemical reactions. In this dissertation, the chemically

reacting flow, originating from the high-temperature decomposition of air and the

ablation processes, will be treated with a non-equilibrium model. The chemical reaction

source term can be written as 76:

 
w s  Ms   s ,r   s ,r R f ,r  Rb,r  ,
Nr
(2.4.27)
r 1

62
where s,r is the stoichiometric coefficient of species s on the product side of reaction r,

s,r is the stoichiometric coefficient of species s on the reactant side of reaction r, Rf,r is

the forward reaction rate for reaction r, and Rb,r is the backward reaction rate for reaction

r. The forward reaction rate is given by:

 s ,r
 Nsgas
 .001 s  
R f ,r  1000k f ,r     , (2.4.28)
 s 1  Ms  

where kf,r is the forward reaction rate coefficient. The backward reaction rate is given by:

 Nsgas  .001   s ,r 
Rb,r  1000kb,r   s
  , (2.4.28)
 s 1  Ms  

where kb,r is the forward reaction rate coefficient. It should be noted that the factors of

1000 and .001 are necessary to convert the reaction rate to SI from cgs units because

most reaction rate data found in the literature is in cgs units. In general the reaction rates

are given by:

  E f ,r 
 
k f ,r T   C f ,rT en  k BT 

k f ,r T 
, (2.4.29)
kb ,r T  
K c ,r T 

E f ,r
where Cf,r, n, and are constants provided for a reaction, and Kc,r is an equilibrium
kB

coefficient usually given as a function of temperature.

As an example consider a mixture of O2 and O. Consider that only one reaction

occurs:

O2  O 
 3O . (2.4.30)

63
For this reaction,  O2  1 ,  O  1 ,  O2  0 , and  O  3 , so the chemical reaction source

terms can be written as:

  
1000k  .001 O 2  .001 O  
 
   MO 2  MO  
f

w O 2  MO 2  
   .001 3  
  1000 k  O
  
b
   M   
 O

. (2.4.31)
  
1000 k  .001 O 2  .001 O  
 
   MO 2  MO  
f

w O  2MO  
   .001 3  
  1000k  O
  
b
   M   
 O

Appendix C.2 contains the details of the chemical reaction models that will be

used in this dissertation.

2.4.6 Surface Chemical Reactions

In non-charring ablation the mass loss occurs only through surface reactions. The
~ , can be defined as the sum of the surface
ablative mass flux vector from the surface, m abl

~ :
production rates of species s, m s

N
~  m
m abl ~s .
s 1
(2.4.32)

The bulk injection velocity, Vinj, can be related to the ablative mass flux by:
~
m
Vinj  abl
, (2.4.33)
 gas

where gas is the density of the gas at the surface. The velocity of the solid surface

recession, Vsurf, can be found by:


64
~
m
Vsurf   abl
, (2.4.34)
 solid

where solid is the solid density. Vsurf and Vinj are taken relative to the inertial reference

frame.

By conservation of mass, the amount of mass leaving the solid phase must be

equal to the amount of mass entering the gas. For each species, this mass balance can be

written as:

 s Vinj  Vsurf  Vs   m
~ ,
s (2.4.35)

where the left hand side comes from Eq. 2.1.11b. Expanding the left hand side gives:
~ ,
 s Vinj   s Vsurf   s Vs  m (2.4.36)
s

where the first term is convective mass flux, the second term is mass flux due to the

surface movement and the third term is diffusive mass flux. Using Eq. 2.4.18, to rewrite

the diffusive mass flux yields:


~ .
 s Vinj   s Vsurf  Dscs  m (2.4.37)
s

Finally applying the definition of mass concentration (Eq. ) and the definitions of

injection velocity and surface velocity (Eq. 2.4.33-34) provides:

~  c  gas m
cs m ~   D c  m~ , (2.4.38a)
 solid
abl s abl gas s s s

or grouping like terms:

   ~ ~
1  gas cs m abl   gasDs cs  m s . (2.4.39b)
  solid 

Equation 2.4.39b can be written in general orthogonal coordinates as:

65
   ~  gas cs ~
1  gas cs m1, abl  Ds  m1,s
  solid  h1 x1
. (2.4.40)
   ~  gas cs ~
1  gas cs m2 , abl  Ds  m2 ,s
  solid  h2 x2

Equation (2.4.39b) is comparable to the surface mass balance provided by Milos and

Rasky 43; Chen and Milos 79 provide the mass balance without accounting for surface

recession as:
~ m
  gasDscs  csm ~ . (2.4.41)
abl s

In the gas phase of the ablation problem there will be a species mass balance

equation (Eq. 2.4.40) in each direction for each species s. For N species, this yields a

system of N equations with 3N + 2 unknowns. The N diffusion coefficients can be found

by Eq. 2.4.24. The ablative mass flux is defined by Eq. 2.4.32. If the mixture density is a

given, only N surface reaction rate equations are needed to close the system of equations.

In the case of graphite ablation, three surface reactions will be considered which

have different temperature ranges where they are significant (see Figure 1.3). These

reactions are oxidation which is important up to 4000K:


O  C solid  CO , (2.4.42a)

nitridation which is important in a narrow temperature range between 3000 and 4000K:


N  C solid  CN , (2.4.42b)

and sublimation of C3 which becomes significant above 3500K:


3C solid  C3 . (2.4.42c)

66
For the first two reactions, the rate equations are found in 79. The reaction rate for

the oxidation reaction is given by:

~  c v  MCO ,
m (2.4.43)
MO
CO O O O

where cO is the concentration of O, vO is a kinetic velocity of the reaction, and O is the

reaction coefficient given by:

1160

 O  0.63e Tw
, (2.4.44)

where Tw is the temperature of the solid surface. For the nitridation reaction, the rate is

given by:

~  c v  MCN ,
m (2.4.45)
MN
CN N N N

where N is assumed to be a constant 0.3. The kinetic velocity for species s is given by:

k BTw
vs  . (2.4.46)
2ms

These reactions also provide surface production rates for O and N:

~  m
~ MO
m , (2.4.47)
MCO
O CO

and

~  m
~ MN
m . (2.4.48)
MCN
N CN

All other surface production rates (with the exception of sublimation of C3) are

identically zero.

67
There are two models of carbon sublimation that will be considered in this

dissertation. In general the sublimation rate is a function of the difference between the

equilibrium value of a property and the actual value (usually concentration, partial

density or partial pressure.

One of the rate equations, based on the Hertz-Knudsen approach, is provided by


79
:
~  c
m C3 
C3 ,eq  cC3 vC3  C3 ,  (2.4.49)

where  C3 is unity, and the equilibrium concentration of C3, cC3 ,eq , is given by:

pC3 ,eq
cC3 ,eq  . (2.4.50)
p

The equilibrium partial pressure of C3, pC3 ,eq , is given as:

90845

pC3 ,eq  5.19 *1015 e Tw


. (2.4.51)

The Knudsen layer is a very thin region near the surface where an evaporating or

subliming material transitions into translational equilibrium. Ytrehus 44, 45 developed a

surface mass production rate based on relations developed for the Knudsen layer. First

three non-dimensional ratios are defined: the speed ratio, S:

Vinj
S , (2.4.52a)
2 RC3 T

the pressure ratio between the actual and equilibrium partial pressures, Z:

pC3 ,eq
Z (2.4.52b)
pC3

and the temperature ratio, :

68
T
 . (2.4.52c)
Tcon

The integration of the half-range Maxwell equations across the Knudsen layer yields the

following functions of the speed ratio:

F  (S )   S  1  erf (S )  e S ,
2
(2.4.53a)

 
G  ( S )  2S 2  1 1  erf S  
2
Se S ,
2
(2.4.53b)

and

  5  S 2  2  S 2
H  (S )  S  S 2   1  erf S    e . (2.4.53c)
2  2  2 

The system of moment equations for the Knudsen layer can be written as:

Z  F  S   2  S
Z  G  S   4S 2  2 , (2.4.54)
 5
Z  H  S   S   S 2  
 2

which if  (an amplitude function) and  are written in terms of S and Z:

4S 2  2  Z
 S , Z  
G  (S ) , (2.4.55)
2  S   S , Z F  ( S )
 S , Z  
Z

becomes the speed-pressure equation:

 5
Z   S , Z  ( S , Z ) H S    S , Z S   S 2    0 . (2.4.56)
 2

For a given pressure ratio, Eq. 2.4.56 can be solved for S, and then the species surface

production rate can be found by:

69
 
~  1  pC3 ,eq  pC3  S , Z F  S  ,
m (2.4.57)
2  RC3 Tw
C3
RC3 T 

which is similar to the other proposed rate (Eq. 2.4.49) with the exception of a non-linear

reaction rate coefficient. In order for the Knudsen layer relations that this rate is based on

to be applicable, the injection velocity must be less than or equal to the local speed of

sound.

A comparison between the Hertz-Knudsen-based (HK) and Ytrehus (Yt) models

in a simple ablation problem has been performed80. Figure 2.2 shows the results of a one-

dimensional transient sublimation problem with the solid and gas at temperatures of

4000K and 4250K. The simulation was run until the equilibrium state was reached. At

both temperatures the behavior of the two models is both quantitatively and qualitatively

different. Oscillations around zero mass flux (equilibrium), occur in both models but in

the Hertz-Knudsen they are more severe. These oscillations were also noted in the

author’s previous laser ablation research 37, 81, 82, 83. It should also be noted that Chen and

Milos experienced similar problems with the behavior of charring ablators 50.

70
15 HK 4000
Yt 4000
HK 4250

Mass Flux ((kg/s)/m )


2
Yt 4250

10

0 5E-07 1E-06 1.5E-06


Time (s)

Figure 2.3:
(b)Comparison of Ytrehus and Hertz-Knudsen Carbon Sublimation
(c)Models at
15 15
4000K and 4250K.
HK 4500 Yt 4500
Mass Flux ((kg/s)/m )
Mass Flux ((kg/s)/m2)

10 10
2.4.7 Surface Energy Balance

At the surface, the heat flux must be balanced. This means that the conduction
5 5

into the gas and the heat flux due to mass diffusion at the surface must be equal to the

enthalpy absorbed by the chemical reactions and the conduction into the solid. The exact
0
0
0 2E-05 4E-05 6E-05 8E-05 0.0001
0 2E-05 4E-05 6E-05
form of this equation
Time (s) can be8E-05 0.0001
found employing a similar process as the surface
Timemass
(s)

balance:

 
Nsgas  
  ~   e  p   e  

kT gas    eT ,s  ps  Ds cs  kT solid  m

abl   T
  T solid  , (2.4.58)
s 1  s     abl
   

hs  h0 

where the subscript abl refers to the gas being created at the surface. Aside from notation,

the use of the total energy instead of the enthalpy, and the elimination of radiation terms,

this is identical to the energy balance proposed by Milos and Rasky 43.
71
2.4.8 Initial and Boundary Conditions

The governing system of equations (Eq. 2.3.1-2) need to have initial and

boundary conditions in order to completely describe the non-charring ablation problem.

Boundary conditions must be specified for the solid and gas boundaries, with the

exception of the surface between the solid and gas; the boundary conditions are

embedded within the governing equations, and do not need to be specified.

In the gas portion of the problem, several conditions are needed including inlet,

outlet, and/or symmetry (see Fig. 3.1). At the inlet, the flow variables will generally be

specified (Dirichlet boundary condition) as the freestream conditions:

U gas inlet  U gas  ; (2.4.59)

it should be noted that if the inflow is subsonic, some information at the boundary must

come from the interior 84. At the outlet, the boundary condition depends again on the

nature of the flow: if the flow is supersonic, the boundary value depends solely on the

interior behavior of the solution, but if the flow is subsonic one physical condition must

be set. If the coordinate system is chosen such that an axis of symmetry exists, the

boundary condition is given on the derivatives of the conservative unknowns (Neumann

boundary condition):

 U gas 
   0, (2.4.60)
 n  symmetry

where n denotes the normal direction.

Within the solid, boundary conditions on temperature must be given as needed.

For example there could be an axis of symmetry where a condition like (2.4.60) could be

72
given, or if there is some inner boundary to the TPS, a constant temperature or heat flux

could be given.

Initial conditions must also be specified for the problem. These will consist of

setting the entire gas to the freestream conditions and the solid to some initial temperature

or to some initial approximation.

2.5 Summary

In this chapter the proposed mathematical model for the non-charring ablation

problem has been presented. This model contains chemically-reacting multi-species gas

dynamics, and heat transfer within the solid. These two sub-models are tightly coupled

by the surface ablation reactions and the resulting movement of the solid surface. This

coupling is implicitly contained within the governing equations, and is not handled by

explicitly stated boundary conditions.

In addition to deriving the governing equations, the auxiliary relations necessary

for their solution have been presented. Various methods for describing the composition

of the gas have been defined. The chemical non-equilibrium model for determining the

species production rates was identified. Boundary and initial conditions were discussed.

The mathematical model for the surface reactions was detailed, and a preliminary

comparison between the Hertz-Knudsen and the Ytrehus models for sublimation was

given.

73
CHAPTER III

NUMERICAL METHODS

In this chapter the numerical methods that are required for the discretization of the

governing equations presented in the previous chapter are presented. First, these

equations are discretized using a finite-volume method that has been modified to allow

for an arbitrary mixture of solid and gas. Second, the standard and reactive Riemann

problems and their solvers are detailed.

3.1 Discretization of Equations

This section contains the discretization techniques for the governing equations

that were developed on the previous chapter. Finite-volumes are defined that can be used

for an arbitrary mixture of solid and gas and then the governing equations are discretized

in terms of space and time.

3.1.1 Definition of Finite Volumes

Consider some arbitrary geometry for an ablation problem such as in Fig. 3.1a.

Now consider that an arbitrary, structured, orthogonal grid is overlaid on the problem

such as in Fig. 3.1b, and it has (imax-imin)*(jmax-jmin) points. A control volume,

corresponding to CvO in the mathematical model that has been developed, can be drawn

74
around any of these points (i, j), with faces corresponding to the halfway points in-

between each point that are labeled (i±½, j) and (i, j±½). If the location of each point is

given by ((x1)i,j, (x2)i,j), then the location of the faces can be defined as:

x1 i 1, j  x1 i , j


x1 i  12, j 
2
x1 i , j  x1 i 1, j
x1 i  12, j 
2
x2 i , j 1  x2 i , j . (3.1.1)
x2 i , j  12 
2
x2 i , j  x2 i , j 1
x2 i , j  12 
2

It should be noted that the body-fitted coordinate requires that no point lies on the

symmetry line. These points are offset by ½x1 so that their faces correspond with the

axis of symmetry

75
M oo
Too
oo
P oo

Solid Body

Axis of Symmetry
(a)

imax, jmax

M oo
Too
oo
poo

i, j
imax, jmin

Solid Body

imin, jmax imin, jmin


Axis of Symmetry

(b)

Figure 3.1: Schematics of a typical TPS ablation problem. (a) Arbitrary ablation problem
(b) Arbitrary grid and associated control volume

76
Within the overall control volume it will be helpful to define a local coordinate

system, (1, 2), such that point (i, j) is located at (0, 0), faces (i±½, j) are located at (±1,

2), and faces (i, j±½) are located at (1, ±1).

If this original control volume contains some ablating material, another control

volume is drawn around the ablating material, and corresponds to Cvsolid. The faces of

this control volume will be labeled (i±w, j) and (i, j±w); this notation is used to conform

to the finite volume formulation notation which uses i±½ etc. to denote control volume

faces. The localized coordinates for faces (i±w, j) and (i, j±w) are given by:

  x1 i  w, j   x1 i , j
 if x1 i  w, j  x1 i , j
x1  1  x1 
 1 i  w, j   x i  2, j  x  i , j
 1 i  w, j 1 i, j
if x1 i  w, j  x1 i , j
  x1 i , j   x1 i  1 , j
 2
  x1 i  w, j   x1 i , j
 if x1 i  w, j  x1 i , j
x1  1  x1 
 1 i  w, j   x i  2, j  x  i , j
 1 i  w, j 1 i, j
if x1 i  w, j  x1 i , j
  x1 i , j   x1 i  1 , j
 2
  x2 i , j  w   x2 i , j
 if x2 i , j  w  x2 i , j
 x2  1   x2 
 2 i , j  w   x i , j  2  x  i , j
 2 i, jw 2 i, j
if x2 i , j  w  x2 i , j
  x2 i , j   x2 i , j  1
 2
  x2 i , j  w   x2 i , j
 if x2 i , j  w  x2 i , j
 x2  1   x2 
 2 i , j  w   x i , j  2  x  i , j
 2 i, j w 2 i, j
if x2 i , j  w  x2 i , j . (3.1.2)
  x2 i , j   x2 i , j  1
 2

The remaining volume is identifiable as Cvgas in the mathematical model. Figure 3.2

gives a schematic of a typical finite volume that contains ablating material.

77
i,j+1/2
X

Cvgas 2
X
i,j+w

i-1/2,j i-w,j i+w,j i+1/2,j


X X X X
x2

i,j 1
Cvsolid
X
i,j-w

X
i,j-1/2

x1

Figure 3.2: Schematic of control volumes associated with point (i, j).

3.1.2 Discrete Form of Governing Equations

Consider the two-dimensional vector form of the governing equations given in Eq

2.3.1-2:

  1 Fsolid 1 G solid 
dV   R solid  S solid dV ,
t Cv 
U dV   
 
solid
Cvsolid  1 
solid
h x1 h2 x2 Cvsold

and


t  U gasdV
Cvgas
 1 Fgas 1 G gas 
 R gas  S gas dV
   Cvgas
  h x  h x dV  .
Cvgas  1 1 2 2    R surf dV
 1 Fsurf 1 G surf  Cvsolid
  
 h x
Cvsolid  1 1

h2 x2
dV

78
Each of these equations consists of a series of volume integrals taken over Cvsolid or Cvgas.

Now assume that the finite volume that was defined in the previous section is small

enough that each integrand is close to some single, discrete value at every point within

the finite-volume. Also, by definition the volume integral of 1 is equal to VCv, the

amount of volume within the control volume Cv that the integral is taken over:

VCv   1dV . (3.1.1)


Cv

If the previous assumption holds then the integrands can be taken out of the volume

integrals which can be rewritten using Eq. 3.1.1 as:


t

U solid i , j VCvsolid 
 1 Fsolid 1 G solid   R solid  S solid i , j VCvsolid , (3.1.2a)
    VCvsolid
 h1 x1 h2 x2 i , j

and


t
 
U gas i , j VCvgas 
 1 Fgas 1 G gas   
R gas  S gas i , j VCvgas
   VCv
 h1 x1

h2 x2  i , j gas

 
 R surf i , j VCvsolid
, (3.1.2b)

 1 Fsurf 1 G surf 
    VCv
 h1 x1 h2 x2  i , j solid

where ( )i,j denotes the discrete value.

A volume ratio, , can be defined as the ratio between the solid volume and

overall volume of the finite-volume:

VCvsolid
  , (3.1.3)
VCvO

which can be calculated from the finite-volume face locations in the local coordinate

system (1, 2) as:


79
 
1
4
  
 1 i  w, j   1 i w, j  2 i, j  w   2 i, j w . (3.1.4)

Using Eq. 2.1.2, the ratio of the volume of the gas to the overall volume can be written

as:

VCvgas
1   . (3.1.5)
VCvO

Dividing both sides of Eq. 3.1.2a-b by VCvO , and using the volume ratio

definitions yields:


t

U solid i , j  
 1 Fsolid 1 G solid   R solid  S solid i , j  , (3.1.6a)
    
 h1 x1 h2 x2 i , j

and


t
 
U gas i , j 1    
 1 Fgas 1 G gas  
R gas  S gas i, j 1   
   1    
 h1 x1

h2 x2  i , j  R surf  i, j  . (3.1.6b)

 1 Fsurf 1 G surf 
    
 h1 x1 h2 x2  i , j

Applying the chain rule to the time derivative and algebraic manipulation yields:


t
   
U solid i, j   1 Fsolid  1 G solid 
 h1 x1 h2 x2  i , j
, (3.1.7a)
U solid i, j 
 R solid  S solid i , j 
 t

and

80
1 Fgas 1 G gas 
   
 h1 x1 h2 x2  i , j
1 Fsurf 1 G surf  
   
 1   
x1 h2 x2
 i, j    R gas  S gas i, j
  h1 i, j
U gas . (3.1.7b)
t

 R surf i , j
1   
U gas i, j 

1    t

3.1.3 Time Discretization

A first-order explicit time discretization can be written 74:


 RHS
t
 n 1   n
 RHS n , (3.1.8)
t
 n 1   n  t * RHS n

where superscript n denotes values taken at the initial time level, superscript n+1 denotes

a value taken after a time t,  is an arbitrary vector or variable, and RHS denotes a

generic right hand side. Applying the time discretization in Eq. 3.1.8 to Eq. 3.1.7a-b

yields:

  1 F 1 G solid 
n

 solid
  
  h1 x1 h2 x2 i , j  
U solid in,j1  U solid in, j  t  , (3.1.9a)
  U   
n

  R solid  S solid in, j   solid i , j   


   t  
 

and

81
  1 Fgas 1 G gas  n 
    
  h1 x1 h2 x2 i , j 
 n 
  1 Fsurf  1 G surf  n 
  h1 x1


h2 x2 i , j 1   n 


U gas in,j1  U gas in, j  t  R gas  S gas in, j . (3.1.9b)
 
 R  S n 
n

 surf

surf i , j

1  n 
  U gas i , j  
 n 

  1    t  
   

Currently the time discretization is restrained to the one described above,

However other explicit time schemes exist and could be employed with this set of

equations in future research. For example in systems of equations that contain gas phase

chemical reactions it may be advantageous to use a second order predictor-corrector

scheme with semi-implicit treatment of the chemical species source terms. Park 85

identifies this scheme as:

 t 
 I  J S dU  tRHS
n

 2 
U  U n  dU
 t  , (3.1.10)
 I  J S dU  tRH S
 2 
dU  dU
U n 1  U n 
2

where I is the identity matrix, the barred values represent an intermediate value, RHS is

the right hand side of 3.1.7b, and Jw is the Jacobian defined by:

S
J S  , (3.1.11a)
U

82
which has the vector of source terms containing only chemical reactions, S :


S  w1  w s  w Nsgas 0 0 0 . 
T
(3.1.11b).

3.1.4 Spatial Discretization of Flux Terms

Fgas
Contained within Eq. 3.1.9a-b are spatial derivatives, such as . These terms
x1

must also be defined in a discretized form. While this is a relatively straightforward

process for a case of all gas or all solid within the finite-volume, for a case of some

arbitrary mixture, there are 13 possible cases which must be dealt with. These cases

result from the location and extent of the solid. Figure 3.3 shows all of these cases.

1 2 3 4

5 6 7 8

9 10 11 12

13
14 15

Figure 3.3: Possible location and extent of solid.

83
In a typical finite-volume scheme, the spatial derivatives in the x1 direction are

found by 84:

 F  Fi  1 , j  Fi  1 , j
   2 2
, (3.1.12)
 x1 i , j x1 i  12, j  x1 i  12, j

where the flux F are required at the faces. In a mixed solid and gas finite-volume, there

are eight faces, instead of four. In a finite-volume containing an arbitrary located and

sized solid, define four zones, as pictured in Fig. 3.4a for F. The volumes of each zone

defined in the local coordinate systems, VF  zone , can be found by:

1   2 i , j  w
VF  I 
2
1   2 i , j  w
VF  II 
2
 1   1 i  w, j   2 i , j  w   2 i , j  w  . (3.1.13)
VF  III   



 2  2 
 1   1 i  w, j   2 i , j  w   2 i , j  w 
VF  IV   



 2  2 

i,j+1/2 i,j+1/2
X X
F-I G-III

X X
i,j+w i,j+w

i-1/2,j i-w,j i+w,j i+1/2,j i-1/2,j i+w,j i+1/2,j


X F-III X X F-IV X X G-I X X G-II X
x2

x2

i,j i,j

X X
i,j-w i,j-w

F-II G-IV
X X
i,j-1/2 i,j-1/2

x1 x1
(a) (b)
Figure 3.4: Zones for flux calculations. (a) F (b) G.

84
The derivative of a flux F in zones F-I and F-II can be discretized using Eq. 3.1.12. For

zone F-III, the derivative can be discretized by:

 F  Fi  w, j  Fi  1 , j
   2
. (3.1.14)
 x1  F  III x1 i  w, j  x1 i  12, j

For zone F-IV, the derivative can be discretized by:

 F  Fi  1 , j  Fi  w, j
   2
. (3.1.15)

 1  F  IV
x  x1 i  1 , j   x1 i  w, j
 2

Finally, the flux for the entire control volume can be written using an averaging

procedure based on the ratio of the volume in each zone to the total volume of gas:

   
VF  I  F  
 
  x1  F  I 
 
  F  
  V 
  

F II
 F  1   x  F  II ,
 
1
 (3.1.16a)
 x1 i , j 1    F  
  VF  III   
  x1  F  III 
 
V  F  
 
 F  IV
 
  x 1  F  IV 

or when the definitions of the different quantities are added in:

85
 1     Fi  1 , j  Fi  1 , j  
 2 i, jw  2 2  


2  x   x1 i  12, j 
 1 i  12, j
 

 
 1   2 i , j  w  Fi  12, j  Fi  12, j  
 2   x1  1   x1  1  
 F  1   i , j
2 
i , j 
  
2

1   , (3.1.16b)
 x1 i , j     
1 i  w, j      
2 i , j  w 
 F  F 
1 i  w , j i , j
1

    2 i , j  w  2
  
  x1 i  w, j   x1 i  12, j  

  2  2
 
  1   1 i  w, j   2 i , j  w   2 i , j  w  Fi  12, j  Fi  w, j  
   


  x  
  2  2  1 i  1 , j   x1 i  w, j 
  2 

where 1- is the volume of the gas in local coordinates.

A similar approach can be taken for the fluxes in the x2 direction, G, with the

definition of the zones provided in Fig. 3.4b. The volumes of each zone defined in the

local coordinate systems, VG  zone , can be found by:

1   1 i  w, j
VG  I 
2
1   1 i  w, j
VG  II 
2
 1   2 i , j  w   1 i  w, j   1 i  w, j . (3.1.17)
VG  III   



 2  2 
 1   2 i , j  w   1 i  w, j   1 i  w, j 
VG  IV   



 2  2 

The derivative of the flux G in zones G-I and G-II can be discretized as:

 G  G i, j  1  G i, j 1
   2 2
, (3.1.18a)

 2 G  I , II
x  x 
2 i, j 1
2
  x 
2 i, j  1
2

for G-III it as:

86
 G  G i, j  1  G i, j w
   2
, (3.1.18b)

 2 G  III
x  x2 i , j  1   x2 i , j  w
 2

and for G-IV as:

 G  G i, j w  G i, j  1
   2
. (3.1.18c)

 2 G  IV
x  x 
2 i, j w   x 
2 i, j 1
2

Using the same averaging procedure as used for F, the derivative for the entire control

volume can be written as:

   
VG  I  G  
 
  x2 G  I 
 
  G  
  VG  II  x  
 G  1   2 G  II ,
   (3.1.19a)

 2 i , j
x 1    G  
  VG  III   
 
 2 G  III
x 
 
V  G  
 
 G  IV
 
  x 2 G  IV 

or substituting the relevant equations:

 1     G i, j 1  G i, j 1  
 1 i  w, j  2 2  


2  x   x2 i , j  12 
 2 i , j  12
 

 
 1   1 i  w, j  G i , j  12  G i , j  12  
 2  x   x2 i , j  12   
 G  1   2 i , j  12 
    . (3.1.19b)
 1 
 2 i , j   1   2 i , j  w   1 i  w, j   1 i  w, j 
  
x G G
i, j 1 i, jw

     2
   x  
  2  2  2 i , j  12  x2 i , j  w  
 
  1   2 i , j  w   1 i  w, j   1 i  w, j  G i , j  w  G i , j  12  
   


 x  
 

2  2  2 i , j  w   x2 i , j  12  

87
In the case of a finite volume containing only gas, the location of the solid faces

are all identically (0,0) in the local coordinate system. This causes the flux equations to

reduce to Eq. 3.1.12 and 3.1.18a.

For fluxes taken over the solid portions of a finite volume (designated by solid or

surf), the spatial derivatives of the fluxes can be discretized in a similar manner to that of

Eq. 3.1.15, with the exception that the faces are the solid faces designated by ±¼ . For

the derivative of F in the x1 direction it is written as:

 Fsolid  Fsolid i  w, j  Fsolid i w, j


   , (3.1.20a)
 x1 i , j x1 i  w, j  x1 i w, j

and for the derivative of G in the x2 direction it is written as:

 G solid  G solid i , j  w  G solid i , j w


   . (3.1.20b)
 x2 i , j x2 i , j  w  x2 i , j w

3.1.5 Scalar Derivative Discretizations

 
Within the flux terms F and G are derivatives of the form or , where  is
x1 x2
an arbitrary scalar such as a velocity component or temperature. The discretized form of

these derivatives takes on different forms depending on the face that the derivative is

being taken for, and the particular configuration of the solid. These derivatives are

needed at each face, and for the overall finite volume. Different forms are needed for the

gas side and for the solid side of the surface as well. In general these derivatives take the

form of :

 


 
 B   C , (3.1.21)
 xd  A xd B  xd C

88
where d is an arbitrary direction, and A, B, and C represent arbitrary faces. Full definitions

of the required discrete forms of the derivatives are found in Appendix C.1.

3.2 Riemann Solvers for Inviscid Fluxes

In the discretizations presented in section 3.1.4, values of the fluxes F and G are

needed at the faces of each finite-volume that contains at least some gas. The flux

vectors Fgas and Ggas, defined in Eq. 2.3.2c-d contain the inviscid flux vectors:

 1u1   1u2 
 .   . 
  s u1    s u2 
     
Fgas_ iv    Ns u1 , G gas_ iv    Ns u2  . (3.2.1)
 gas
  gas

 u1  p   u1u2 
2

 u1u2   u22  p 
e  p u  e  p u 
 T 1  T 2

One method of finding the value of these inviscid fluxes at an interface, for example

(Fgas_iv)i ½,j, is to solve the associated Riemann problem.

The standard Riemann problem is an initial value problem that has two states UL

and UR, on either side of a point at some initial time zero. A system of waves composed

of shock fronts, contact discontinuities and/or expansion waves evolves after this initial

time; Fig 3.5a shows a typical set of waves. For the inviscid Riemann problem there are

10 different cases86 that determine the particular combination and properties of the

various wave patterns. The goal of the standard Riemann solver is to find the value of a

specified variable or set of variables at a given point. Standard Riemann solvers fall into

two camps: exact solvers like the method detailed in section 3.2.1, and approximate

solvers such as the Steger-Warming Flux Splitting method detailed in section 3.2.2. A

89
first-order method is described in each of these sections and their extension to second-

order is provided in section 3.2.3.

The reactive-Riemann problem is similar to the standard Riemann problem, but

the left and right states can also interact through a reaction front. Le Metayer et al. 87

provide an overview of how a reactive-Riemann problem and an associated solver are

used for solving a problem that involves evaporation fronts. In the particular case of the

non-charring ablation problem considered in this derivation, the movement of the solid

surface due to the surface reactions corresponds to the reaction front, which will be

referred to as the ablation front. This reactive-Riemann problem has four more possible

cases resulting from the direction of the reactive front and the location of the solid. Figure

3.5b shows a typical wave pattern for this problem. It should be noted that in the case

where the ablation front is stationary (i.e. reaction rates are zero), the standard wall

boundary conditions should be reproduced. A reactive-Riemann solver is developed in

Section 3.2.4.

t F an t
ion
ns
e

e
urfac

pa
urfac

n Ex
Fa
sion
act S

act S

n
pa Ab
Ex la
tio
Cont

Cont

n
k

Fr
oc

oc

on
Sh

Sh

(U gas)R t (U gas)R
(U gas)L (U solid)L

i+1/2, j x1 i+1/4, j x1

(a) (b)

Figure 3.5: Riemann Problems. (a) Standard for inviscid gas dynamics
(b) Reactive for non-charring ablation.

90
3.2.1 Exact Riemann Solver

Toro86 provides the details of an exact Riemann solver for a single species gas; a

straightforward modification to multiple species gases is detailed below. The procedure

will be outlined for determination of the flux (Fgas_iv)i+½,j when both sides of the face are

gas. A similar procedure can be used for the flux G.

First, vectors of primitive variables WL and WR are defined from (Ugas)i,,j and

(Ugas)i+1,,j respectively, where W is defined as:

W  c1  cs  cN  u1 u2 p T  .
T
(3.2.2)

The solution process requires a ratio of specific heats, , for each side:

N
 T    cs s , (3.2.3a)
s 1

where s is defined as:

C p T 
 s T   . (3.2.3b)
C p T   Rs

A frozen speed of sound, , can then be calculated for each side by:

   T 
p
. (3.2.4)

The next step involves finding the intermediate values of pressure, p* , and

velocity, u1*. An algebraic equation can be written for p* as:

f L  p* , WL   f R  p* , WL   u1 R  u1 L  0 , (3.2.5)

where the function fL is given by:

91
  2 
  
 p  p      1 
L L
if p*  pL
 * L
 1
 p*  L p
  L 1 L
f L  p* , WL    , (3.2.6a)

   L 1

  L 
  2 L
 1
2 p
  *
if p*  p L
   1  p  
 L  L
 

and the function fR is given by:

  2 
  

 p  p   R R   1  if p*  p R
 * R
 R 1
 p*  p
  R 1 R
f R  p* , WR    , (3.2.6b)

   R 1

 2 R  p*  2 R  1 if p*  p R
   1  p  
 R  R  

Newton’s method can be used to solve Eq. 2.2.5 by iterating:

p*k 1  p*k 
  
f L p*k , WL  f R p*k , WL  u1 R  u1 L 
 
f L p*k , WL  f R p*k , WL ,
 (3.2.7)

until a convergence criterion is reached. The first derivatives can be written as:

 1

 

2 

2 
 


  d  d  1 
   p*  pd
  if
 1  1    
p*  pd
 p*  d 
  2 p*  d  1
 p  
f d  p* , Wd     p

 d 1 d  
    d  1 d   , (3.2.8)

   d 1
  p*  2 d
 1
  if p*  pd
  d  d  pR 

92
where d can be either L or R. Once the pressure is found by Eq. 2.2.7, the contact surface

velocity, u1 * , can be found by:

u1 L  u1 R f R  p* , WL   f L  p* , WL 
u1 *   . (3.2.9)
2 2

A sampling procedure can then be implemented to find the value of Wi+1/2,j:

S L  0 Wi  1  W*shock
p*  pL
,j
2
SL  0 Wi  1 ,j
 WL
2
u1 *  0 S HL  0 Wi  1
2
,j
 WL
p*  pL S HL  0  STL Wi  1 ,j
 WL , fan
2
STL  0 Wi  1 ,j
 W*fan
L
2
. (3.2.10)
S R  0 Wi  1  W*shock
p*  pR
,j
2
SR  0 Wi  1 ,j
 WR
2
u1 *  0 S HR  0 Wi  1
2
,j
 WR
p*  pR S HR  0  STR Wi  1 ,j
 WR , fan
2
STR  0 Wi  1 ,j
 W*fan
R
2

The left shock speed, SL, can be found by:

1
   1  p   1 2
S L  u1 L   L  L  *  L  , (3.2.11a)
 2 L  pL 2 L 

the right shock speed, SR, can be found by:

1
   1  p   1 2
S R  u1 R   R  R  *  R  , (3.2.11b)
 2 R  pR 2 R 

the left fan head speed, SHL, can be found by:

SHL  u1 L   L , (3.2.11c)

the right fan head speed, SHR, can be found by:

S HR  u1 R   R , (3.2.11d)
93
the left fan tail speed, STL, can be found by:

 L 1
p  2 L
STL  u1 L   L  *  , (3.2.11e)
 pL 

and the right fan head speed, STR, can be found by:

 R 1
 p  2 R
STR  u1 R   R  *  . (3.2.11f)
 pR 

Assuming that the composition and u2 velocity remain constant behind the contact

surface, the sampled vectors can be written as:

 c1 L 
  
 
 cs L 
 
  


 
c Nsg a s
L


W*shock    p*  L  1   ,
 (3.2.12a)
  pL  L  1  
L

 L  
  
 1 
 L 1 p*
   L  1 pL  
 u1 * 
 
 u2 L 
 
 p* 

94
 c1 R 
  
 
 cs R 
 
  


 
c Nsg a s
R


W*shock    p*  R  1   ,
 (3.2.12b)
  pR  R  1  
R

 R  
  
 1 
 R 1 p *
   R  1 p R  
 u1 * 
 
 u2 R 
 
 p* 

 c1 L 
  
 
 cs L 
 
  
 c Ns

  
,
W*fan 
gas L
(3.2.12c)
  p  L 
L 1

L  *  
  p L  
 u  
 1 *

 u 2 L 
 
 p* 

 c1 R 
  
 
 cs R 
 
  
 c Ns

  
,
W*fan 
gas R
(3.2.12d)
  p  R 
R 1

L  *  
  p R  
 u  
 1 *

 u 2 R 
 
 p* 

95
 c1 L 
  
 
 cs L 
 
  



c Nsgas 
L


  
2

WL , fan     2   L  1 u   L  ,
 1
(3.2.12e)
   L  1  L  1 L
1 L
 
L

 2   L 1  
 
 L  u 
1 L

  L 1 2  
 u2 L 
 2 L 
  2  L 1   L 1 
 pL   u1 L  
   L  1  L  1 L  

and:

 c1 R 
  
 
 cs R 
 
  



c Nsgas
R
 

   1 
2

WR , fan     2   R  1 u   R  . (3.2.12d)
   L  1  R  1 R
1 R

R

 2   1 
 R  R u1 R  
  R 1  2  
 u2 R 
 2 R 
  2  R 1   R 1 
 pR   u1 R  
   R  1  R  1 R  

The sampled primitive variables can then be used to calculate the flux (Fgas_iv)i+½,j

by plugging the values into Eq. 3.2.1. The species densities can be calculated by

multiplying the sampled concentrations times the sampled density. Calculating the total

energy requires finding the temperature using the ideal gas law and the mixture gas

constant (Eq. 2.4.8):


96
p
T Nsgas
, (3.2.13)
  cs R s
s 1

using this value to calculate the internal energy by applying Eq. 2.4.14-15, and plugging

this result into Eq. 2.4.13.

3.2.2 Steger-Warming Flux Splitting: An Approximate Riemann Solver

The exact method detailed in the previous section is computationally expensive

because of the iterative solution of the pressure equation at every point (between three to

eight iterations required at each point for simple test cases 86). In order to avoid this

computational expense, approximate Riemann solvers have been developed. One such

scheme is the so-called Steger-Warming flux splitting scheme.

Before deriving this scheme, it is useful to rewrite the fluxes using the chain rule

as:

Fgas_ iv U gas G gas_ iv U gas


 AF ,  AG , (3.2.14a)
x1 x1 x2 x2

where AF and AG are the Jacobians of Fgas_iv and Ggas_iv with respect to Ugas_iv:

Fgas_ iv G gas_ iv
AF  , AG  . (3.2.14b)
U gas U gas

The Jacobians, and the associated eigenvalues and eigenvectors can be found in 76. The

Jacobians are given as:

97
 u1 1  c1  .  c1u1 .  c1u1 c1 0 0 
 . . . . . . . . 
 
  cs u1 . u1 1  cs  .  cs u1 cs 0 0 
 
 . . . . . . . . 
AF  
 c Ns u1
 ~ gas 2
.  c Nsgasu1 . 
u1 1  c Nsgas  cN 0 0 

, (3.2.15)

u1 2   
~ ~
 1  u1 . s  u12 .  Nsgas  u12  u2  
 u u .  u1u2 .  u1u2 u2 u1 0 
 ~1 2 

u1 1  h  ~

. u1 s  hT  ~

. u1  Nsgas  hT  hT  u12  u1u2 u1 1   

and

 u2 1  c1  .  c1u2 .  c1u2 c1 0 0 
 . . . . . . . . 
 
  cs u 2 . u2 1  cs  .  cs u 2 cs 0 0 
 
 . . . . . . . . 
AG  

 c Nsgasu2 .  c Nsgasu2 . 
u2 1  c Nsgas  cN 0 0 

,(3.2.16)

  u1u2 .  u1u2 .  u1u2 u2 u1 0 


 ~  u 2 u2 2   
~ ~
. s  u22 .  Nsgas  u22  u1  
 1~ 2 

u2 1  hT  ~

. u2 s  hT  ~

. u2  Nsgas  hT   u1u2 hT  u22 u2 1   

where hT is the total enthalpy related to total energy by:

p
hT  eT  , (3.2.17a)

 is defined by:

p Rs  s
Nsgas


eT 
  c C T  ,
s 1
(3.2.17b)
s v,s

~
and  s is defined by:

~ p  u 2  u22 
s   RsT    1   es . (3.2.17c)
 s  2 

The frozen speed of sound can be defined as:

98
 1    p . (3.2.17d)

The eigenvalue matrixes F, and G can be found to be:

u1  0  0 0 0 0
        

0  u1  0 0 0 0
 
       
ΛF   , (3.2.18a)
0  0  u1 0 0 0
 
0  0  0 u1   0 0
0  0  0 0 u1   0
 
 0  0  0 0 0 u1 

and

u2  0  0 0 0 0
        

0  u2  0 0 0 0
 
       
ΛG   . (3.2.18b)
0  0  u2 0 0 0
 
0  0  0 u2   0 0
0  0  0 0 u2   0
 
 0  0  0 0 0 u2 

The right eigenvectors of AF and AG are:


~ ~ ~
 2  c11   c1s   c1 Nsgas u1c1 u2 c1  c1 
 
         
  cs~1    css
2 ~

~
 cs Nsgas u1cs  u 2 cs  cs 
 
         , (3.2.19a)
RF  
cNsgas 
~ ~ ~
c    c Nsgass    c Nsgas Nsgas
2
u1cNsgas u2 cNsgas
 Nsgas 1 
  u2   u2   u2 0 1 0 
 0  0  0 0 0 0 
~ ~ ~
 1  u1  s  u1   Nsgas  u1    u1  u2  

and

99
~ ~ ~
 2  c11   c1s   c1 Nsgas u1c1 u2c1  c1 
 
         
  cs~1    css
2 ~

~
 cs Nsgas u1cs  u 2 cs  cs 
 
         . (3.2.19b)
RG  
cNsgas 
~ ~ ~
c    c Nsgass    c Nsgas Nsgas
2
u1cNsgas u2 cNsgas
 Nsgas 1 
  u1   u1   u1 1 0 0 
 0  0  0 0 0 0 
~ ~ ~
 1  u2  s  u2   Nsgas  u2  u1    u2  

Finally, the left eigenvectors are given by:

 1 c1 
  0s  0 0 0 
2 2 2
         
 1 cs 
 0   0 0 0 
 2 2 2 
         
 1 c Nsgas , (3.2.20a)
LF   0  0  0 0
2 2 2
 u1 u1 u1 u1   
   0 0 
 2  2
2 2 2 
 u2 u2 u2 u2 
   1 0
2  2
2 2 2 

   u12  u22   s  u12  u22    Ns 
~ ~ ~
  u1  u2  1 hT  u1 
2 2
  gas
u2 0
  2  2  2
2 2 

and

 1 c1 
  0s  0 0 0 
2 2 2
         
 1 cs 
 0   0 0 0 
 2 2 2 
         
 1 c Nsgas . (3.2.20b)
LG   0  0  0 0
2 2 2
 u1 u1 u1 u1 
   1 0 
 2  2
2 2 2 
 u2 u2 u2 u2   
   0 0
2  2
2 2 2 

   u12  u22   s  u12  u22    Ns 
~ ~ ~
  u1  u2  1 hT  u2 
2 2
  gas
u1 0
  2  2  2 2 2 

100
The derivation of the Steger-Warming Flux splitting scheme can be found for a
74 , 84
single species gas in most CFD textbooks, such as or 86. A similar derivation can

be followed for multi-component gases. Any nonsingular Jacobian A, can be rewritten

as:

A  L* Λ*R . (3.2.21)

For the Steger-Warming scheme, the eigenvalue matrix is split into matrixes containing

only positive eigenvalues, + , and matrixes containing only negative eigenvalues, -.

By substituting these into Eq. 3.2.21, split Jacobians A+, and A- can be found.

Multiplying these split Jacobians by U yields the split fluxes:

F   A F U, F   A F U
. (3.2.22)
G   A G U, G   A G U

By applying this process to the fluxes Fgas_iv and Ggas_iv and simplifying, the split fluxes

can be found by:

 c1  
  
 
 cs  
 
     ,
Fgas  (3.2.23)
21     cN  
_ iv

 
 
u1    2  3  
 u2  
 
2 

 H  2 1  u1 2  3
 



and

101
 c1  
  
 
 cs  
 
    ,
G gas_ iv  (3.2.24)
21     cN  
 
 u1  


u2    2  3   

 H  2 1  u2 2  3
2   
 


where the function is defined as:

  21  2  3 , (3.2.25)

the eigenvalues for F are given by:

1  u1 2  u1   3  u1   , (3.2.26a)

and the eigenvalues for G are given by:

1  u2 2  u2   3  u2   . (3.2.26b)

The fluxes at the faces can be found by:

F 
gas_ iv i  1 , j
2


 Fgas _ iv 
i 1, j


 Fgas _ iv  i, j
(3.2.27a)

and

G 
gas_ iv i , j  1
2

 G gas_ iv 
i , j 1

 G gas_ iv i, j
. (3.2.27b)

By examination it can be assumed that the fluxes can be computed much faster

using this approximate Riemann solver rather than the exact Riemann solver of the

previous section. However, this increase in efficiency comes with a possible decrease in

accuracy. The scheme becomes inaccurate near sonic points when two of the eigenvalues

approach zero (u1±, or u2±. Also, this scheme can be inaccurate in boundary layers

where one eigenvalue (u1, or u2) approaches zero and the others approach each other in
102
absolute value (±). These difficulties can be addressed by using the exact Riemann

solver near these regions, and using the approximate Riemann solver of Steger and

Warming away from these regions.

3.2.3 MUSCL-Hancock Scheme for Second-Order Fluxes

In the course of this research program it was determined that the first-order

upwind Riemann solvers provide inaccurate solutions near stagnation points, such as the

typical case shown in Figure 3.6. In order to resolve this inaccuracy a second-order

scheme of the MUSCL-Hancock (Monotonic Upstream-centered Scheme for

Conservation Laws) variety was implemented. This implementation follows Toro86, with

the exception of the choice of variables: Toro uses the conservative variables, but for this

case the primitive variables (Eq. 3.2.2) were required. This particular method can be

used with any appropriate Riemann Solver. The details are provided for the fluxes in the

x1 direction, F, but a similar procedure is used for G.

The first step of the MUSCL-Hancock Scheme is to assume that instead of a

constant value within the finite-volume, each variable has a value that can vary according

to a piecewise-linear function:

1
W 1   Win, j  ΔW , (3.2.28)
2

where W is the vector containing the slopes of the primitive variables, defined by:

Δ W i  1 ,j  Δ W i  1 ,j
ΔW  2 2
2 . (3.2.29)
Δ W i  1 ,j  Win, j  Win1, j Δ W i  1 ,j  Win1, j  Win, j
2 2

In order to make this scheme TVD (total variant diminishing), this slope is modified as:

103

max 

 0, min  Δ W i  1 , j , Δ W i  1 , j , 
, Δ W i  1 , j  0
 
2 2

 

min Δ  ,  Δ  

2

 
W i 1 , j W i 1 , j
ΔW   2 2
, (3.2.30)
    
 0, max  Δ W i  1 , j , Δ W i  1 , j ,
 , Δ 
  W i 1 , j  0
2 2
 min
 max Δ W  1 ,  Δ W  1 

2
 i , j
2
i , j
2 

where = 1 is equivalent to the minmod limiter, and = 2 is equivalent to the superbee

limiter. In this implementation  = 1 provides the best results. Once the slopes have

been determined, W is found at the faces:

ΔW ΔW
WiR, j  Wi , j  WiL, j  Wi , j  , (3.2.31)
2 2

where R represents the right face (i+1/2, j) and L represents the left face (i-1/2, j). The

conservative variables, UiR, ,jL , at these faces are then calculated.

Next, the values at the cell faces are evolved by a half-time step:

~
U iL, j  U iL, j  

L
 
t  Fiv Wi , j  Fiv Wi , j
R
   R  
iv i , j 
2  xh1 i , j 
 

, (3.2.32)
~
U iR, j  U iR, j  
 L
 
t  Fiv Wi , j  Fiv Wi , j
R
 
  R iv  

2  xh1 i , j 

i, j


where the inviscid fluxes are calculated according to their definition (Eq. 3.2.1). The

primitive variables WiR, j, L are then calculated.

Finally, a Riemann problem is set up at the cell faces with:


~ ~
WL  WiR, j WR  WiL1, j . (3.2.33)

Either one of the previous described Riemann solvers can be used to calculate the flux at

the cell faces.

104
1st Order
2nd Order
4500

4400

Temperature (K)
4300

4200

4100

4000 -4 -3 -2
10 10 10
X

Figure 3.6: Comparison of near-surface temperature for 1st and 2nd Order Riemann
solvers.

3.2.4 Exact Reactive-Riemann Solver

In the standard Riemann solver, the goal is to determine the values of the flow

variables at a fixed point in space at a given time, which in the current application is the

overall finite-volume face (i+½ , j etc.). The reactive-Riemann solver for this application

differs in that the information is desired only along one of the waves, the ablation front,

(depicted in Fig 3.5b), instead of the value at a fixed point.

A Reactive-Riemann solver for evaporation fronts has been presented in 87, and a

more detailed analysis of a generic Reactive-Riemann problem is given by Ben-Artzi and

Falcovitz 88. In both of these cases it has been assumed that both sides of the interface

contain compressible fluids, but in the non-charring ablation problem, one side of the

interface contains an incompressible solid.

105
As with the standard Riemann solver, the value of one of the flux vectors Fgas_iv or

Ggas_iv is desired at the chosen interface. The derivation will be given for the solver for

Fgas_iv only, but the same procedure can be implemented on Ggas_iv.

Consider that the left side of an interface is occupied by solid material with a

vector of unknowns (Usolid)L, and the right side is occupied by gas with the vector of

unknowns (Ugas)R. The solid undergoes some set of reactions which cause mass to leave

~ . The values of primitive variables of the


with a mass flux rate in the x1 direction of mabl,1

gas at the surface are desired:

W 
gas surf  
 c1  cs  cNsgas  gas uinj,1 p T .
T
(3.2.34)

The determination of these values starts with the surface mass balance equations

developed in Section 2.4. The total ablation mass flux rate is equal to the sum of the

individual mass flux rates (Eq. 2.4.32):

N
~  m
m abl ~s .
s 1

The species mass flux rates can be written in a general form as:

~ m
m ~  , c, T , (3.2.35a)
1,s 1,s gas w

where c is the vector of Nsgas species concentrations. It can then be seen that the overall

ablation mass flux rate can be written as a function of the concentrations, mixture density,

and wall temperature:

~  , c, T   m
 ~1,s  gas, c, Tw .
N
mabl gas w (3.2.35b)
s 1

The diffusion coefficient Ds (Eq 2.4.24) can be written in terms of the concentration and

binary diffusion coefficients Ds,r using Eq. 2.4.7:


106
  
1  c s  M  
  
  Ms  
Ds  . (3.2.36)
M
cr  
 Mr 
N


r 1 Ds ,r
rs

The mass balance for each species in x1 (Eq. 2.4. 40) can be rewritten as:

  
1  c s  M  
   N ~  gas   M 
 s   cs ~
1  gas cs  m  , c, T    m1,s  gas , c, Tw  . (3.2.37)
  solid  s 1  M  x1
1,s gas w
h1
cr  
 Mr 
N


r 1 Ds ,r
rs

Equation 3.2.37 contains a partial derivative of species concentration with respect to x1,

which must be discretized, using relations similar to Appendix C.1. For example, in the

case of a finite volume containing a mixture of solid and gas (i.e 0 <  < 1) , the derivative

at face (i+w, j) is:

cs i , j  cs i  w, j
 
, (3.2.38)
1
x1 i  12, j  x1 i  w, j  x1 i  w, j
2

which when substituted into Eq. 3.2.37 yields:

 
  
  
N
1  gas i  w, j c  ~  , c, T
m
  solid 
s i  w, j 1,s gas w i  w, j
  s 1
  M 
1  c s     
 
 gas i  w, j 


i  w, j 
 Ms   
 cs i, j  cs i  w, j m 
~  , c, T i  w, j . (3.2.39)
  1,s gas w
 M   1  x  
   1 i  1 , j  x1 i  w, j   x1 i  w, j
h1 
N c s i  w, j 
 
M  2  2  

r 1

 r

Ds , r i . j
rs

This equation represents a system of Nsgas equations. The values of the solid and the gas

variables away from the surface are already known. This leaves the values of the

107
concentration vector (ci+w,j) and the mixture density at the surface as unknowns, which

means there is one more unknown than equation.

The density of the mixture at the wall is usually assumed to be the same as the

density of the overall finite-volume (see for example 79). For low ablation rates, this

would probably be a safe assumption. However, at rates of large mass flux rate, this

assumption starts to break down. A possible solution can be proposed by the previous

mentioned requirement that the injection velocity remain less than or equal to the local

speed of sound (see Section 2.4.5). For the purposes of this research, it is proposed that

the wall density be found in the following way:

 gas i , j if u1,inj calc   i , j


 
gas i  w, j
 ~
 m abl,1
if u     . (3.2.40)
   1, inj calc i , j
 i, j

Now that an equation for wall density is found, the resulting system of equations

must be solved. For the purpose of solution, this system (Eq. 3.2.39) can be rewritten as:

f c  0 . (3.2.41)

The method that will be employed is to use Newton’s method for non-linear systems89,

where the following equations are iterated until some convergence criteria is met:

J f c k δ   f c k 
, (3.2.42)
c k 1  c k  δ

where Jf is the Jacobian defined by:

f c 
Jf  , (3.2.43)
c

and the vector  is found by multiplying the inverse of Jf and the right hand side, which

will be handled using the Gauss-Seidel method which is provided in Appendix C.2.
108
Once the concentration, mixture density, and ablative mass flux rate are known,

the rest of the values of Wgas surf can be found. The injection velocity is found by Eq.

2.4.33:
~
m
u1,inj  abl,1

 
gas i  w, j
,

and the surface velocity, which is the ablation wave velocity is provided by Eq. 2.4.34:
~
m
u1, surf   abl,1
.
 solid

This leaves the pressure and the temperature to be solved for.

A solution has been determined for the concentration of each species at the

surface. It is assumed that the temperature of the gas at the surface is equal to the

temperature of the gas at the nearest point. From this and the density found by 3.2.40 the

partial pressures of each species can be found using Eq. 2.4.17 and then the mixture

pressure can be found using Eq. 2.4.11 (Dalton’s Law). These values are then used to

calculate the (Fgas_iv)i+w,j using Eq. 2.3.2c.

It should be noted that this reactive-Riemann solver contains mass diffusion terms

which account for the loss of particles of a species that is involved in the surface

chemical reactions while an ordinary Riemann solver treats diffusion terms as sources 88.

In ordinary Riemann problems, all of the wave forms arise from inviscid fluid dynamics.

In the non-charring ablation problem the diffusive terms are required because the ablation

wave behavior is directly linked to the diffusion of the reacting species to the surface.

Finally, if the ablative mass flux is zero, the reactive-Riemann solver provides the same

conditions at the wall as would be expected for solid wall boundary conditions.

109
3.3 Calculation of Remaining Fluxes

The solution of the governing equations requires the evaluations of other fluxes

on the finite-volume faces, including the viscous/heat transfer fluxes (_vh), mass diffusion

fluxes (_md), solid fluxes (_solid) and surface fluxes (_surf). The methods of calculating

these fluxes are included in this section.

3.3.1 Viscous/Heat Transfer Flux

The viscous and heat transfer fluxes, Fgas_vh, and Ggas_vh are given by: 2.3.2c-d as:

 0   0 
    . 
   
 0   0 
   
     
Fgas_ vh   0 , and G gas_ vh  0 ,
   
   11     12 
     
 12
  22

 T 
k  k T 
 h1 x1  
h2 x2 

G gas_ vh

where the shear stress terms 11, 22, and 12 can be written using Eq. 2.1.13b and 2.2.12:

2   1 u u h   1 u u h   u h u h  
 11    2 1
 2 1
   2
 2 2
   2 3
 1 3 

3   h1 x1 h1h2 x2   h2 x2 h1h2 x1   h2 h3 x2 h1h3 x1  
2   1 u u h   1 u u h   u h u h  
 22    2 2
 2 2
 1
 2 1
 2 3
 1 3 
 .(3.3.1)
3   h2 x2 h1h2 x1   h1 x1 h1h2 x2   h2 h3 x2 h1h3 x1  
 h1   u1  h2   u2  
 12         

 h2 x2  h1  h1 x1  h2  

The derivatives for the velocities, temperature and metrics (hi’s) are discretized using the

appropriate formula found in Appendix C.1. In general, these variables can be

interpolated to the face values using simple linear averaging. The exception is for surface
110
faces (i.e i±w, j ) where the temperature at the interface is taken as the solid temperature,

and the velocities are provided by the reactive-Riemann solver.

The viscosity  and thermal conductivity k are taken for the finite-volume for

which the flux is being calculated, with the exception of the surface faces, where the

thermal conductivity is found by 90:

1
ksurf  . (3.3.2)
1 1

k gas k solid

3.3.2 Mass Diffusion Fluxes

The mass diffusion fluxes Fgas_md and Ggas_md are given in Eq. 2.3.2c-d:

 1u1,1   1u2,1 
 .   . 
   
  s u1, s    s u 2, s 
   
 .    
Fgas_ md   Nsgasu1, Nsgas , and G gas_ md    Nsgasu2, Nsgas ,
   
 0   0 
 0   0 
N  N 
  s eT  s   ps u1s    s eT  s   ps u2 s 
 s 1   s 1


G gas_ md

where the diffusion velocities for species s, u1,s and u2,s, can be found in terms of the

concentration gradient using Eq. 2.4.19:

 cs
 s u1, s   Ds
h1 x1
.
c
 s u 2, s   Ds s
h2 x2

111
The spatial derivatives of concentration can be discretized using the formulas contained

in Appendix C.1. The species mass diffusion coefficient Ds, and mixture density r are

taken for the finite volume under consideration. The concentrations at the finite-volume

faces are found by a linear averaging between neighboring finite volumes, with the

exception of the solid surfaces, where the concentration is found using the reactive-

Riemann solver.

3.3.3 Solid and Surface Fluxes

The solid fluxes Fsolid and Gsolid are given by 2.3.1c-d:

k T 
 eT u1, surf  
Fsolid 
 u    h1 x1  ,
 1, surf   0 

and

k T 
 eT u2, surf  
G solid  
 u   h2 x2  .

 2 , surf   0 

The first set of terms is only needed for calculation of the solid flux at the surface, and

the individual variables are calculated directly from the results of the reactive-Riemann

solver. The conductive heat flux is calculated in two ways. For an interior face, the

temperatures are interpolated using a linear averaging, and the derivative is approximated

using the discretizations of Appendix C.1. However for a surface face, the heat flux is

found using the surface energy balance (Eq. 2.4.58), for example for face i+w ,j:

112
 k T   k T   p  D c ~   p 
Nsgas
         eT ,s  s  s s  m abl,1  
 eT    eT solid  , (3.3.3)
 h1 x1  solid  h1 x1  gas s 1   s  h` x1    abl 

where again k should be calculated using Eq. 3.3.2.

The surface fluxes Fsurf and Gsurf are calculated by substituting the results of the

reactive-Riemann solver into Eq. 2.2.2e:

 1u1, surf   1u2, surf 


 .   . 
   
  s u1, surf    s u2, surf 
   
     .
Fsurf  , G surf 
  Ns u1, surf    Ns u2, surf 
 gas   gas 
 0   0 
 0   0 
   
 eT u1, surf   eT u2, surf 

3.4 Calculation of Source Terms

The governing system of equations contains geometric and physical source terms

that must be calculated. The values of these terms must be calculated for each finite

volume (i, j) instead of each face.

3.4.1 Geometrical Source Terms

The geometrical source terms are represented by R in the governing vector

equations, with individual components defined by Eq. 2.3.1e, and 2.3.2g-i. These contain

the derivatives of the coordinate system metrics (hi’s):

h2 h1 h2 h3  h1h3  h3 h3


, , , , , and ,
x1 x2 x1 x2 x1 x2

113
which are discretized using the appropriate formula of Appendix C.1. The other

variables such as densities, velocities, etc are taken at their value for the entire finite

volume. Depending on the location and extent of the solid, an averaging procedure such

as the one used for averaging the fluxes in Section 3.1.4 may be necessary for finite

volumes that contain both solid and gas.

3.4.2 Dissipation Terms

The physical source terms S contain the dissipation function , which can be

defined using Eq. 2.2.12 and 2.2.16:

   1 u  
2
   
2
   
2

2  
u h 1 u u h u h u h
1
 2 1
   2
 2 2
   2 3
 1 3

   h1 x1 h1h2 x2   h2 x2 h1h2 x1   h2 h3 x2 h1h3 x1   
 
  h   u  h   u  2 
     1  1   2  2    , (3.4.1)
  h2 x2  h1  h1 x1  h2   
 2

 2   1 u1 u2 h1   1 u2 u2 h2   u2 h3 u1 h3   
              

 3  h 
 1 1 x h h  x 2   2 h x h h  x1   2 3 h h  x h h  x1  
 
1 2 2 1 2 2 1 3

which can be discretized using the appropriate forms of the equations in Appendix C.1.

Averaging based on the solid location and size might be needed for this term as well.

3.4.3 Gas Phase Chemical Reactions

The source term S also contains terms resulting from the volumetric mass

production rates w s , which are given by Eq. 2.4.27:

 
w s  Ms   s ,r   s ,r R f ,r  Rb,r  ,
Nr

r 1

114
where s,r is the stoichiometric coefficient of species s on the product side of reaction r,

s,r is the stoichiometric coefficient of species s on the reactant side of reaction r, Rf,r is

the forward reaction rate for reaction r, and Rb,r is the backward reaction rate for reaction

r. The set of w s equations actually forms a system of ordinary differential equations.

As an example consider the net production rate of atomic oxygen within the 10

species model proposed in Appendix C.2. By defining Rnet,r as:

Rnet,r  R f ,r  Rb,r ,

Eq. 2.4.27 can be written for atomic oxygen as:

2 RNet ,1  2 RNet , 2  2 RNet ,3  2 RNet , 4  2 RNet ,5 


 
  2 RNet ,6  2 RNet , 7  2 RNet ,8  2 RNet ,9  RNet ,19 ,
w O  MO   (3.4.2)
R  RNet , 21  RNet , 22  RNet , 23  RNet , 24
 Net , 20 
 RNet , 25  RNet , 26  RNet , 27  RNet , 28  RNet , 29  RNet ,30 

where the individual net reaction rates are given by:

  
2
    
O   
2

Rnet,1  1000 k f ,1  .001 2    kb,1  .001  O   .001 O2  . (3.4.3a)
   MO2     MO   MO 2  
     

  O        
2
  
Rnet, 2  1000 k f , 2  .001 2  .001 N 2   kb, 2  .001

O
  .001 N 2  , (3.4.3b)
  MO 2  MN 2   MO   MN 2  
       

  O         
3

Rnet,3  1000 k f ,3  .001 2  .001 O


  kb,3  .001 O
  , (3.4.3c)
  MO 2  M  M 
   O  
   O   

  O      
2
   N  
Rnet, 4  1000 k f , 4  .001 2  .001 N
    k  .001 O
  .001  , (3.4.3d)
  MO2  M 

b,4 
M   M 
   N  
   O   N  

115
  O    
2
   
Rnet,5  1000 k f ,5  .001 2  .001  NO   kb,5  .001  O   .001  NO   , (3.4.3e)
  MO 2  MNO    MO   MNO  
    

  O       
2
  
Rnet,6  1000 k f ,6  .001 2  .001 NO    kb,6  .001  O   .001 NO   , (3.4.3f)
  MO 2  MNO     
MO   MNO   
       

  O    
2
   
Rnet,7  1000 k f ,7  .001 2  .001  CO   kb,7  .001  O   .001  CO   , (3.4.3g)
  MO 2  MCO    MO   MCO  
    

  O    
2
   
Rnet,8  1000 k f ,8  .001 2  .001  CN   kb,8  .001  O   .001  CN   , (3.4.3h)
  MO 2  MCN    MO   MCN  
    

  O     
2
   
Rnet,9  1000 k f ,9  .001 2  .001 C3   kb,9  .001  O   .001 C3  , (3.4.3i)
  MO2  MC3    MO   MC3  
      

    N      
Rnet, 28  1000 k f , 28  .001 O  .001 2   kb, 28  .001  N  .001  NO   , (3.4.3j)
 MO  MN 2   MN  MNO  
     

          O  
Rnet, 29  1000 k f , 29  .001 NO  .001 O   kb, 29  .001 N  .001 2  , (3.4.3k)
  MNO  MO    MN  MO 2  
  

and

             
Rnet,30  1000 k f ,30  .001 N  .001 O   kb,30  .001 NO  .001 e   . (3.4.3l)
 MN  MO    MNO   Me-  
     

The production rate is calculated by substituting the values of (s)i,j into the appropriate

net reaction rate equations, and using Ti,j to determine the rate coefficients.

The inclusion of these ordinary differential equations can introduce problems with

the solution of the overall system because the time step requirement for the chemical

reactions might be significantly less than the time steps allowed for the other processes,
116
which is called stiffness 85. This is the reason why the semi-implicit scheme of Eq. 3.1.10

may be required.

3.5 Time Step Calculation

In order to proceed with time-marching calculations using either Eq. 3.1.9 or

3.1.10, a method for determining the time step t is required. There are various stability

and accuracy limits for t arising from the individual processes, such as the inviscid,

viscous, heat transfer, mass diffusion, gas-phase chemical reaction and surface reactions.

The current solver allows for varying grid spacing in both the x1 and x2 direction, so the

formulas should be modified to account for this. For the inviscid gas dynamics, the time

step is given by 74:

 x1* x2* 
tiv  min  , , (3.5.1)
imin i imax
jmin  j  jmax  u1   u 2   

where x1* and x2* are the distances between face for a given computational cells (so a

finite volume with a mixture of solid and gas may have up to 4 of each to compare). For

the viscous gas dynamics, the time step is given by91:

1  x12  x22 
tvisc  min   , (3.5.2)
4 ijmin i j imaxj   
min max

where  =  / The time step for conductive heat transfer is found similarly by 90:

1  x12  x22 
tcond  min  , (3.5.3)
4 ijmin i j imaxj   
max  
diff
min

where diff = k/(Cp). The chemical reaction characteristic time is more complicated

because it must be taken for Nr reactions at every point 91, 85 :


117
  1

 .001 s     
 s ,r
   Nsgas

tchem  min  min  1000k f ,r T        , (3.5.4)


imin i imax
jmin  j  jmax 
r 1, Nr 
   M     
   s 1 s


using the terminology developed in Section 2.3.4. Finally, there is a time step

dependence on the movement of the surface:

 x1 i  w, j  x1 i  w, j x2 i , j  w  x1 i , j w 


t surf  min 
imin i imax  u       
, , (3.5.5)
 u u  u 
jmin  j  jmax  1, surf i  w, j 1, surf i  w, j 2 , surf i, j w 2 , surf i, j w 

which comes from the fact that the solid volume cannot become negative. The overall

time step is then found by:

t  CFL * min tiv , tvisc , tcond , tchem , tsurf  , (3.5.6)

where the CFL (Courant-Freidrichs-Levy) number is a factor less than one (usually taken

as less than or equal to 0.8). Depending on the particular flow parameters the CFL

number may need to be varied to ensure stability.

3.6 Numerical Boundary Conditions

A brief description of appropriate boundary conditions was presented in Section

2.4.7. The definition of conditions will be for a mesh that has imax points in the x1

direction and jmax points in the x2 direction. The use of second order inviscid flux

approximations requires the use of two ghost volumes on each boundary, so the actual

grid has a range for i of -1 to imax+2, and for j of -1 to jmax+2. The boundary conditions

will be presented for a typical body-fitted grid such as the one shown in Figure 3.7.

118
Figure 3.7 Boundary conditions and mesh for a typical problem.

For the inlet boundary condition located at jmax, the two ghost volumes are set to

the constant free stream condition as is appropriate for a supersonic inlet:

U 
gas i , j 1
max
 U gas i , j
max  2
 U gas freestream . (3.6.1)

It should be noted that the inlet velocity which is typically given as u∞ in Cartesian

coordinates must be transformed into the appropriate u1 and u2 components in the body-

fitted coordinate system at each point.

119
The gas boundary condition at imax is a supersonic outlet given by:

U gas i  2, j
max
 U gas i
max 1, j
 U gas i
max , j
. (3.6.2)

The solid temperature boundary condition is defined in the same manner.

The symmetry boundary condition located at i=1 given by Eq. 2.4.60 can be

numerically approximated by 86:

 1 0, j   1 1, j   1 1, j   1 2, j 


           
       
  s 0, j    s 1, j    s 1, j    s 2, j 
       
         
  N 0, j    N 1, j ,   N 1, j    N 2, j  . (3.6.3)
       
 u1 0, j   u1 1, j   u1 1, j   u1 2, j 
u    u   u    u  
 2 0, j   2 1, j
  2 1, j   2 2, j

eT 0, j   eT 1, j  eT 1, j   eT 2, j 

For the solid boundary condition located at j=1, several boundary conditions can

be implemented. The simplest is a constant temperature boundary condition where the

two ghost volumes are set to the specified temperature. An insulated boundary condition

can be set to:

T
 0  Ti ,0  Ti , 1  Ti ,1 . (3.6.4)
x2

3.7 Algorithms

The numerical methods that have been detailed must be integrated into an

algorithm in order to solve the non-charring ablation problem. First, a serial algorithm

will be outlined that can be implemented on a single processor computer. Second, the

120
details of a parallel algorithm that can be implemented on a cluster computer with

multiple processors will be provided.

3.7.1 Serial Algorithm

The serial algorithm begins with the values of (Ugas)n and (Usolid)n for each point (i,

j) that is not on a boundary. From these values, primitive variables must be calculated at

each point including the species concentrations, mixture density, velocity components u1

and u2, pressure, solid and gas temperatures, values for  and , and the speed of sound .

Once these values are found linear averaging is employed to interpolate values to the

finite volume faces. Next, the transport properties are calculated for each point using the

relations of Section 2.4.4.

The fluxes and source terms are then calculated. First, the reactive-Riemann

solver detailed in 3.2.3 is implemented to determine the value of the inviscid fluxes at the

surface and to provide the surface velocities (u1,surf, u2,surf), and the primitive variables on

the surface faces. Second, the appropriate Riemann solver, either the exact (Section

3.2.1) or Steger-Warming (Section 3.2.2) is chosen for each point based on the

eigenvalues; the exact solver is needed if the velocity is near the speed of sound, or the

velocity is near zero, which in general means the exact solver is used in the boundary

layer, and possibly in small regions of the shock layer. Then the other fluxes are solved

in accordance with Section 3.3. Finally the source terms are calculated using the methods

of Section 3.4.

121
Throughout the previous calculations, the time steps are being calculated as per

Eq. 3.5.1-6. The new location of the surfaces must be found, for example:

x1 inw1, j  x1 in w, j  t u1,surf in w, j . (3.7.1)

The change in volume ratio  with time is found by:

 1
t 4
 n n
 n n

 u1, surf i  w, j  u1, surf i  w, j u2,surf i , j  w  u2,surf i , j  w . (3.7.2)

Once these values and all of the fluxes, sources and the time step are calculated, the new

values of (Ugas)n+1 and (Usolid)n+1 are found using 3.1.9a-b. The boundary conditions are

applied to the boundary points, and then the process is repeated until some final time is

reached.

3.7.2 Parallel Algorithm

The small time steps required by an explicit scheme and number of calculations

involved in each time step cause the serial algorithm running on a single processor to be

computationally expensive; even on a fast desktop a one millisecond of simulated time

can take on the order of days of physical time to complete. In order to achieve

meaningful results in a shorter amount of time, it is necessary to modify the serial

algorithm to one that takes advantage of the scaling of cluster computing.

Consider that the problem is to be solved on a grid that has N points in the x1

direction and M points in the x2 direction. The simplest way to decompose this domain

for solution on P processors is to divide it into sub-domains in the x1 direction that have

N/P points in the x1 direction and points in the x2 direction. Figure 3.8 shows this grid

122
decomposition for a body-intrinsic coordinate system (with each color representing the

domain handled by a different processor) using 8 processors.

Figure 3.8: Domain decomposition for 8 processors.

A message passing algorithm can then be implemented to transfer information

between the various processors. Once the grid has been decomposed the serial algorithm

is followed within each sub-domain until the time step is calculated. Each processor

must send a message containing their individual time step to a processor that has been

designated as the controller processor (see Fig. 3.9a). This processor then determines the

minimum of all the time steps and then sends a message to all of the other processors

containing this new time step. Then each processor determines its values of (Ugas)n+1 and

(Usolid)n+1. Then each processor must send its data to its neighboring processors, as

shown in Fig. 3.9b. Details of this algorithm can be found in92 .

123
This algorithm was implemented using MPI and FORTRAN90. All

computational runs were completed on the Ohio Supercomputer Center Glenn cluster93.

A typical node on this cluster has four processor cores. Figure 3.10 shows the typical

performance of the algorithm for the fourth test case which will be discussed in the next

chapter. Speedup is the ratio of the computational time required for a given number of

processors to some reference number. Efficiency is the ratio of the actual speedup to the

theoretical speedup (ratio of number of processors to reference number). In this case the

reference number is four processors which corresponds to a single typical node of the

Glenn system. A best practice for parallel computing is to increase the number of

processors until the efficiency drops below fifty percent. For test case four this occurs at

44 processors, so 40 processors (10 nodes) are used to obtain the results.

Left Communicates to Right


0
0 U imax(0),j 1 U imax(1),j
2 U imax(2),j
3
t
1
t

tmin

after
t2

left-to-right
t
3
is complete
in

then
m

m
t

in

0 1 2 3
1 2 3 U imin(1),j U imin(2),j U imax(3),j

Right Communicates to Left

(a) (b)
Figure 3.9: Parallel Communication Schemes. (a) Time Step
(b) Sub-domain Boundary Data.

124
1.2

1
4
speedup

efficiency
efficiency

speedup
0.8
3

0.6
2

0.4 1
10 20 30 40 50
# of processors

Figure 3.10: Speedup and efficiency vs. number of processors.

3.8 Summary

This chapter has provided the numerical methods that are proposed to solve the

governing equations of the non-charring ablation problem that were given in Chapter 2.

These numerical methods when combined into an algorithm should provide a means of

tracking the time-dependent solution of this problem.

The discretization schemes for time and for space operators that have been

provided are based on standard procedures, but have been modified to handle finite-

volumes that contain a mixture of solid and gas. These modifications were needed in

order to handle arbitrary location and volume ratios of the solid. A first-order explicit in

time (Eq. 3.1.9a-b) scheme have been provided that accounts for the movement of the

solid faces.

125
An exact Riemann solver and an approximate Riemann solver (Steger-Warming

Flux Splitting) are used to calculate the inviscid fluxes at gas faces. They have been

modified to deal with the multiple-species and high temperature gas dynamics that are

associated with this problem. The exact solver is needed for regions where the flow is

either transonic or near zero due to the inaccuracy of the Steger-Warming method at these

areas. The extension of these schemes to second order has also been given, as the use of

either first order scheme is inappropriate in the stagnation region.

A reactive-Riemann solver has been developed in this work that provides the

inviscid gas fluxes at the surface of the solid, and also accounts for the surface recession

velocity. This solver consists of solving the mass balance equation for the surface

chemical reactions, and then calculating the other variable necessary for the flux. It

should be noted that this solver only provides the value on the gas side of the ablation

front, whereas the exact and approximate Riemann solvers provide the value at a

particular point in space.

The methods for calculating the diffusive fluxes within the gas flow require the

discretization schemes given in Appendix C.1 for primitive variables. These schemes are

modified from typical forms to account for the possibility of a mixture of solid and gas

within a finite-volume. The variables are interpolated to the faces from the discrete (i,j)

value using linear averaging.

The solid and surface fluxes depend on the surface recession velocity determined

by the reactive-Riemann solver. At the surface, the heat flux to the solid is found using

the surface energy balance.

126
The source terms can be separated into geometrical and physical sources. The

geometrical sources are solved using a similar procedure to the diffusive fluxes, with an

averaging procedure such as the one used in calculating the flux derivatives being applied

when necessary. The chemical source terms require solution of a system of ordinary

differential equations.

The time step requirements for the overall solution depend on the values of the

time step limitation for each separate process. If one of these individual time steps is

much smaller than the others (usually from the chemical reactions), a computational issue

known as stiffness is introduced; a semi-implicit time scheme is provided in Eq. 3.1.10

can be used to alleviate this issue. However in the test cases considered in the next

chapter this problem does not occur due to the small grid spacing needed for adequate

resolution of the boundary layer, so only a first order explicit time is required.

A serial algorithm has been presented to show how the various numerical methods

can be combined to achieve the solution of the non-charring ablation problem. The serial

algorithm is of limited practical use because meaningful simulation times take excessive

amounts of computational time on a single processor. Parallel computing offers a way to

decrease the computational time required for a given simulation time. A parallel

algorithm is provided that divides the solution domain into several (one per processor)

sub-domains on which the serial algorithm is implemented. This parallel algorithm sends

messages between the various processors to communicate the time step and results at

each time step.

127
CHAPTER IV

RESULTS AND DISCUSSION

In this chapter the computational methods proposed and presented in the previous

chapter will be tested. First, separate parts of the computational methodology will be

validated against test cases. Second, results from a simplified model with a stair-case

representation of the surface will be presented to illustrate the necessity of the inclusion

of a body fitted grid and more precise chemistry and gas dynamics. Third, several

ablative test cases in dissociated air will be presented including enhanced ablation in the

presence of a surface defect and compared to appropriate experimental results where

possible.

4.1 Validation

The developed computational method couples solver for many different physical

processes such as shock wave gas dynamics, boundary layers, cavity flow and conduction

heat transfer. In the development of the solvers for these various phenomena it was

useful to test them individually and in concert against established test cases.

128
4.1.1 One-Dimensional Inviscid Gas Dynamics

A simple but important test case for any CFD code that employs Riemann solvers

for inviscid gas dynamics is the Sod problem86. The Sod problem is essentially a model

of a shock tube with a high pressure and density gas on one side of an interface at the

initial time moment, and a low pressure and density gas on the other side. For testing the

exact and approximate Riemann solvers employed in this research, the domain was taken

as one meter in length with the interface at 0.5 meters from the origin. The initial

conditions for each side of the interface are given in Table 4.1. An analytical solution at a

final time of 7.5*10-4 seconds was determined for this problem using the techniques

given in86. This solution consists of a rarefaction fan expanding to the left of the initial

interface, and a shock and contact surface moving to the right (see for example Figure

4.1). Each Riemann solver was used to solve the problem with various number of grid

points involved.

Table 4.1: Initial conditions for the Sod Problem.


Left of the membrane Right of the
Variable
(x1<0.5 m) membrane (x1>0.5 m)
(kg/m )
3
1.000 0.125
u1 (m/s) 0.000 0.000
p (Pa) 101325.000 10132.500

129
1

Analytical
100 points
200 250
0.8
400
800
200
0.6
 (kg/m3)

u1 (m/s)
150

0.4
100 Analytical
100 points
200
0.2 400
50 800

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x1 (m) x1 (m)

100000

Analytical
80000 100 points
200
400
800
60000
p (Pa)

40000

20000

0
0 0.2 0.4 0.6 0.8 1
x1 (m)

Figure 4.1: Results of Grid Refinement for Exact Riemann Solver.

Figure 4.1 shows the density, velocity and pressure profiles for the exact Riemann

solver developed in Section 3.2.1. As the number of grid points increases, the solution

converges to the analytical solution. The solutions at the shock and along the rarefaction

have converged much faster than the solution at the contact surface (middle discontinuity

in the density plot).

130
1 300

0.9
Analytical 250
0.8
100 Points
200
0.7 400 200
800
0.6
 (kg/m3)

u1 (m/s)
0.5 150 Analytical
100 Points
0.4 200
400
100
0.3 800

0.2
50
0.1

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x1 (m) x1 (m)

100000

Analytical
100 Points
80000 200
400
800

60000
p (Pa)

40000

20000

0
0 0.2 0.4 0.6 0.8 1
x1 (m)

Figure 4.2: Results of Grid Refinement for the Steger-Warming Riemann Solver.

Figure 4.2 shows the same profiles for the Steger-Warming approximate Riemann

solver developed in Section 3.2.2. Again as the number of grid points increases, the

solution converges to its analytical value. The contact discontinuity is again less

converged than the other waves for the same amount of grid points.

131
1 300

0.9 Analytical
Exact Riemann Solver 250
0.8 Steger-Warming Solver

0.7
200
0.6
 (kg/m3)

u1 (m/s)
0.5 150

0.4
100
0.3

0.2
50
0.1

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x1 x1

100000

Analytical
80000 Exact Riemann Solver
Steger-Warming Solver

60000
p (Pa)

40000

20000

0
0 0.2 0.4 0.6 0.8 1
x1

Figure 4.3: Comparison between Riemann solvers.

Figure 4.3 shows a comparison between the solution obtained by each Riemann

solver with 800 grid points and the analytical solution. The performance of each solver is

nearly identical with the exception of the right edge of the rarefaction, where the exact

solver is converging faster than the Steger-Warming solver.

132
1 300

Analytical 250
0.8 S-W 1st order
S-W 2nd order

200
0.6
 (kg/m )

u1 (m/s)
3

150

0.4
100
Analytical
S-W 1st order
S-W 2nd order
0.2
50

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x1 x1

100000

Analytical
80000 S-W 1st order
S-W 2nd order

60000
p (Pa)

40000

20000

0
0 0.2 0.4 0.6 0.8 1
x1

Figure 4.4: Comparison between 1st and 2nd order Riemann solvers.

The use of the second order Riemann solver allows a much more accurate

solution for the same number of points. Figure 4.4 shows a comparison between the first

and second order Steger-Warming solver for 200 points. The second order solver

reproduces the tail and head of the rarefaction wave almost exactly, and gets much closer

to reproducing the contact discontinuity.

For one-dimensional gas dynamics, the Riemann solvers that are implemented in

this study can be seen to reproduce the solution of the Sod problem. They converge to

133
the analytical solution with increasing numbers of grid points. In each solver the contact

discontinuity requires more grid refinement to converge than the shock or the rarefaction

fan. Implementation of the second order solver allows a better solution to be obtained

with a much coarser grid. For this particular case, the approximate Riemann solver of

Steger and Warming can be used without significant decrease in accuracy.

4.1.2 Supersonic Flow over a Wedge

In order to test the two-dimensional gas dynamics and the solid wall boundary

condition provided by the reactive-Riemann solver the supersonic flow over a wedge was

used as a test case. The particular case chosen was a Mach 3 air flow over a 30° wedge.

Figure 4.5 shows the pressure across the shock wave and the shock angle using the

Steger-Warming solver and a 50 x 50 grid. The numerical results are almost identically

equal to the theoretical values94 for the pressure after the shock wave (644.01*103 Pa),

and the theoretical angle (52.014°). The numerical results provide the Mach number after

the shock of 1.405, whereas theoretical value is 1.406. Figure 4.6 compares the results

of the Exact and Steger-Warming Riemann solvers.

134
P
750000
700000
650000
1.2 600000
550000
500000
450000
1 400000
350000
300000
3 250000
0.8 p1=101.325 *10 Pa
200000
150000
Y
0.6 3
p2=644.150*10 Pa

0.4
 = 52
o

0.2
 = 30 o

0
-0.4 -0.2 0 0.2 0.4 0.6 0.8
X

Figure 4.5: Pressure contours for Mach 3 flow over a 30° wedge.

0.4
Exact Riemann Solver

0.2 P
650000
600000
550000
500000
Y

0 450000
400000
350000
300000
250000
-0.2 200000
150000

-0.4
Steger-Warming

0.2 0.4 0.6 0.8 1


X

Figure 4.6: Comparison of Riemann solvers for Mach 3 flow over a 30° wedge.

135
4.1.3 Flat Plate Boundary Layer

In order to ascertain the function of the viscous flux calculation portion of the

solver, the boundary layer resulting from flow over a flat plate was used as a test case.

The numerical results for u1∞ = 100 m/s were compared with the Blasius profile95 at

various lengths along a 1 meter long plate. The exact Riemann solver was used for

inviscid portions of the flux. The solid surface was located at 2 = 0.5 within the first

finite volume. The grid was 50 points in the x1 direction, and 25 points in the x2 direction

with x2 = 1.2*10-4 m. Figure 4.7a shows the results for 0.25m, Figure 4.7b shows the

results at 0.5m, and Figure 4.7c shows the results for 0.75m. As can be seen the

numerical results follow the Blasius solution.

136
3.0x10-03 3.0x10-03

2.5x10-03 2.5x10-03

Numerical
Blasius
2.0x10-03 2.0x10-03
x2 (m)

x2 (m)
1.5x10-03 1.5x10-03

1.0x10-03 1.0x10-03

5.0x10-04 5.0x10-04

0.0x10+00 0.0x10+00
0 20 40 60 80 100 0 20 40 60 80 100
u1(m/s) u1 (m/s)

(a) (b)

3.0x10-03

2.5x10-03

2.0x10-03
x2 (m)

1.5x10-03

1.0x10-03

5.0x10-04

0.0x10+00
0 20 40 60 80 100
u1 (m/s)

(c)
Figure 4.7: Velocity Profiles for Flat Plate Boundary Layer. (a) 0.25 m (b) 0.5 m
(c) 0.75 m.

4.1.4 Driven Cavity

Another important test case for the viscous portion of the code is the lid-driven

cavity problem. In this case a square cavity has a top wall that is moving at some speed

U∞. Depending on the Reynolds (Re) number of a problem, a particular vortex pattern is

expected to form within the cavity. Ghia, Ghia and Shin96 provide results for the velocity

profiles along the vertical and horizontal center lines of the cavities for various values of

137
Re. For this study, the cavity was chosen to have a grid spacing of x1 = x2 = 2.5*10-6

m, which is typical of the resolution required for boundary layer resolution. The

temperature is specified at 300K, and the pressure as 101.325 kPa, which provides a

density of 1.17 kg/m3. Based on these properties the viscosity is determined to be

1.6967*10-5 N*s/m2. The grid was set to 134x134 with 2 solid points on each wall which

yield a gas grid of 129x129. For Re = 5000 the wall velocity is 224.83 m/s. It should be

noted that the results were normalized after the computations.

Figure 4.8(a) shows the non-dimensional horizontal velocity profile along a

vertical line passing through the geometrical center of the cavity. As the grid is refined

the solution is trending towards the solution provided by Ghia et al. Figure 4.8(b) shows

the non-dimensional velocity vectors for every tenth point in each direction. The location

of the main vortex and the presence of weaker secondary vortices near the corners are as

expected.

This test case is important because it demonstrates the ability of the model to

handle driven cavity flow which is important in the second ablation test case that will be

discussed later in this chapter (Section 4.3.2).

138
1

0.8

x2
* 0.6

0.4 Ghia et al.


129x129
358x358

0.2

0
-0.4 -0.2 0 0.2 0.4 0.6 0.8 1
*
u1

(a)

(b)
Figure 4.8: Driven cavity results for Re=5000. (a) Horizontal velocity profiles along
vertical line through geometrical center (b) Velocity vectors.

139
4.1.5 Solid Heat Transfer

The conduction heat transfer validation case is a semi-infinite solid with a

constant heat-flux at its surface. The analytical solution is90:

" t
2qwall
 exp   x   qwall x erfc x  ,
2 "
T x, t   Tinitial   4t  (4.1.1)
k   k  2 t 

"
where qwall is the heat flux at the wall,  is the thermal diffusivity, k is the thermal

conductivity, and x is the distance from the wall. Figure 4.9 shows a comparison between

the analytical solution and two numerical solutions with differing x. The initial

temperature is 300K, the diffusivity is 1.05*10-4 m2/s, and the thermal conductivity is

357.16 W/(m*K). The analytical surface temperature at a time of 10-3 s for this case is

310.24K; the coarser grid gives this temperature as 310.12K, and the more refined grid

gives 310.22K. It should be noted that the diffusivity and conductivity are taken to be

constant which differs from the complete ablation numerical model that allows these to

vary with temperature.

140
310
Analytical
x=10
-5

x=2.5*10 -6

Temperature (K)
305

300

0 0.0005 0.001 0.0015 0.002


X (m)

Figure 4.9: Numerical vs. analytical solution for conduction with a constant heat flux in
a semi-infinite solid.

4.2 Preliminary results obtained using a simplified model

In the course of researching the non-charring ablation problem, an intermediate

model was developed80, 97. This model was used to simulate several test cases, and the

results from these tests have provided direction to the enhancement of the final version of

model presented in this dissertation.

The early model defined  as a phase variable, which was either one (solid) or

zero (gas). The governing equations were written such that when  was one the equations

reduced to conduction heat transfer within the solid, and when was zero they reduced to

the Navier-Stokes equations.

141
Only ablation via sublimation was accounted for. The surface was tracked, but the

ablative flux was calculated at the original interface instead of at the actual location of the

surface. This continued until the entire finite-volume was ablated.

The ratio of specific heats, for the air and ablation product (C3) were assumed

to be constant, and gas phase chemical reactions were not accounted for.

As an example of the results from the preliminary model 80 was a two-

dimensional axisymmetric circular-tipped graphite cone with a nose radius of 2 cm,

overall length of 3 cm, and cone half-angle of 10°, which is similar to the case given by

Chen and Milos79. The cone was given an initial temperature of 3750K (to be within the

sublimation range for the entire simulation), and the air was given an initial density of

0.235 kg/m3, pressure of 101.325 kPa, and Mach number of 7.0596. The computational

grid was Cartesian, with a stair-cased region representing the solid (=1). The 160 x160

point grid was structured in such a way as to be fine (r = z = 5*10-4m) near the

surface, and coarse (r = z = 5*10-3) away from it. This refinement was an attempt to

resolve the boundary layer.

0.05
T(K)
20000
18150
16300
0.04 14450
12600
10750
8900
7050
0.03 5200
3350
R (m)

1500

0.02

0.01

0
-0.01 0 0.01 0.02 0.03
Z (m)

Figure 4.10: Temperature contours of the converged flowfield in the simplified model.
142
Figure 4.10 shows the temperature contours after the flowfield has been allowed

to converge (ablation was suppressed until this occurs). The temperature near the

stagnation point is on the order of 18000K, and in some near surface regions approached

20000K. These are unrealistic temperatures, which result from the use of constant  and

neglecting the gas phase chemical reactions. This led to the inclusion of these into the

current version of the model.

Figure 4.11 shows the concentration of the ablation product C3 at various times.

It was expected that the ablation contours would be smooth with higher concentrations

near the stagnation point and the surface. Instead, there are “plumes” of much higher

concentration that are short-lived and vary in location. An investigation into the cause of

these plumes led to several possible culprits. First, the already identified unrealistically

high temperatures cause a corresponding large initial ablative mass flux resulting in a

high concentration which suppresses the mass flux until the high concentration dissipates.

Second, in the stair-cased approach the surface is not smooth, which does not allow the

near surface gas dynamics to behave correctly; this lead directly to the inclusion of the

general orthogonal coordinates (specifically body fitted coordinate systems) in the current

model. The use of the Ytrehus model of sublimation in place of the Hertz-Knudsen

model helped to partially suppress these oscillations but they were still present.

143
Figure 4.11: Concentration of C3 at various times within the simplified preliminary
model. (a) 1 ms (b) 5 ms (c) 10 ms.

4.3 Test Cases

The ablation test cases for this dissertation consist of varying conditions on the

same basic geometry which consists of a spherical tipped cone with a nose radius of 1.95

cm, a cone angle of 10°, and an overall length of 3 cm. This is similar to the case

presented by Chen and Milos79, and is in the range of the experiments conducted by

Lundell and Dickey30 which will allow comparisons to Figure 1.5 in terms of ablative

mass flux versus surface temperature.

Two different sets of freestream conditions were considered which are shown in

Table 4.2. These cases were chosen to represent this model’s capability of handling a

vast range of initial conditions. In each of these cases the initial composition is 22% O2

144
and 78% N2. The first set represents flight at sea level conditions, which would be the

maximum freestream pressure and density that a vehicle could encounter. The second set

represents the arc-jet conditions (provided by79) under which TPS materials are usually

tested.

For each of these sets of conditions there are two distinct test cases that are meant

to illustrate the capabilities of this model. Test case 1 has a perfect surface and test case

2 has the same surface with a small defect. In Test case 3 the surface temperature is

within the range where oxidation and nitridation are expected to dominate. Test 4 has

allowed the surface to heat to the point where sublimation is expected to dominate in the

stagnation region.

Table 4.2: Freestream conditions for the varying test cases.

Case T∞(K) P∞(Pa) ∞(kg/m3) u∞(m/s) M∞


1, 2 300 101325 1.17 3305 9.5
3, 4 1428 1672 0.004 5354 7.24

4.3.1 Test Case 1: Sea level flight with defect-free surface

The first test case checks the functionality of the complete numerical method.

The numerical procedure was implemented as follows: a steady inviscid solution was

obtained with reflective boundary conditions at the TPS surface, then the viscous and

heat transfer portions of the solver is included and the temperature of the solid is allowed

to reach a suitable temperature, after which the hypersonic chemical reactions of the

partially dissociated air and mass diffusion portions are added until the concentrations

stabilize and then finally the ablation solvers including the reactive-Riemann solver and

surface energy balance are turned on. This procedure was required to decrease the
145
amount of computational time needed to reach the point where an ablation solution could

be obtained. For this particular case the viscous and heat transfer terms were run for

1.25*10-4 s of physical time and then the chemical reaction terms were run for an

additional 4*10-6 s prior to ablation being switched on.

Figure 4.12 shows the grid for this test case. The grid has 240 nodes in the x1 (i)

direction and 294 nodes in the x2 (j) direction. The surface is located at j = 208, and has

an initial local coordinate of (2)i,j+w = 0.0. The coarse grid has x1 = 3.466*10-4 m and

x2 = 3.4*10-4m. The boundary layer is resolved with 20 points in the x2 direction with a

x2 = 2.5*10-6m, and this spacing is used for 56 point in the x2 direction below the

surface. There is an exponential expansion factor of 1.1 used to expand between the

refined and the coarse grid spacing over 12 points.

Figure 4.12: Grid for test cases 1 and 2.

146
Figure 4.13 shows the computed location and shape of the shock versus the

location provided by the correlations given by Anderson75. In this correlation the shock

shape is assumed to be hyperbolic and its location is given by:

 y 2 tan 2  
 
x  r    rc cot 2   1 
rc2
 1 , (4.3.1a)
 

where r is the radius of the spherical tip, is the shock detachment distance at the

stagnation point given by:

 3.24 
  0.143r * exp  ,
2 
(4.3.1b)
 M 

rc is the radius of curvature for the shock wave along the stagnation line given by:

 1.8 
rc  1.143r * exp  ,
0.75 
(4.3.1c)

 M  1 

and  is the oblique shock angle for the given cone angle. The numerical solution

provides a shock location that corresponds with the experimental correlation near the

spherical tip, and then provides a shock location closer to the surface along the cone

portion of the body. This is as expected because the angle  is provided by the oblique

shock relations, not the angle of the shock that would be expected from a cone.

147
Figure 4.13: Predicted vs. actual shock shape for M∞ = 9.5.

Figure 4.14: Comparison of temperature between chemically reacting and non-reacting


flow at M∞= 9.5.

148
The addition of gas phase chemical reactions significantly alters the computed

temperature of the flow behind the shock, as can be seen in Figure 4.14. This is

especially true in the stagnation region as can be seen in the inset. The maximum

temperature for the non-reacting case is 4548K, whereas for the reacting case it is 4109K.

Figure 4.15 and 4.16 show this comparison for the density and the pressure. The

maximum density for the non-reacting case is 12.48 kg/m3, and for the reacting case it is

13.18 kg/m3. The maximum pressure for the non-reacting case is 11.85*106 Pa and for

the reacting case it is 12.62*106 Pa.

149
Figure 4.15: Comparison of pressure between chemically reacting and non-reacting flow
at M∞= 9.5.

Figure 4.16: Comparison of density between chemically reacting and non-reacting flow
at M∞= 9.5.

150
7
10
12

10

106

Pressure (Pa)
8
(kg/m )
3

5
4 10

4
0 -5 -4 -3 -2 10 -5 -4 -3 -2
10 10 10 10 10 10 10 10
x2(m) x2(m)

3500
4000

3000

3000 2500
Temperature (K)

2000
u (m/s)

2000
1500

1000
1000

500

0 -5 -4 -3 -2 0 -5 -4 -3 -2
10 10 10 10 10 10 10 10
x2(m) x2(m)

Figure 4.17: Stagnation line profiles for test case 1 with chemically reacting flow without
ablation.

The profiles along the stagnation line for the chemically reacting flow without

ablation are shown above, with the surface located at x2=0. Thee profiles show that the

second order Riemann solvers capture the shock within a few grid points. The pressure

remains nearly constant after the shock, so as the gas is cooled by heat transfer into the

solid the density increases. This is as expected from the ideal gas equation of state.

151
100

-1
10

-2
10

mass concentration
10-3 O2
N2
10-4 O
N
-5 NO
10 +
NO
-6
10

10-7

10-8

-9
10 -5 -4 -3 -2
10 10 10 10
x2(m)

Figure 4.18: Profiles of mass concentration for test case 1 along the stagnation line with
chemically reacting flow and without ablation.

In the current test case the most prevalent gas-phase chemical reaction is the

dissociation of O2. At the maximum temperature behind the shock wave this is expected,

as Anderson75 states that the dissociation of N2 begins around 4000K, and if the gas were

at equilibrium conditions the O2 should be almost completely dissociated. Figure 4.18

shows the mass concentration along the stagnation line without ablation. The peak

concentration of O is 3.8%, so the O2 is not completely dissociated and the decision to

include non-equilibrium chemical reactions is justified. The next largest component of

gas is NO, which is being produced mainly by reaction 28 in Table B.4 which has atomic

oxygen and molecular nitrogen as its reactants. Figures 4.19-20 show the contours of

concentration vs. temperature for O2 and O. These are presented in computational space,

152
for ease of display, with an inset that shows the near surface region. In the cooler

temperatures of the boundary layer some of the O has recombined into O2.

Figure 4.19: Contours of temperature and O2 concentration for test case 1 with
chemically reacting flow without ablation.

153
Figure 4.20: Contours of temperature and O concentration for test case 1 with chemically
reacting flow without ablation.

Figure 4.21: Tangential velocity profiles along lines normal to the surface for test case 1
with chemically reacting flow without ablation.

154
Figure 4.22: Temperature profiles along lines normal to the surface for test case 1 with
chemically reacting flow without ablation.

10-2

x1=
-3
1.73E-4 10
1.91E-3
2.81E-3 x2 (m)
5.04E-3
9.20E-3
1.27E-2
1.61E-2 -4
10
1.96E-2
2.31E-2
2.65E-2
3.00E-2
3.35E-2

10-5

-6 -5 -4 -3 -2
10 10 10 10 10
mass concentration of O

Figure 4.23: O concentration profiles along lines normal to the surface for test case 1
with chemically reacting flow without ablation.

155
Profiles of the tangential velocity, u1, for chemically reacting flow without

ablation are shown in Figure 4.21. Each profile is along a line normal to the surface. At

the shock there is a sharp decrease in the velocity, then it steadily decreases towards the

surface, and then in the boundary layer there is a more rapid decrease to zero at the

surface. The inset shows the profiles very near the surface within the boundary layer.

From this figure it can be seen that the entire volume between the shock wave and surface

(the shock layer) contains viscous flow.

Profiles of the temperature are shown in Figure 4.22. The temperature goes from

its free stream value to a much higher temperature at the shock. It then increases due to

aerodynamic heating to some mild maximum and then begins to decrease due to heat

transfer into the surface. Again an inset shows the profiles near the surface.

Figure 4.23 shows profiles of atomic oxygen (O) concentration along lines normal

to the surface for various positions along the surface. As the temperature undergoes a

jump at the shock the molecular oxygen (O2) beings to dissociate, and the concentration

of O reaches a maximum at approximately the same point as the temperature. As the

temperature cools the O begins to recombine.

156
3000
Solid
Gas

2500

Temperature (K)
2000

1500

1000

0.01 0.02 0.03


x1(m)

(a) (b)
Figure 4.24: Solid Temperature. (a) Contours. (b) Surface Profile.

The solid temperature just prior to the implementation of ablation is shown in

Figure 4.24. The contours in computational space are shown in Figure 4.24a. The

temperature of the solid surface at the stagnation point has reached 3155 K. The profile

of the surface temperatures along the surface is shown in Figure 4.24b.

157
1.5x10-03 0.30

0.25

Sublimation Oxidation
mass flux (kg/(m s)

mass flux (kg/(m s)


-03
1.0x10 0.20
2

2
0.15

5.0x10-04
0.10

0.05

+00
0.0x10
0.01 0.02 0.03 0.01 0.02 0.03
x1(m) x1(m)

(a) (b)

2.0x10-06
-1
10

-2
10
1.5x10-06
Nitridation
mass flux (kg/(m2s)

mass flux (kg/(m s)


10-3
2

-06
10-4
1.0x10
Overall
-5
10 Oxidation
Nitridation
Sublimation
5.0x10
-07 10-6

10-7

0.0x10+00 10
-8
0.01 0.02 0.03 0.01 0.02 0.03
x1(m) x1(m)

(c) (d)
Figure 4.25: Ablative mass fluxes for test case 1. (a) Sublimation (b) Oxidation (c)
Nitridation (d) Comparison to overall.

The ablative mass fluxes are shown in Figure 4.33 at 7.66*10-6 s after the ablation

portion of the solver was implemented. At this point the mass flux results have reached a

quasi-stable solution for the current conditions. The peak flux is at the stagnation point

and has a value of 0.276 kg/(m2s) or 0.0276 g/(cm2s), of which 0.274 kg/(m2s) is from

oxidation, 1.66*10-6 kg/(m2s) is from nitridation and 1.38*10-3 kg/(m2s) is from

158
sublimation. The area averaged flux is 0.113 kg/(m2s). These results are compared to the

experimental results of Lundell and Dickey30 in Figure 4.46.

Figure 4.26 shows the position of the surface at each point in the local non-

dimensional coordinate 2. The initial surface was set to (2)i,j+w = 0.0 for each point. At

the stagnation point, the wall has receded to (2)i,j+w = 7.5*10-4. This translates to a

surface recession rate of .016 cm/s, which falls within the range predicted by 30. This rate

(~1 cm/min) means that significant shape change would occur over the course of a flight

at these conditions.

Figure 4.27 shows contours of the mass concentration of the ablation products

CO, CN, and C3. The contours show that the products have started to diffuse and convect

away from the surface.

Figure 4.28 shows the profile of the gas species along the stagnation line. Near

the surface the N and O profiles have steep drops, which correspond to the production of

CN and CO at the surface. This can be compared to the non-ablative case shown in

Figure 4.18. The concentration of NO rises near the surface because there is less O to

react with in reaction 29 in Table B.4.

159
0.0x10+00

-2.0x10-04

-04
-4.0x10

(2)i,j+w
-6.0x10-04

-04
-8.0x10

-03
-1.0x10
0.01 0.02 0.03
x1(m)

Figure 4.26: Surface position in local non-dimensional coordinates for test case 1.
5.0x10-05
5.0x10-05

c(CO) c(CN)
4.0x10-05 4.0x10-05 1.2x10-07
0.03 -07
0.028 1.1x10
-07
0.026 1.0x10
0.024 9.0x10-08
0.022 -05 8.0x10-08
3.0x10
-05 3.0x10 7.0x10-08
0.02 -08
6.0x10
x2(m)

0.018
x2(m)

-08
0.016 5.0x10
-08
0.014 4.0x10
-08
0.012 3.0x10
2.0x10-05 0.01 2.0x10-05 2.0x10
-08

-08
0.008 1.0x10
0.006
0.004
0.002 -05
-05
1.0x10 1.0x10

+00 +00
0.0x10 0.0x10
0 0.01 0.02 0.03 0 0.01 0.02 0.03
x1(m) x1(m)

(a) (b)
5.0x10-05

c(C3)
4.0x10-05 6.5x10-05
-05
6.0x10
-05
5.5x10
5.0x10-05
-05 4.5x10-05
3.0x10 4.0x10-05
-05
3.5x10
x2(m)

-05
3.0x10
-05
2.5x10
-05
2.0x10
2.0x10-05 1.5x10
-05

-05
1.0x10
5.0x10-06

-05
1.0x10

+00
0.0x10
0 0.01 0.02 0.03
x1(m)

(c)
Figure 4.27: Contours of ablation products for test case 1. (a) CO (b) CN (c) C3.

160
100

10-1

-2
10

mass concentration
-3
10
O2
N2
10-4 O
N
10-5 NO
+
NO
-6 CO
10
CN
C3
-7
10

10-8

10-9 -5 -4 -3 -2
10 10 10 10
x2(m)

Figure 4.28: Mass concentration profiles along the stagnation line for test case 1 with
ablation.

4.3.2 Test Case 2: Sea level flight with defect at TPS surface

One of the motivations behind this study was to be able to ascertain the effects of

a small surface defect in the overall performance of an ablative TPS. Preliminary results

in our study97 had indicated that the ablative mass fluxes would have an order of

magnitude increase near the upstream and downstream edges of the defect. However

these results were obtained on a coarse stair-case grid with the same issues noted in the

preliminary results section. In the current version of the model the grid for test cases 1

and 2 was designed to allow for adequate resolution of a defect. It should be noted that

the developed model is not limited to small defects.

The defect was taken to have a length of 1.25% of the overall surface length in the

x1 direction and an aspect ratio (length/depth) of ~2. Table 4.3 contains details of

location of the defect and Figure 4.29 shows a comparison between the original surface
161
and the surface with the defect. The defect extends from i = 41 to 115 and j = 123 to 208

(74 x 85 nodes) with varying grid spacing ensuring adequate boundary layer resolution

near the surfaces. Excluding the corner grid points the upstream side of the defect is

given an initial location in the local coordinate system of (1)i+w,j = 0.0, the downstream

side is given (1)i-w,j = 0.0, and the bottom is given (2)i,j+w = 0.0. This corresponds to

these finite volumes having an initial volume fraction  = 0.5. The upstream top corner

has an initial top surface located at (2)i,j+w = 0.0 and a right side given by (1)i+w,j = 0.0,

which corresponds to  = 0.25. The downstream top corner is given the same surface

coordinates and volume fraction with the exception of the surface being on the opposite

side. Finally the bottom corners are initialized as all solid as current version of the

numerical methods can only deal with convex surfaces within a given finite-volume.

Table 4.3: Geometry of the defect.


x1(m) x2(m) x(m) y(m) i j
Upstream top corner 2.7330E-03 0 1.9120E-04 2.7240E-03 41 208
Downstream top corner 3.1763E-03 0 2.5812E-04 3.1623E-03 115 208
Bottom -2.1250E-04 123
Length 4.4333E-04
Depth 2.1250E-04

.
Figure 4.29: Original and defect surfaces.
162
The numerical procedure to reach the point of simulating ablation was similar to

that in case 1 with each portion of the simulation being run until the solution has reached

a quasi-steady state. In this case, the simulation starts with a pre-converged solution from

test case 1 and then the defect is added. This is necessary to prevent the development of

complex shock patterns and interactions within the defect that would occur if the

simulation was initialized with every gas point at the freestream conditions. The

ablation-free simulation is run until the vortex within the defect has had sufficient time to

form. Finally the ablation portion of the solver is switched on.

Figure 4.30 shows the contours of temperature for the region of the defect, with

(a) showing the overall defect, and (b) and (c) showing the region near the top corners.

These results are shown in computational space. The overall defect shows a structure of

high temperature at the top that results from the boundary layer being disturbed by the

defect and hotter air from the shock layer being pulled into the defect where it is

impinges on the downstream side of the crack a small distance below the original surface.

This gas is also being heated by shear dissipation. There is another region along the

downstream side that has a higher temperature. Examination of the velocity vectors (see

Figure 4.31) in this region shows that this is where the main vortex detaches from the

wall and interacts with a weaker vortex in the lower downstream corner. The bottom of

the defect and the majority of the upstream side of the defect have been cooled by

conduction into the surrounding solid. The upstream top corner has a region of hotter gas

surrounding it. Based on these temperatures it would be reasonable to expect there to be

three peaks in the ablative mass fluxes that correspond to these regions of elevated gas

163
temperature (greater than 3500K) once the ablation portion of the solver is enabled.

These temperatures are also high enough for sublimation to dominate.

(a)

(b) (c)
Figure 4.30: Temperature contours in the defect pre-ablation. (a) Overall (b) Upstream
Top Corner (c) Downstream Top Corner

164
(a)

(b) (c)
Figure 4.31: Velocity vectors in defect region. (a) Overall (b) Upstream top corner (c)
Center region near original surface.

165
Figure 4.31 shows the velocity vectors in the region of the defect. Figure 4.31(a)

shows vectors at every point in the i direction and every third point in j direction. The

center of the main vortex is located at x1 = 3.14*10-3m and x2 = -1*10-4 m. There is a

secondary vortex in the bottom downstream corner with a center located at x1 = 3.15*10-
3
m and x2 = -1.7*10-4 m. Figure 4.31(b) provides the velocity vectors near the upstream

top corner and illustrates how the flow reemerging from the cavity disrupts the boundary

layer. Figure 4.31(c) shows the flow in middle of the x1 range of the defect near the

original surface.

Figure 4.32 provides the mass concentration of atomic oxygen in the defect

region. From this data, it can be expected that significant ablative mass fluxes due to

oxidation will occur near the top corners of the defect, along the downstream side, on the

bottom near the downstream side.

c(O)
0.0002 0.034
0.032
0.03
0.028
0.026
0.0001 0.024
0.022
0.02
0.018
0.016
0 0.014
x2(m)

0.012
0.01
0.008
0.006
0.004
-0.0001 0.002

-0.0002

0.0027 0.0028 0.0029 0.003 0.0031 0.0032


x1(m)

Figure 4.32: Concentration of atomic oxygen in the defect region.

166
Away from the defect the surface temperature of the solid at a given point for test

case 2 is lower than the temperature of the same point in test case 1 by about 100K as can

be seen in Figure 4.33. This is due to the shorter amount of physical time that was

simulated at each stage of the solution; additional amount of time was required to run the

computations for test case two because the defect increases the number of points at which

a gas solution is needed which takes a significantly longer time than a solid solution. For

example the ablation portion of the solver was run to simulate 1.5*10-7 s of physical time

for the surface with the defect compared to 7.66*10-6 s for the original surface. Multi-

time step simulations based on the use of different time steps for refined and coarse girds

will be implemented in future research to facilitate a decrease in the amount of computer

time required.

While an exact quantitative comparison between the two test cases is not available

a detailed qualitative comparison can be made.

3500

Defect
3000 Original
Surface Temperature (K)

2500

2000

1500

1000

500

0.01 0.02 0.03


x1(m)

Figure 4.33: Comparison between surface temperatures for test case 1 and 2 prior to
ablation.

167
0
10

10-1

ablative mass flux (kg/(m s))


-2
10

2
-3
10

-4
10

10-5

-6
10 Overall (defect)
Overall (original)
10
-7 Oxidation (defect)
Oxidation (original)
Sublimation (defect)
10-8
Sublimation (original)

0.005 0.01 0.015 0.02


x1(m)

(a)

0
10

10-1
ablative mass flux (kg/(m s))

-2
10
2

-3
10

-4
10

10-5

-6
10
Overall (defect)
-7 Overall (original)
10 Oxidation (defect)
Oxidation (original)
10-8 Sublimation (defect)
Sublimation (original)

0.0026 0.0028 0.003 0.0032 0.0034


x1(m)

(b)
Figure 4.34: Comparison between ablative mass fluxes for original and defect surfaces.
(a) Overall (b) Defect region

168
Figure 4.34 shows a comparison between the overall, oxidation and sublimation

ablative mass fluxes for the original surface in test case 1 and the same surface with a

defect in case 2. Near the stagnation point the oxidation ablative mass flux for the two

cases are of the same order: 0.275 kg/(m2s) for test case 1 vs. 0.207 kg/(m2s) for test case

2, with the difference due to the temperature difference. The sublimation ablative mass

flux has a much larger difference, with 5.4*10-3 kg/(m2s) for test case 2 and 1.4*10-3

kg/(m2s) for test case 1; the difference results from the concentration of C3 not having

enough time to develop in test case 2. Despite these differences and away from the

defect the ablative mass fluxes for the two cases exhibit the same behavior with the peak

at the stagnation point and a steady decline with temperature. Away from the defect the

oxidation flux is several orders of magnitude larger than the sublimation flux.

Figure 4.35 and Table 4.4 detail the behavior of the ablative mass flux in the

region of the defect. These illustrate the ablation environment occurring along the

surface of the defect. The gas entering the defect has an original concentration of atomic

oxygen that is not replenished. This affects the oxidation rate along the entire surface of

the defect.

Figure 4.35(a) shows the mass fluxes and temperature along the upstream side of

the defect, near the top corner. The sublimation flux is negligible compared to the

oxidation flux except very near the corner at which point sublimation reaches the same

order of magnitude as oxidation and even becomes larger, as the surface temperature

exceeds 3500K.

Along the downstream side of the defect the peak mass flux is occurring at the

location predicted when examining Figure 4.30. The concentration of O has dropped

169
from the .02 shown in Figure 4.32 to the 3.15*10-4 provided by Table 4.4. This is due to

both the increase in the concentration of C3 from sublimation and the reaction of the

atomic oxygen with the solid carbon.

Figure 4.35(c) shows the ablation profile along the entire bottom surface. Again

the sublimation rate is negligible compared to the oxidation flux. The peak oxidation and

sublimation rates occur at different locations, with the oxidation occurring farther away

from the corner. This can be accounted for the flowfield within the defect. The location

of the peak oxidation is where the main and secondary vortexes interact. As the

secondary vortex travels away from this point towards the downstream corner the atomic

oxygen content is decreasing via the oxidation reaction, and the surface temperature is

increasing which leads to an increase in the sublimation.

170
Table 4.4: Data for peak mass fluxes within the defect.
Top Peak on
Peak on
Location Upstream downstream
bottom
Corner side
x1 2.733E-03 3.176E-03 3.109E-03
x2 0.000E+00 -7.500E-06 -2.125E-04
x*1 0.000 1.000 0.847
x*2 0.000 0.035 1.000
T(K) 3675.4 3595.0 2760.8
c(O) 2.15E-04 3.15E-04 2.24E-05
ablative mass overall 2.185 2.171 9.715E-02
fluxes(kg/(m2s)) oxidation 0.951 1.574 9.686E-02
sublimation 1.234 0.597 2.940E-04

171
Tsolid(K) Tsolid(K)
1500 2000 2500 3000 3500 1500 2000 2500 3000 3500
0.0x10+00 0.0x10+00

Overall
-2.0x10-06 Oxidation -1.0x10-05 Overall
Sublimation Oxidation
Temperature Sublimation
Temperature
-06 -05
-4.0x10 -2.0x10
x2(m)

x2(m)
-6.0x10-06 -3.0x10-05

-06 -05
-8.0x10 -4.0x10

-05 -05
-1.0x10 -5 -4 -3 -2 -1 0 -5.0x10 -2 -1 0
10 10 10 10 10 10 10 10 10
2 2
ablative mass flux (kg/(m s)) ablative mass flux (kg/(m s))

(a) (b)

Overall
Oxidation
ablative mass flux (kg/(m2s))

Sublimation
Temperature 2500
10-2

Tsolid(K)
2000
10-3

1500
10-4
0.0028 0.0029 0.003 0.0031
x1(m)

(c)
Figure 4.35: Peak mass fluxes within the defect. (a) Top upstream corner
(b) Downstream side (c) Bottom.

This test case has shown that a small defect in a surface can drastically alter the

ablation environment in its vicinity. Based on the data of test case 1 it would seem that in

order to increase the computational efficiency of the method it would be appropriate to

ignore sublimation. However the current simulation shows that the increase in the

temperature near a defect can cause a sublimation rate that is of the same order as the

oxidation rate, and quickly begins to dominate as the temperature increases.

172
The higher local rates will begin to alter both the size and shape of the defect,

which has a feed back loop to the local rates. The evolution of the defect and its ablation

rates over a long time can be explored in future studies when the computational

efficiency of the solver will be enhanced by using implicit schemes and multiple time

steps where appropriate.

4.3.3 Test case 3: Arc-jet conditions in oxidation and nitridation dominant

regime

The third test case consists of the same basic geometry exposed to conditions

similar to the typical conditions found in an arc-jet test. The inlet conditions are the same

as those provided by Chen and Milos79 with the exception of the composition; at the inlet

temperature of 1428K they report a composition that has a significant portion of the O2

dissociated. In this test case the surface temperature at the stagnation point has reached

2473K.

Figure 4.36 shows the grid that was used for test cases 3 and 4. It has 121 points

in the x1 direction and 308 in the x2 direction, of which 4 points in each direction are

ghost points used for the boundary conditions. The surface is again located at j = 208,

and the boundary layer is resolved using the same refined grid spacing and transition as

test cases 1 and 2.

173
Figure 4.36: Grid for test cases 3 and 4.

Figure 4.37: Temperature contour comparison between test case 1 and 3.

174
Figure 4.37 shows a comparison between temperature contours for test cases 1

and 3. The peak temperature behind the shock wave for test case 3 is ~9800K compared

to ~4100K for test case 1. The shock displacement distance d for test case 3 can be seen

to be slightly larger than that of test case 1 which is as expected given the lower Mach

number (see Eq. 4.3.1b). The maximum pressure behind the shock wave is 125kPa

which is two orders of magnitude lower than test case 1. The density behind the shock

wave is 0.147 kg/m3, which is also much lower than the density obtained in test case 1.

100 10000

-1
10

Temperature(K)
-2
10 8000
mass concentration

-3
10

10-4 6000

-5
10

-6 O2
10 4000 N2
O
10
-7 N
NO
+
NO
10-8 2000 e
-

CO
10-9 CN
-5 -4 -3 -2
10 10 10 10 C3
x2(m) T

Figure 4.38: Profiles of species mass concentration and temperature along the stagnation
line for test case 3.

At the shock the sudden increase in temperature causes the molecular oxygen and

nitrogen to begin to dissociate. Figure 4.38 shows the profiles of the species mass

concentrations and temperature along the stagnation line. At the shock the molecular

oxygen and nitrogen undergo chemical reactions and their concentrations undergo
175
significant decreases with the O2 dropping from 0.22 to 2.5*10-5. In this case enough

nitrogen is present (a maximum of ~10%) to make nitridation important. The ablation

products CO, CN and C3 have diffused throughout the boundary layer and then decrease

in concentration out to the shock.

1e-6 s
2e-6 s
3e-6 s
0.3
4e-6 s
5e-6 s
6e-6 s
0.25 7e-6 s
ablative mass flux (kg/(m 2))

8e-6 s
9e-6 s
2

1e-5 s
0.2

0.15

0.1

0.05

0
0.005 0.01 0.015
x1(m)

Figure 4.39: Convergence of ablative mass flux for test case 3.

Figure 4.39 shows the profiles of the ablative mass flux for various times. After

10-5s have elapsed with the ablation portion of the solver enabled the flux appears to have

converged to a curve that is not varying with time.

Figure 4.40 shows the ablative mass flux profiles for this converged curve. The

overall stagnation point mass flux is 0.198 kg/(m2s), of which 0.132 kg/(m2s) is due to

oxidation, 0.066 kg/(m2s) is due to nitridation and a miniscule 4.76*10-6 kg/(m2s) is due

to sublimation; The area-averaged rate is 0.142 kg/(m2s). The low sublimation rate is due

176
to the fact that the surface temperature is still relatively low despite the high gas

temperature.

Figure 4.40: Ablative mass fluxes for test case 3.

4.3.4 Test case 4: arc-jet conditions in sublimation dominant regime

The final test case is at a stagnation point surface temperature (3850K) where

sublimation should be the dominant contributor to the ablative mass flux. The freestream

conditions are the same as those under test 3. Under the given conditions the equilibrium

partial pressure of the C3 is higher (293.4 kPa) than the actual mixture pressure (125

kPa), which provides a very strong test for the sublimation model. The simulation is run

with ablation enable for 10-7 s after when the mass flux rates have begun to converge.

177
100 10000

-1
10

Temperature(K)
-2
10 8000
mass concentration
-3
10

10-4 6000

-5
10

-6 O2
10 4000 N2
O
10
-7 N
NO
+
NO
10-8 2000 e
-

CO
10-9 CN
-5 -4 -3 -2
10 10 10 10 C3
x2(m) T

Figure 4.41: Profiles of species mass concentration and temperature along the stagnation
line for test case 4 using Hertz-Knudsen.

Figure 4.41 shows the profiles of the gas temperature and mass concentrations

along the stagnation line. As can be seen the sublimation has released enough C3 to make

the concentration at the surface equal to 0.98. There are only trace amounts of O, and N

which mean that the oxidation and nitridation reactions do not have enough reactants to

produce significant amounts of CO and CN.

Figure 4.42 shows the ablative mass fluxes for test case 4 and the surface

temperature of the solid using the Hertz-Knudsen equation for sublimation (Eq. 2.4.49).

The stagnation point mass flux is 22.77 kg/(m2s), of which 22.75 kg/(m2s) is from

sublimation, .015 kg/(m2s) is from oxidation and the remainder from nitridation.

Downstream the oxidation and nitridation fluxes increase until the oxidation and

sublimation fluxes are equal after which the oxidation begins to dominate and the

178
sublimation quickly drops to a negligible amount. The ablative flux averaged over the

surface area is 1.95 kg/(m2s). In this severe case the stagnation flux is 11.7 times greater

than the average surface flux. Further downstream the local ablative flux is almost 40

times less than the average. Clearly using the average rate to design a graphite TPS

under these conditions would be a mistake.

Figure 4.43 shows the ablative mass fluxes using the Ytrehus sublimation mass

flux (Eq. 2.4.57). The overall ablative mass flux has increased to 36.40 kg/(m2s), which

is almost entirely from sublimation. The average mass flux has also increased to 2.82

kg/(m2s).

Figure 4.44(a) compares the sublimation flux provided by the Hertz-Knudsen and

Ytrehus equations. The Ytrehus provides a higher rate across the entire temperature

range, especially in the hotter region. An examination of the two equations shows that in

the Ytrehus equation the term that deals with the actual partial pressure of C3 present at

the surface is multiplied by a coefficient that is less than unity under the current

conditions. Figure 4.46 in the next section compares the results of this test case with

experimental data, which shows that the Hertz-Knudsen rate is closer to the experimental

data. Nevertheless the Ytrehus model has been shown (see Figure 2.2) to eliminate

oscillations in the mass flux rate as it approaches zero, which would occur in strong

sublimation situations where the mixture pressure is larger than the equilibrium pressure

of C3.

Figure 4.44(b) compares the injection velocities at the surface versus and the

speed of sound. The injection velocities remain subsonic, but as the mass flux increases

at higher temperature they could approach the speed of sound (which may occur if there

179
were a surface defect similar to the defect found in test case 2). The Ytrehus model also

implicitly contains the requirement that the injection velocity not become supersonic

while this restriction must be added in as a separate condition on the Hertz-Knudsen rate.

Tsolid
Overall
Oxidation
102 Nitridation 4000
Sublimation
Area Average
101
ablative mass flux (kg/(m s))

3500
2

100

Temperature (K)
10-1
3000
10-2

10-3
2500

10-4

-5
10 2000
0.01 0.02 0.03 0.04
x1(m)
Figure 4.42: Ablative mass fluxes for test case 4 with Hertz-Knudsen sublimation.

180
Tsolid
Overall
Oxidation
102 Nitridation 4000
Sublimation
Area Average
101

ablative mass flux (kg/(m s))


2 3500
100

Temperature (K)
10-1
3000
-2
10

10-3
2500

10-4

-5
10 2000
0.01 0.02 0.03 0.04
x1(m)
Figure 4.43: Ablative mass fluxes for test case 4 with Ytrheus sublimation.

1500
Sublimation mass flux (kg/(m s))

30 Hertz-Knudsen
2

Ytrehus
Velocity (m/s)

1000
20

Hertz-Knudsen
Ytrehus
10 500
Speed of Sound

0 0
0.01 0.02 0.03 0.04 0 0.01 0.02 0.03 0.04
x1(m) x1(m)

(a) (b)
Figure 4.44: Comparison between Hertz-Knudsen and Ytrehus models (a) mass fluxes
(b) injection velocities.

Figure 4.45(a) shows a comparison of the shear stresses at the wall for the

conditions of test case 4 with and without ablation. The presence of ablation reduces the

shear stress especially near the stagnation point. The drag due to skin friction can be

181
found by integrating the shear stress over the surface area. For the non ablative case this

provides a drag force of 5.27N, and the ablative case provides a drag force of 4.28N

which is a reduction of 23%. This observation agrees with the experimental results of

Kruse35 which showed a reduction in drag from ablation.

The reason for this decrease in drag can be shown by examining Figure 4.45(b)

which contains profiles of the tangential velocity (u2) near the wall at various x1

coordinates along the wall. The profiles nearest the stagnation point (x1 = 0) have slopes

that have been significantly flattened close to the wall due to the large injection velocity.

Away from the region of high ablative mass fluxes the velocity profiles return to their

non-ablative behavior.

182
2000
-04
1.0x10
w/o ablation
ablation

1500 8.0x10-05
wall(N/m2)

6.0x10-05

x2(m)
1000 x1 =
0.176E-03 (abl)
0.176E-03 (w/o abl)
4.0x10
-05 0.721E-02 (abl)
0.721E-02 (w/o abl)
0.107E-01 (abl)
500 0.107E-01 (w/o abl)
0.213E-01 (abl)
2.0x10-05 0.213E-01 (w/o abl)
0.353E-01 (abl)
0.353E-01 (w/o abl)
0
0.01 0.02 0.03 0.04 0 500 1000 1500
x1(m) u1(m)

(a) (b)

-100
-05
1.0x10

-150 8.0x10-06

w/o ablation
q"wall(MW/m2)

ablation -06
6.0x10
x2(m)

-200

4.0x10-06

-250
2.0x10-06

+00
-300 0.0x10
0.01 0.02 0.03 0.04 2000 2500 3000 3500 4000 4500
x1(m) Temperature (K)

(c) (d)
Figure 4.45: Comparison of wall conditions for ablative and non-ablative test case 4.
(a) Shear stress (b) Velocity profiles (c) Heat flux (d) Temperature Profiles.

Figure 4.45(c) shows a comparison of the wall heat fluxes between the ablative

and non-ablative cases. At the stagnation point the ablative case shows the gas providing

a heat flux of 272 MW/m2 compared to 289 MW/m2 for the non ablative case. It should

be noted that in addition to this reduction, two of the ablation process also absorb large

amounts of heat (22.7 MJ/kg for sublimation, and 16.7 MJ/kg for nitridation) which

reduces the heat flux going into the solid even more. The oxidation reaction releases 3.9
183
MJ/kg but this is small compared to the reduction in heat flux being provided. The exact

contribution of these reactions to the overall heat flux entering the surface can be found

by multiplying the MJ/kg values by the ablative mass fluxes for each process. It is

interesting to note that downstream the heat flux for the ablative case becomes larger than

the non-ablative case, although it remains below the peak stagnation point flux.

Figure 4.45(d) displays profiles of the temperature near the wall. The wall

temperature has increased as time has elapsed during the ablation simulation, so the

temperatures at the point nearest the wall do not coincide between the ablative and non-

ablative cases. The green profiles are within the region where the flux is higher, and as

can be seen the temperature gradient is much higher in the ablative case. The exact reason

for this higher gradient in this region is unclear, although it is in the same general

location as where the ablative shear profile in Figure 4.45(a) begins to become parallel to

the non-ablative profile. It should be noted that the scales of (b) and (d) are different

because the temperature boundary layer is thinner than the velocity boundary layer.

4.3.5 Comparison to experimental data

The four test cases considered in this section have shown that the developed

model is capable of handling a wide range of input conditions, and the ability to resolve

localized peaks in the ablation environment. Even though the flight or experimental data

may correspond to the ablation being primarily oxidation localized regions may transition

to sublimation dominated ablation. Table 4.5 summarizes the stagnation point ablative

mass fluxes and the area averaged mass fluxes for test case 1, 3 and 4.

184
Table 4.5: Stagnation point ablative mass fluxes
2 Area
Stagnation point mass Flux (kg/(m s))
Averaged
T(k) Overall Oxidation Nitridation Sublimation Overall
Case 1 3150.50 0.276 0.274 1.661E-06 1.378E-03 0.113
Case 3 2473.06 0.198 0.132 6.591E-02 4.758E-06 0.142
Case 4
3849.00
(H-K) 22.769 1.532E-02 4.211E-04 22.753 1.951
Case 4 (Yt) 3848.91 36.405 1.661E-03 3.208E-05 36.403 2.818
H-K: Hertz Knudsen. Yt: Ytrehus.

0.4
Lundell and Dickey
exponential curve fit
avaerage ablative mass flux (g/(cm s))

test case results


2

0.3
Ytrehus
4

0.2 Hertz-Knudsen

0.1
3 1

0
2500 3000 3500 4000
Tsolid(K)

Figure 4.46: Comparison between computational results and the experiments of Lundell
and Dickey.

185
Figure 4.46 shows a comparison between the experimental results of Lundell and

Dickey30 and the results obtained in this dissertation. The exact method at which they

arrived at these values and an error analysis are lacking, but enough detail exists to

ascertain that they determined the amount of mass loss, and an average surface area and

then used the overall time of the run to generate an ablative mass flux. The curve fit is:

~ T 6.454*103 *Tsolid   0.013 ,


exp solid   1.122 *10 * e
12
m (4.4.1)

where the temperature is in K and the ablative mass flux is in g/(cm2s). The conversion

factor is 1 g/(cm2s) = 0.1 kg/(m2s). The area averaged values for the mass flux have

been used to compare to the experimental results. At the lower temperature range where

oxidation and nitridation dominate the numerical results have very good agreement with

the experimental results. At the higher temperatures of test case 4, sublimation

dominates and the agreement with the results are not as good, but are still comparable

with the Hertz-Knudsen model provides a result that is closer to the experiment. The

discrepancy between the numerical and experimental results in the sublimation range is

probably due to the choice in equilibrium pressure function for C3, which is disagreed

upon within the relevant literature.

4.4 Summary

This chapter has provided results for the numerical methods against a variety of

test cases of increasing complexity. Where possible the obtained computational results

were compared to analytical or experimental results.

First the various sub-models such as inviscid and viscous gas dynamics were

tested against corresponding standard numerical tests. These tests ranged from one-
186
dimensional inviscid flow in the Sod problem using first and second order numerical

fluxes, to flat plate boundary layer and driven cavity flow for viscous flow, to a semi-

infinite solid solution to test the conduction heat transfer portion of the solver.

Second a variety of test cases were run across a wide range of hypersonic flow

conditions to validate the functionality of the full model. The shock shape has been

shown to follow the experimental correlations and the general behavior of the flow

properties and composition is as expected. The decision to use non-equilibrium chemical

reactions was shown to be correct by the composition of the gas behind the shock wave.

The model has been shown to handle small defects in the surface and the different

ablation reactions. The results have been compared to experimental results and shown to

agree with the experiment, especially at the relatively lower temperature range where

oxidation and nitridation dominate.

The results of the validation cases, the agreement of the ablation test cases with

experimental results, and the ability to handle surface defects verifies the functionality

and flexibility of the novel reactive-Riemann solver and the overall numerical model. The

overall models ability to use different sub-models such as the different Riemann solvers

or the choice between Ytrehus or Hertz-Knudsen sublimation models has been

demonstrated.

187
CHAPTER V

CONCLUSIONS

In Chapter 1 the existing state of the art in ablation research was detailed. A need

was identified for a computational method that couples all of the various ablation sub-

processes in as tight a manner as possible. The objectives for a research program aimed

at providing a solution for this need were given as:

1. Develop a comprehensive mathematical model of the non-charring ablation

process based on first principles that takes into account the various gas-dynamic,

heat transfer, and chemical reaction processes including the development of

appropriate surface mass and energy balance equations

2. Develop new or modify existing numerical methods to simulate the developed

model with varying levels of resolution that are capable of handling the moving

ablation surface intrinsically and account for surface defects

3. Validate the mathematical model and numerical methods against experimental

arc-jet results for hypersonic ablation.

All of these objectives have been met during the course of this research program.

Chapter 2 details the development of the comprehensive mathematical model.

Starting from first principles and using the Reynolds Transport Theorem, a novel coupled

system of governing conservation equations for the solid and gas is developed that

188
account for the surface chemical reactions (oxidation, nitridation, and sublimation) and

recession of the surface. The final version of these equations is in a form suitable for

problem geometries that can be described with two-dimensional general orthogonal

coordinates. The associated gas phase thermodynamic relations, transport models, and

hypersonic chemical reaction equations are also provided. Novel forms of the surface

mass and energy balances that account for the ablation of the surface are also presented.

The numerical methods necessary for the solution of the mathematical model are

provided in Chapter 3. Finite volume spatial discretization methods are modified to

account for both the overall finite volume associated with a grid point, and the various

configurations of gas and solid contained within a finite volume encompassing the

ablative surface. A first order explicit time discretization for the governing equations is

presented with terms that account for the changing volume of the solid. In order to

calculate the inviscid portion of the fluxes, existing Riemann solvers are modified to take

into account multiple chemical species originating from dissociation of the air and high

temperature effects such as varying thermophysical properties; these solvers are

presented in both first and second order of approximation and the necessity for second

order is illustrated. A new reactive-Riemann solver is proposed to determine the fluxes at

the ablative surface using the surface mass and energy balances. A time step limitation

resulting from the movement of the solid surface is proposed and is used in conjunction

with the existing time step limitations for other processes to determine the overall

maximum allowable time step. Finally an algorithm that combines these methods is

developed for both serial and parallel methods.

189
The validation tests for the different sub-processes are presented in Chapter 4

including the one-dimensional Sod problem and two-dimensional supersonic wedge flow

problem for inviscid gas dynamics, flat plate boundary layer and driven cavity flow for

the viscous gas dynamics, and conduction into a semi-infinite solid. The convergence of

the results to the known solutions with increasing grid resolution is presented.

A summary of results from a preliminary simplified model based on a staircase

approximation of the surface are then presented to illustrate the reasons for some of the

choices made in the development of the final version of this model. The use of variable

specific-heat and chemical reactions within the gas phase are justified by the non-

physical high temperatures of the shock layer. The need for a body-fitted coordinate

system was also shown. The chief driving force for these enhancements was to eliminate

the presence of non-physical ablation plumes.

Finally several TPS ablation test cases were provided that illustrate the

functionality of the model against a range of conditions. The first two cases are at sea

level flight conditions and represent the highest free-stream pressures and densities that a

real aerospace vehicle could be expected to encounter. The second set of cases is at arc-

jet conditions which are the standard method for testing the performance of TPS

materials.

The first test case is a defect-free surface at sea-level pressure, 300 K temperature

and Mach 9.5 freestream conditions. The choice to use non-equilibrium versus

equilibrium chemical reactions is justified. The shape and offset of the shock wave are

shown to agree with experimental correlations. The ablation results show that as

expected for the surface temperature range that the ablation is primarily via oxidation.

190
The second test case is at the same initial free-stream conditions as the first one

with a small cavity-type defect located on the surface. In the region of this small defect

there are locations that have significant increases in surface temperature and ablation

rates. At these locations the ablation process transitions from being oxidation to

sublimation dominated. The developed model has been shown capable of handling the

complex interaction between ablation, surface heat transfer and the behavior of vortexes

within the defect.

The third test case is a defect-free surface under arc-jet conditions that has a

surface temperature that is within the oxidation and nitridation range. A comparison is

made between the shock wave location and shape for the arc-jet and sea-level flight

conditions. The convergence of the ablative mass fluxes to a quasi-steady rate is shown.

The fourth test case is the same as the third with the exception of an elevated

surface temperature that places the stagnation region within the sublimation dominated

temperature range. This provides the most extreme test of the model because of the

strong blowing from the surface. A comparison is provided between the Hertz-Knudsen

and Ytrehus sublimation rates. It is shown that in this strong sublimation case that the

use of a surface average ablative mass flux for design purposes would severely

underestimate the stagnation region ablation and overestimate the ablation further

downstream. The wall shear stress and heat flux are shown for the ablative and non-

ablative cases are compared, which shows that ablation significantly reduces the surface

shear (and therefore drag) and the heat flux within the stagnation region.

The three test cases without surface defects are compared to the experiments of

Lundell and Dickey30, which are interpreted to be equivalent to a surface average. The

191
surface average ablative mass fluxes of the first and third test cases (which fall within the

oxidation and nitridation dominated temperature ranges) agree very well with the

experimental results. The results from the fourth test case are higher than the

experimental results, with the Hertz-Knudsen result being closer. The models for

oxidation and nitridation are consistent throughout the literature but a function (of the

form of Eq. 2.4.51) for the equilibrium pressure of C3, which is important in both the

Ytrehus and Hertz-Knudsen models, has not been agreed upon, with a wide range being

reported. The value used in the current study falls in the middle of the range, and was

recommended by Chen and Milos79.

To summarize, this dissertation has provided a proof-of-concept for a tightly

coupled ablation solver that incorporates the different sub processes that are important for

the non-charring ablation problem. A novel mathematical model for ablation has been

developed starting from first principles that provides the coupling intrinsically. A set of

numerical methods including a novel reactive-Riemann solver has been integrated into an

algorithm for the solution of hypersonic ablation problems. The model and algorithm and

their various components have been validated against test cases of increasing complexity

up to experimental results for hypersonic ablation of graphite. The importance of being

able to resolve local defects is shown: a defect can produce local regions where the

temperature increases to the point that the sublimation reaction begins to dominate

leading to ablative mass fluxes much higher than the same surface without a defect.

192
The research conducted in this dissertation program is by no means at a dead-end.

Many opportunities exist to continue this research including enhancements to the

physical model and numerical efficiency.

The current physical model has assumed that the gas phase is in thermal

equilibrium, but more accurate gas dynamics would include the possibility of a two-

temperature model allowing vibrational non-equilibrium. Radiative heating should be

included in the future especially at the higher temperature ranges where sublimation is

expected to dominate. The physical model could also be expanded to allow for charring

ablators provided a detailed model of the charring process including the composition and

properties of the pyrolysis gas. A variety of the C3 equilibrium pressure functions could

be tested against experimental results to determine which if any provides better

agreement.

The current numerical methods are computationally expensive, which is in part

due to the explicit time step limitations due to the resolution of the boundary layer. A

multiple time step scheme could be implemented in order to resolve this. Also better

parallel computing techniques need to be explored including different domain

decompositions such as two-dimensional decomposition.

193
REFERENCES

1
Gnoffo, P.A., “Planetary-Entry Gas Dynamics,” Annual Review of Fluid Mechanics,
Vol. 31, 1999, pp. 459-94.
2
Heppenheimer, T. A., Facing The Heat Barrier: A History of Hypersonics, NASA
History Series, NASA SP-2007-4232, Washington, DC, 2007.
3
Sutton, G. W., “The Initial Development of Ablation Heat Protection, An Historical
Perspective,” Journal of Space Craft and Rockets, Vol. 19, No. 1, 1982, pp. 3-11.
4
Erb, R. B., and Jacobs, S., “Entry Performance of the Mercury Spacecraft Heat Shield,”
NASA TMX-57097, 1964.
5
NASA, “Project Gemini Quarterly Status Report, Period Ending 30 Nov. 1962,” NASA
TM-X-51176, 1962.
6
Avco Corp., “Apollo Heat Shield Monthly Progress Report,” NASA CR-127995, 1964.
7
Houston, R. S., and Hallion, R. P., “Transiting from Air to Space: The North American
X-15,” The Hypersonic Revolution: Case Studies in the History of Hypersonic
Technology, Edited by R. P. Hallion, Vol. 1, Air Force History and Museums Program,
Washington, DC, 1998.
8
Price, A. B., “Design Report – Thermal Protection System X-15A-2,” NASA CR-
82003, 1968.
9
Geiger, C. J., “Strangled Infant: The Boeing X-20A Dyna-Soar,” The Hypersonic
Revolution: Case Studies in the History of Hypersonic Technology, Edited by R. P.
Hallion, Vol. 1, Air Force History and Museums Program, Washington, DC, 1998.
10
Hallion, R. P., “ASSET: Pioneer of Lifting Reentry,” The Hypersonic Revolution: Case
Studies in the History of Hypersonic Technology, Edited by R. P. Hallion, Vol. 1, Air
Force History and Museums Program, Washington, DC, 1998.
11
Vitelli, J. L., and Hallion, R. P., “Project PRIME: Hypersonic Reentry from Space,”
The Hypersonic Revolution: Case Studies in the History of Hypersonic Technology,
Edited by R. P. Hallion, Vol. 1, Air Force History and Museums Program, Washington,
DC, 1998.

194
12
Chandler, H. H., “Low-cost ablative heat shields for space shuttles,” NASA CR-
111800, 1970.
13
Hiltz, A. A., “Exploratory Development of a Flexible Ablative Covering for Space
Shuttle Application,” NASA NAS-9-12067, 1972.
14
Walberg, G.D., and Sullivan, E. M., “Ablative Heat Shields for Planetary Entries – A
Technology Review,” NASA TMX-66946, 1970.
15
Gnoffo, P. A., Weilmuenster, K. J., Hamilton II, H. H., Olynick, D. R., and
Venkatapathy, E., “Computational Aerothermodynamic Design Issues for Hypersonic
Vehicles,” Journal of Spacecraft and Rockets, Vol. 36, No. 1, 1999, pp. 21-43.
16
Pallix, J., “High-Temperature Behavior of Advanced Spacecraft TPS,” NASA CR-
195832, 1994.
17
Lycans, R. W., “X-33 Base Region Thermal Protection System Design Study,” 7th
AIAA/ASME Joint Thermophysics and Heat Transfer Conference, ASME, New York,
1998.
18
Davis, D., “Thermal Analysis of the MC-1 Chamber/Nozzle,” Space Technology and
Applications International Forum 2001, Albuquerque, NM, Feb. 11-14, 2001.
19
Smith, D., Carroll, C., Marschall, J., and Pallix, J., “Materials Testing on the DC-X and
DC-XA,” NASA TM-110430, 1997.
20
Steltzner, A., Desai, P., Lee, W., and Bruno, R., “The Mars Exploration Rovers Entry
Descent and Landing and the Use of Aerodynamic Decelerators,” AIAA 2003-2125.
AIAA ADS Conference, Monterey, CA, May 20-22, 2003.
21
Zhong, J., Ozawa, T., and Levin, D., “Modeling of Stardust Reentry Reacting Thermal
and Chemical Ablation Flow,” AIAA-2007-455, 39th AIAA Thermophysics Conference,
Miami, FL, June 25-28, 2007.
22
Olynick, D., Chen, Y.K., and Tauber, M.E., “Forebody TPS Sizing with Radiation and
Ablation for the Stardust Sample Return Capsule,” AIAA-1997-2474, 32nd AIAA
Thermophysics Conference, Atlanta, GA, June 23-25, 1997.
23
Ko, W.L., Gong, L., and Quinn, R. D., “Reentry Thermal Analysis of a Generic Crew
Exploration Vehicle Structure,” NASA TM-2007-214607, 2007.
24
Rahsis, B., and O’Hare, R. J., “Free-Flight Investigation of Reentry Heat Transfer and
Ablation at Velocities up to 22,500 Feet per Second,” NASA TMX-970, 1964.

195
25
Wick, B. H., “Ablation Characteristics and Their Evaluation by Means of Arc Jets and
Arc Radiation Sources,” 7th Congress International Aeronautique, Paper 10, Paris, June
14-16, 1965.
26
Scala, S. M., and Gilbert, L. M., “The Sublimation of Graphite at Hypersonic Speeds,”
AIAA-1964-1320, AIAA Entry Technology Conference, Williamsburg & Hampton, VA,
October 12-14, 1964, pp. 239-258.
27
Metzger J. W., Engel, M. J., and Diaconis N. S., “The Oxidation and Sublimation of
Graphite in Simulated Re-Entry Environment,” NASA CR-60395, 1965.
28
Maahs, H. G., “Ablation Performance of Glasslike Carbons, Pyrolytic Graphite and
Artificial Graphite in the Stagnation Pressure Range 0.035 to 15 Atmospheres,” NASA
TN-D-7005, 1970.
29
Maahs, H. G., “A Study of the Effect of Selected Material Properties on the Ablation
Performance of Artificial Graphite,” NASA TN D-6624, 1972.
30
Lundell, J. H. and Dickey, R. B., “Ablation of ATJ Graphite at High Temperatures,”
AIAA Journal, Vol. 11, No. 2, 1973, pp. 216-222.
31
Wool, M. R., “Passive Nosetip Technology (PANT) Program. Volume X. Summary of
Experimental and Analytical Results,” SAMSO-TR-74-86, 1975.
32
Chen, Y. K., Milos, F. S., Reda, S. C., and Stewart, D.A., “Graphite Ablation and
Thermal Response Simulation Under Arc-Jet Flow Conditions,” AIAA 2003-4002, 36th
AIAA Thermophysics Conference, Orlando, FL, June 23-26, 2003.
33
Marvin, J. G., and Pope, R.B., “Laminar Convective Heating and Ablation in the Mars
Atmosphere,” AIAA Journal, Vol. 5, No. 2, 1967, pp. 240-248.
34
Williams, M. E., and Chapman, G. T., “Free Flight Determination of Boundary Layer
Transition on Small Scale Cones in the presence of Surface Ablation,” NASA TM-X-
68867, 1972.
35
Kruse, R. L., “Comparison of the Aerodynamic Characteristics of an Ablating and
Nonablating Blunted Conical Body,” NASA TN D-7196, 1973.
36
Kruse, R. L., “The Effect of Asymmetric Ablation on Trim Angle of Attack,” NASA
TM 84309, 1982.
37
Mullenix, N., “A Coupled Gas Dynamics and Heat Transfer Method for Simulating the
Laser Ablation Process of Carbon Nanotube Production,” M.S. Thesis, Mechanical
Engineering Dept., University of Akron, Akron, OH, 2005.

196
38
Laub, B., “Ablator Modeling: Historical perspectives, surface ablation and in-depth
modeling,” AFOSR/NASA/SNL Workshop on Future Research Directions for Ablation
and Environment-Material Interactions, 2008.
39
Park, C., “Review of Chemical-Kinetic Problems of Future NASA Missions, I: Earth
Entries”. Journal of Thermophysics and Heat Transfer, Vol. 7, No. 3, 1993, pp. 385-398.
40
Park, C., Jaffr, R. L., and Partridge, H., “Chemical-Kinetic Parameters of Hyperbolic
Earth Entry,” Journal of Thermophysics and Heat Transfer, Vol. 15, No. 1, 2001, pp. 76-
90.
41
Zhluktov, S. V., and Abe, T., “Viscous Shock-Layer Simulation of Airflow past
Ablating Blunt Body with Carbon Surface,” Journal of Thermophysics and Heat
Transfer. Vol. 13, No. 1, 1999, pp. 50-59.
42
Havstad, M. A., and Ferencz, R. M., “Comparison of Surface Chemical Kinetic Models
for Ablative Reentry of Graphite,” Journal of Thermophysics and Heat Transfer, Vol. 16,
No. 4, 2002, pp. 508-515.
43
Milos, F. S., and Rasky, D. J., “Review of Numerical Procedures for Computational
Surface Thermochemistry,” Journal of Thermophysics and Heat Transfer, Vol. 8, No. 1,
1994, pp. 24-34.
44
Ytrehus, T., “Kinetic Theory Description and Experimental Results for Vapor Motion
in Arbitrary Strong Evaporation,” Technical Note 112, Von Karman Institute for Fluid
Dynamics, 1975.
45
Yrehus, T., and Ostmo, S., “Kinetic Theory approach to Interphase Processes,”
International Journal of Multiphase Flow, Vol. 22, 1996, pp. 133-155.
46
Pekker, L., Keidar, M., and Cambier, J-L., “Effect of thermal conductivity on the
Knudsen layer at ablative surfaces,” Journal of Applied Physics, Vol. 103, No. 3, 2008,
pp. 034906-1-6.
47
Hogan, R. E., Blackwell, B. F., and Cochran, R. J., “Numerical Solution of Two-
Dimensional Ablation Problems Usinag the Finite Control Volume Method with
Unstructured Grids,” AIAA-94-2985, 6th AIAA/ASME Joint Thermophysics and Heat
Transfer Conference, Colorado Springs, CO, June 20-23, 1994.
48
Storti M., “Numerical modeling of ablation phenomena as two-phase Stefan problem,”
International Journal of Heat and Mass Transfer, Vol. 38, No. 15, 1995, pp. 2843-2854.
49
Potts, R. L., “Application of Integral Methods to Ablation Charring Erosion, a
Review,” Journal of Spacecraft and Rockets, Vol. 32, No. 2, 1995, pp. 200-209.

197
50
Chen, Y. K., and Milos, F. S., “Ablation and Thermal Response Program for
Spacecraft Heatshield Analysis,” AIAA-98-0273, 36th AIAA Aerospace Sciences Meeting
and Exhibit, Reno, NV, Jan. 12-15, 1998.
51
Shih, Y. C., Cheung, F. B., Koo, J. H., and Yang, B. C., “Numerical Study of Transient
Thermal Ablation of High-Temperature Insulation Materials,” Journal of Thermophysics
and Heat Transfer, Vol. 17, No. 1, 2003, pp. 53-61.
52
Dec, J. A., and Braun, R. D., “An Approximate Thermal Protection System Sizing Tool
for Entry System Design,” AIAA 2006-780, 44th AIAA Aerospace Sciences Meeting,
Reno, NV, Jan. 9-12, 2006.
53
Amar, A. J., Blackwell, B. F., and Edwards, J. R., “One-Dimensional Ablation Using a
Full Newton’s Method and Finite Control Volume Procedure,” Journal of Thermophysics
and Heat Transfer, Vol. 22, No. 1, 2008, pp. 71-82.
54
Mitchell, S. L., and Meyers, T. G. “Heat Balance Integral Method for One-
Dimensional Finite Ablation,” Journal of Thermophysics and Heat Transfer, Vol. 22, No.
3, 2008, pp. 508-514.
55
Engel, C. D., Farmer, R. C., Pike, R. W., “Ablation and Radiation Coupled Viscous
Hypersonic Shock Layers,” AIAA Journal, Vol. 11, No. 8, 1973, pp. 1174-1181.
56
Konuraskai, K., and Candler, G., “Laminar-to-Turbulent Transitions over an Ablating
Reentry Capsule,” Acta Astronautica, Vol. 47, No. 10, 2000, pp. 745-751.
57
Zhong, J. Ozawa, T., and Levin, D., “Modeling of Stardust Reentry Reacting Thermal
and Chemical Ablation Flow,” AIAA-2007-455, 39th AIAA Thermophysics Conference,
Miami, FL, June 25-28, 2007.
58
Barbante, P. F., Dgrez, G., and Sarma, G. S. R., “Computation of Nonequilibrium
High-Temperature Axisymmetric Boundary Layer Flows,” Journal of Thermophysics
and Heat Transfer, Vol. 16, No. 4, 2002, pp. 490-497.
59
Ivankov, A. A., “A Method for Calculating Intense Injection from Surfaces of Blunt
Bodies Entering Planetary Atmospheres at Hypersonic Velocities,” Fluid Dynamics, Vol.
41, No. 1, 2006, pp. 137-146.
60
Candler, C. V., and MacCormack, R. W., “The Computation of Hypersonic Ionized
Flows in Chemical and Thermal Nonequilibirum,” Journal of Thermophysics and Heat
Transfer, Vol. 5, No. 3, 1991, pp. 266-273.
61
Thompson, R. A., and Gnoffo, P. A., “Implementation of a Blowing Boundary
Condition in the LAURA Code,” AIAA-2008-1243, 46th AIAA Aerospace Sciences
Meeting and Exhibit, Reno, NV, Jan 7-10, 2008.

198
62
Conti, R. J., MacCormack, R. W., Groener, L. S., and Fryer, J. M., “Practical Navier-
Stokes Computation of Axisymmetric Reentry Flowfields with Coupled Ablation and
Shape Change,” AIAA-1992-752, 30th AIAA Aerospace Sciences Meeting. Reno, NV,
Jan. 6-9, 1992.
63
Bhutta, B. A., and Lewis, C. H., “New Technique for Low-to-High Altitude
Predictions of Ablative Hypersonic Flowfields,” Journal of Spacecraft and Rockets, Vol.
29, No. 1, 1992, pp. 35- 50.
64
Olynick, D., Chen, Y. K., and Tauber, M. E., “Forebody TPS Sizing with Radiation
and Ablation for the Stardust Sample Return Capsule,” AIAA-1997-2474, 32nd AIAA
Thermophysics Conference, Atlanta, GA, June 23-25, 1997.
65
Kuntz, D. W., Hassan, B., and Potter, D. L., “Predictions of Ablating Hypersonic
Vehicles Using an Iterative Coupled Fluid/Thermal Approach,” Journal of
Thermophysics and Heat Transfer, Vol. 15, No. 2, 2001, pp. 129-139.
66
Suzuki, T., Furudate, M., and Swada, K., “Unified Calculation of Hypersonic
Flowfield for a Reentry Vehicle,” Journal of Thermophysics and Heat Transfer, Vol. 16,
No. 1, 2002, pp. 94-100.
67
Cybyk, B. Z., Hunter, L. W., Drewry, D. G., and Van Wie, D. W., “A Unified
Methodology for Simulation of Aerothermochemistry at Gas/Solid Interfaces,” AIAA-
02-1086, 40th AIAA Aerospace Sciences Meeting & Exhibit, Reno, NV, Jan. 14-17, 2002.
68
Daimon, Y., Shimada, T., Tsuboi, N., and Fujita, K., “Predictions of Nozzle Shape
Change Using a Coupled Fluid/Thermochemical Approach,” AIAA 2007-5778, 43rd
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Cincinnati, OH, July
8-11, 2007.
69
Ayasoufi, A., Rahmani, R. K., Cheng, G., Koomullil, R., and Neerookar, K.,
“Numerical Simulation of Ablation for Reentry Vehicles,” AIAA 2006-2908, 9th
AIAA/ASME Joint Thermophysics and Heat Transfer Conference, San Francisco, CA, 5-8
June 2006.
70
Gosse, R., and Candler, G., “Ablation Modeling of Electro-Magnetic Launched
Projectile for Access to Space,” AIAA 2007-1210, 45th AIAA Aerospace Science
Meeting, Reno, NV, Jan. 8-11, 2007.
71
Prakash, A., Zhong, X., “Numerical Simulation of Planetary Reentry Aeroheating over
Blunt Bodies with Non-equilibrium Reacting Flow,” AIAA 2008-744, 46th AIAA
Aerospace Sciences Meeting, Reno, NV, Jan. 7-10, 2008.
72
White, F. W., Fluid Mechanics, 4th ed., McGraw Hill, Boston, 2000.

199
73
Kee, R. J., Cotrin, M. E., and Glarborg, P., Chemically Reacting Flow: Theory and
Practice, John Wiley & Sons, Hoboken, NJ, 2003.
74
Tannehill, J. C., Anderson, D. A., and Pletcher, R. H., Computational Fluid Mechanics
and Heat Transfer, 2nd ed., Taylor & Francis, Washington, DC, 1997.
75
Anderson, J. D., Hypersonic and High-Temperature Gas Dynamics, 2nd ed., AIAA,
Reston, VA, 2006.
76
Gnoffo, P. A., Gupta, R. N., and Shinn, J. L., “Conservation Equations and Physical
Models for Hypersonic Air Flows in Thermal and Chemical Nonequilibrium,” NASA
Technical Paper 2867, 1989.

Esch, D. D., Siripong, A., and Pike, R. W., “Thermodynamic Properties of Polynomial
77

Form for Carbon, Hydrogen, Nitrogen, and Oxygen Systems from 300 to 15000°K,”
NASA CR-111989, 1970.
78
NIST, Chemistry WebBook, [online database], URL: http://webbook.nist.gov/.
79
Chen, Y. K, and Milos, F. S., “Finite-Rate Ablation Boundary Conditions for a Carbon-
Phenolic Heat-Shield,” AIAA 2004-2270, 37th AIAA Thermophysics Conference,
Portland, OR, June 28 – July 1, 2004.
80
Mullenix, N., Povitsky, A., “Comparison of Hertz-Knudsen and Ytrehus carbon
ablation rates using a Reactive-Riemann solver,” AIAA Paper 2008-1222, 46th AIAA
Aerospace Sciences Meeting and Exhibit, Reno, NV, Jan. 7-10, 2008.
81
Mullenix, N., and Povitsky, A., “Comparison of 1D and 2D coupled models of gas
dynamics and heat transfer for the laser ablation of carbon,” Journal of Computational
and Theoretical Nanoscience, Vol. 3, No. 4, 2006, pp. 513-524.
82
Pathak, K., Mullenix, N., and Povitsky, A., “Combined Thermal and Gas Dynamics
Numerical Model for Laser Ablation of Carbon,” Journal of Nanoscience and
Nanotechnology, Vol. 6, No. 12, 2006, pp. 1271-1280.
83
Mullenix, N., and Povitsky, A., “Exploration of Pulse Timing for Multiple Laser Hits
within a Combined Heat Transfer, Phase Change, and Gas Dynamics Model for Carbon
Ablation,” Applied Surface Science, Vol. 253, No. 15, 2007, pp. 6636-6370.
84
Hirsch, C., Numerical Computation of Internal and External Flows. Volume 2:
Computational Methods for Inviscid and Viscous Flows, John Wiley & Sons, Chichester,
England, 1990.
85
Park, C., Nonequilibrium Hypersonic Aerodynamics, Wiley Intersciences, New York,
1990.

200
86
Toro, E. F., Riemann Solvers and Numerical Methods for Fluid Dynamics, Springer-
Verlag, Berlin, 1997.
87
Le Metayer, O., Massoni, J., and Saurel, R., “Modeling Evaporation fronts with
reactive Riemann solvers,” Journal of Computational Physics, Vol. 205, No. 2, 2005, pp.
567-610.
88
Ben-Artzi, M., and Falcovitz, J., Generalized Riemann Problems in Computational
Fluid Dynamics, Cambridge University Press, Cambridge, UK, 2003.
89
Ortega, J. M., and Rheinboldt, W.C., Iterative Solution of Nonlinear Equations in
Several Variables, Academic Press, New York, 1970.
90
Incropera, F. P., and Dewitt, D. P., Introduction to Heat Transfer, 4th ed., John Wiley
& Sons, New York, 2002.
91
Hoffmann, K, A., Chiang, S. T., Siddiqui, S., and Papadakis, M., Fundamental
Equations of Fluid Mechanics, Engineering Education Systems, Wichita, KS, 1996.
92
Mullenix, N., and Povitsky, A., “Parallel Implementation of a Tightly Coupled
Ablation Prediction Code Using MPI,” IEEE Cluster 2009, New Orleans, LA, Aug. 31-
Sep. 4, 2009.
93
The Ohio Supercomputer Center website [online] URL: http://www.osc.edu.
94
Anderson, J. D., Modern Compressible Flow: With Historical Perspective, 2nd ed.,
McGraw-Hill, Boston, 1990.
95
Kays, W. M., and Crawford, M. E., Convective Heat and Mass Transfer, 3rd ed.,
McGraw-Hill, New York, 1993.
96
Ghia, U., Ghia, K. N., and Shin, C. T., “High-Re Solutions for Incompressible Flow
Using the Navier-Stokes Equations and a Multigrid Method,” Journal of
Computatational Physics, Vol. 48, 1982, pp. 397-411.
97
Mullenix, N., Povitsky, A., and Gaitonde, D., “Modeling of Local Intense Ablation in
Hypersonic Flight,” AIAA Paper 2008-2555, 15th AIAA International Space Planes and
Hypersonic Systems and Technologies Conference, Dayton, Ohio, Apr. 28-May 1, 2008.

201
APPENDICES

202
APPENDIX A

LIST OF SYMBOLS

Latin

A Jacobian
Ai,s,r Coefficients for polynomial curve-fits
c Vector of mass concentrations
Cf,r, Ef,r, nf,r Parameters for forward reaction rate coefficients
CFL Courant-Friedrichs-Levy number
Cv,s Specific heat at constant volume for species s
Cp,s Specific heat at constant pressure for species s
Cs Molar concentration of species s (moles of s per unit mass of mixture)
cs Mass concentration of species s (mass of s per unit mass of mixture)
Cs Control surface
Cv Control volume
Ds Mass diffusion coefficient of species s
Ds,r Binary mass diffusion coefficient between species s and r
e Internal energy per unit mass
eij Strain
eT Total energy (internal plus kinetic) per unit mass
f Vector of body forces
F Flux vector in the x1 direction
F-I, G-I Flux calculation Zones
G Flux vector in the x2 direction
h Enthalpy per unit mass
h0 Enthalpy of formation per unit mass
203
h1, h2, h3 Metrics of coordinate transformation
J Jacobian
k Thermal conductivity
kb,r Backward reaction rate coefficient of reaction r
Kc,r Equilibrium constant of reaction r
kf,r Forward reaction rate coefficient of reaction r
L Matrix of left eigenvectors
m Mass
M Molar mass (kg/kmol)
~
m abl Overall ablation mass flux vector (mass/(time*area))
~
m s Species ablation mass flux vector
N Extensive property of a system
n Unit normal to control surface
p Pressure tensor
p Pressure
R Geometrical source term vector
R Specific gas constant
R Universal gas constant
R Matrix of right eigenvectors
Rb,r Backward reaction rate of reaction r
Rf,r Forward reaction rate of reaction r
S Physical source term vector
S Speed ratio in Ytrehus model
SHR Right rarefaction head speed
SL Left shock speed
STR Right rarefaction tail speed
t Time
T Temperature
U Vector of conservative unknowns
ud Velocity component in direction d
ud, inj Component of injection velocity in direction d

204
ud, s Diffusion velocity of species s in direction d
ud, surf Component of surface velocity in direction d
V Velocity vector relative to inertial frame of reference
V Volume
v Specific volume (volume per unit mass)
Vinj Velocity vector of injected gas relative to inertial frame of reference
Vr, solid Velocity vector relative to motion of solid surface
Vsurf Velocity vector of surface relative to inertial frame of reference
W Vector of primitive variables
w s Mass production rate of species s (mass/(time*volume))
x, y, z Cartesian coordinates
x1, x2, x3 General orthogonal coordinates
Z Pressure ratio in Ytrehus model

Greek

 Speed of Sound
diff Thermal diffusivity
s, r Stoichiometric coefficient of species s on reactant side of reaction r
 Change in pressure with respect to change in total energy
 Amplitude function in Ytrehus model
s, r Stoichiometric coefficient of species s on product side of reaction r
 Eigenvalue function
1s ,r Modified collision integral
t Time step
 Dissipation function
 Arbitrary scalar
 Ratio of specific heats
 Intensive property of a system
 Matrix of eigenvalues
 Eigenvalue

205
 Viscosity
 Momentum diffusivity
s1,r,1 Collision integral
 Temperature ratio in Ytrehus model
 density
 Stress Tensor
 Shear stress tensor
 Ratio of solid to overall volume
 Local non-dimensional spatial coordinates

Superscripts

±
Variables split with sign of eigenvalues
L
Left side of interface
n
Time level n
n+1
Time level n+1
poly
Polynomial curve fit
R
Right side of interface

Subscripts

_iv Inviscid
_md Mass diffusion
_vh Viscous and heat transfer
chem Gas phase chemical reactions
cond Conduction
d Direction (1, 2, 3)
gas Term is evaluated within the gas
i±w, j±w Solid finite volume faces
i±½, j±½ Overall finite volume faces
L, R Left or right side of interface
r Reaction r
206
s Species s
solid Term is evaluated within the solid
surf Term is evaluated on the solid surface
surf Surface chemical reactions
O Overall
visc Viscous

207
APPENDIX B

CONSERVATION EQUATION DERIVATIONS

The particular forms of the Reynolds’ Transport Theorem (RTT) for the non-

charring ablation problem were derived in section 2.1.2. For the solid, the RTT is given

by:

dN solid 
t Cv
 dV   Vr ,solid  n solid dA , (2.1.6)
dt solid Cssolid

where

Vr ,solid  Vinj  Vsurf . (2.1.7)

For the gas, the RTT is given by:

dN gas 
dt
  dV  Cs V  nO dA  Cs Vr ,solid  n soliddA .
t Cv
(2.1.10)
gas O solid

B.1 Global Continuity Equation

The global continuity equation can be found using the principle of conservation of

mass. Assuming there are no nuclear reactions, this states that the mass of any system

stays constant, or equivalently that the time rate of change of the mass of any system is

zero:

208
dmsys
0. (B.1.1)
dt

This applies to all three of the systems under consideration.

The RTT can be used to convert Eq. B.1.1 into a form suitable for use on control

volumes, by setting N equal to m (mass) and  = 1 in Eq. 2.1.6 to provide the global

continuity equation for Cvsolid:


dV    Vinj  Vsurf  n solid dA  0 ,
t Cv
(B.1.2a)
solid Cssolid

or Eq. 2.1.10 for Cvgas:


 dV  Cs V  nO dA  Cs  Vinj  Vsurf  n soliddA  0
t Cv
(B.1.2b)
gas O solid

B.2 Species Continuity Equation

The materials contained within any of the systems defined above are arbitrary,

and can be composed of many different species. Consider that the overall system, SO, is

composed of Ns species. The total mass of a system is equal to the sum of the different

species:

Ns
msys   ms , (B.2.1)
s 1

where the subscript s denotes each individual species. For any system, the conservation

of mass can then be written as:

dmsys d  Ns 
   ms   0 , (B.2.2)
dt dt  s 1 

209
which does not mean that the time rate of change for each individual species is identically

zero.

The RTT can be applied to the time rate of change in mass for each individual

species by setting N equal to ms, and  equal to the mass concentration cs (s/which still

has units of 1). A new velocity vector, Vrs, must be defined to account for any variations

in velocity of each individual species caused by gradients in concentration73:

Vrs  V  Vs , (B.2.3)

where Vs is the diffusion velocity for species s. Finally the time rate of change in mass of

species s can be defined as:

dms
  w s dV , (B.2.4)
dt Cv

where w s is the specific mass production in units of mass per unit volume per unit time.

For species s, the species continuity equation for Cvsolid is:


 s dV    s Vinj  Vsurf  Vs  n solid dA  0 ,
t Cv
(B.2.5a)
solid Cssolid

if chemical reactions are assumed to not take place within the solid. For Cvgas the species

continuity equation is:


  s dV  Cs  s V  Vs  nO dA
t Cv
  w s dV . (B.2.5b)
  V  Vsurf  Vs  n solid dA
gas O

 s inj CvI
Cssolid

210
B.3 Momentum Equation for Gas

The development of the momentum equation follows from Newton’s 2nd Law,

which states that the change in momentum for a system is equal to the sum of the forces

acting on the system:

d mV sys
  Fsys . (B.3.1)
dt

The forces acting on a system can be split into body forces and forces acting on the

surface:

F sys   Fbody   Fsurface . (B.3.2)

The body forces, Fbody, which could include gravity, electromagnetic or other such forces,

can be written as:

F body   fdV , (B.3.3)


Cv

where f is a specific force term with units of force per unit mass. The surface forces,

Fsurface, can be written as the dot product of the stress tensor, , with the unit outward

normal integrated along a surface:

F surface   σ  ndA . (B.3.4)


Cs

The stress tensor can be split into a pressure tensor, p, and a shear stress tensor, :

 p 0 0   11  12  13 
σ    0 p 0    12  22  23  , (B.3.5)
 0 0 p   13  23  33 
 
p τ

where (under the assumption of a Newtonian fluid), the individual ij can be written as 74:

211
 2e11  e22  e33 
2
 11 
3
 22   2e22  e11  e33 
2
3
 33   2e33  e22  e11  ,
2 (B.3.5b)
3
 12  e12
 13  e13
 23  e23

where the eij’s will be defined later. Eq. B.3.4 can then be written as:

F surface    p  ndA   τ  ndA . (B.3.6)


Cs Cs

The RTT can be used to write the momentum equation for a control volume by

setting N equal to mV (linear momentum), and  equal to V. For Cvgas, the momentum

equation is:


 VdV  Cs VV  nO dA
t Cv
 fdV   p  n dA   τ  n dA
0 0


Cvgas CsO CsO
. (B.3.7)
 VV  Vsurf  n solid dA
gas O

 inj

 p  n solid dA 
Cssolid
 τ  n solid dA
Cssolid
Cssolid

212
B.4 Energy Equations

The derivation of the energy equation begins with the first law of thermodynamics

for a system 72:

d ET sys d Q sys d W sys


  , (B.4.1)
dt dt dt

where ET represents the total energy within the system, Q represents heat added to the

system, and W represents work done on the system.

The total amount of heat added to the system is the sum of conduction heat

transfer, Q cond , heat transfer caused by mass diffusion, Q md , radiation heat transfer, Q rad ,

and other heat generation, Q gen . The conduction heat transfer is written for a generic

control volume using Fourier’s Law 73:

Q conduction   kT  ndA     kT dV , (B.4.2)


Cs Cv

where k is the thermal conductivity, and T is the temperature. The heat transfer caused by

mass diffusion is defined by:

N N
Q md    hs  s Vs  ndA      hs  s Vs dV , (B.4.3a)
s 1 Cs s 1 Cv

where hs is the enthalpy of species s. Equation B.4.3a can be rewritten by using the

definition of enthalpy, h = et +p/as:

 p  
 
Q md        et  s   s   s Vs dV       s eT  s Vs      p sVs  dV ,
N N
(B.4.3b)
s 1 Cv  s   s 1 Cv

where ps is the partial pressure of species s. For now the other two terms will be defined

as the integral of a volumetric rate ( q ):

213
Q rad   qrad dV , (B.4.4a)
Cv

and

Q gen   q gendV . (B.4.4b)


Cv

The rate of work done on a system (neglecting shaft work) can be defined as the

dot product of the velocity vector and stress tensor integrated along the surface of the

control volume 72:

W   σ  Vr   ndA   p  Vr   ndA   τ  Vr   ndA , (B.4.5)


Cs Cs Cs

where Vr can be either V or Vr,surf

Using the above definitions for the time rates of heat addition and work, the RTT

can be used to define the energy equation by setting N equal to ET, and  equal to eT. For

Cvsolid, the energy equation can be written as:

 kT  n
Cssolid
solid dA

  e    p V  n
N
 dA
 s T s s s solid

t Cv
eT dV s 1 Cssolid

   q dV   q dV , (B.4.6a)


 eT Vinj  Vsurf  n soliddA
solid rad gen
 Cvsolid Cvsolid

Cssolid   p  V  V  n dA
Cssolid
inj surf solid

  τ  V
Cssolid
inj  Vsurf  n solid dA

214
and for Cvgas it can be written as:

 kT  n
CsO
O dA   kT  n
Cssolid
solid dA

    s eT  s   ps Vs  n O dA
N

s 1 CsO

  e    p V  n
N

  s T s s s solid dA
 eT dV  Cs eT V  nO dA
t Cv
s 1 Cssolid

   q rad dV   q gendV . (B.4.6b)


 eT Vinj  Vsurf  n solid dA
gas O

 Cvgas Cvgas

Cssolid   p  V   n O dA   τ  V   n O dA
CsO CsO

  p  V
Cssolid
inj  Vsurf  n solid dA

  τ  V
Cssolid
inj  Vsurf  n solid dA

215
APPENDIX C

MATERIAL PROPERTIES

The species that will be considered include O2, N2, O, N, NO, NO+, e-, CO, CN,

C3, and (C)solid. Unless otherwise noted the properties for O2, N2, O, N, NO, NO+, and e-

come from 76, and for the carbon containing species 78.

Table C.1: Molecular Weights and Enthalpies of Formation


s Ms (kg/kmol) h o (J/kg)
O2 31.9980 0
N2 28.0135 0
O 15.9994 1.55739E+07
N 14.0067 3.37468E+07
NO 30.0061 3.00909E+06
+
NO 30.0056 3.30926E+07
-
e 0.0005486 0
CO 28.0101 -3.94597E+06
CN 26.0174 1.67248E+07
C3 36.0321 2.27593E+07

216
C.1 Specific Heat Curve Fits

The specific heat at constant pressure is provided by a fourth order polynomial

given by 76:


, s T   Rs A1, s  A2, sT  A3, sT  A4, sT  A5, sT
C ppoly 2 3 4
, (C.1.1)

where the coefficients Ai,s of are also functions of temperature:

 Ai , s ,1 if T  800 K
  T  800  T  800
 1   Ai , s ,1  Ai , s , 2 if 800 K  T  1200 K
  400  400
Ai , s T    Ai , s , 2 if 1200 K  T  5500 K , (C.1.2)
 T  5500  T  5500
1   Ai , s , 2  Ai , s ,3 if 5500 K  T  6500 K
  1000  1000
 Ai , s ,3 if 6500 K  T  15000 K

where the Ai,s,r are found in Table C.1.

217
Table C.2: Coefficients for Specific Heat Polynomial Curve Fits
s r A 1,s,r A 2,s,r A 3,s,r A 4,s,r A 5,s,r
O2 1 3.6255980E+00 -1.8782180E-03 7.0554540E-06 -6.7635130E-09 2.1555990E-12
2 3.6219530E+00 7.3618260E-04 -1.9652200E-07 3.6201550E-11 -2.8945620E-15
3 3.7210000E+00 4.2540000E-04 -2.8350000E-08 6.0500000E-13 -5.1860000E-18
N2 1 3.6748260E+00 -1.2081500E-03 2.3240100E-06 -6.3217550E-10 -2.2577250E-13
2 2.8963190E+00 1.5154860E-03 -5.7235270E-07 9.9807390E-11 -6.5223550E-15
3 3.7270000E+00 4.6840000E-04 -1.1400000E-07 1.1540000E-11 -3.2930000E-16
O 1 2.9464280E+00 -1.6381660E-03 2.4210310E-06 -1.6028430E-09 3.8906960E-13
2 2.5420590E+00 -2.7550610E-05 -3.1028030E-09 4.5510670E-12 -4.3680510E-16
3 2.5460000E+00 -5.9520000E-05 2.7010000E-08 -2.7980000E-12 9.3800000E-17
N 1 2.5030710E+00 -2.1800180E-05 5.4205280E-08 -5.6475600E-11 2.0999040E-14
2 2.4502680E+00 1.0661450E-04 -7.4653370E-08 1.8796520E-11 -1.0259830E-15
3 2.7480000E+00 -3.9090000E-04 1.3380000E-07 -1.1910000E-11 3.3690000E-16
NO 1 4.0459520E+00 -3.4181780E-03 7.9819190E-06 -6.1139310E-09 1.5190700E-12
2 3.1890000E+00 1.3382280E-03 -5.2899320E-07 9.5919330E-11 -6.4847930E-15
3 3.8450000E+00 2.5210000E-04 -2.6580000E-08 2.1620000E-12 -6.3180000E-17
NO+ 1 3.6685060E+00 -1.1544580E-03 2.1755610E-06 -4.8227470E-10 -2.7847900E-13
2 2.8885490E+00 1.5217120E-03 -5.7531240E-07 1.0051080E-10 -6.6044290E-15
3 2.2141700E+00 1.7760600E-03 -4.3038600E-07 4.1737700E-11 -1.2828900E-15
e- 1 2.50E+00 0 0 0 0
2 2.50E+00 0 0 0 0
3 2.5080000E+00 -6.3320000E-06 1.3640000E-09 -1.0940000E-13 2.9340000E-18
CO 1 3.7870000E+00 -2.1710000E-03 5.0760000E-06 -3.4740000E-09 7.7220000E-13
2 3.2540000E+00 9.6980000E-04 -2.6470000E-07 3.0370000E-11 -1.1770000E-15
3 3.3660000E+00 8.0720000E-04 -1.9680000E-07 1.9400000E-11 -5.5490000E-16
CN 1 3.3530000E+00 -2.7630000E-03 6.8570000E-06 -5.4130000E-09 1.4910000E-12
2 3.4110000E+00 4.8970000E-04 1.0050000E-07 -3.4730000E-11 2.3610000E-15
3 3.4730000E+00 7.3370000E-04 -9.0880000E-08 4.8470000E-12 -1.0180000E-16
C3 1 2.6330000E+00 9.4190000E-03 -9.5930000E-06 5.5800000E-09 -1.4240000E-12
2 4.0020000E+00 3.5450000E-03 -1.3180000E-06 2.0640000E-10 -1.1440000E-14
3 2.2130000E+01 -1.4590000E-02 5.5650000E-06 6.7580000E-10 2.8250000E-14
Csolid all -2.6896041E+02 4.0723033E+00 -2.5273539E-03 5.3492230E-07 0
77 37
Note: Gas phase species are provided by . Csolid is provided by .

C.2 Gas Phase Chemical Reactions

Consider a reduced gas phase model that contains ten species: O2, N2, O, N, NO,

NO+, e-, CO, CN, C3. The first two species are present in ambient air, and the next five

are results of dissociation reactions, ionization, and exchange reactions involving NO.

The last three species are products of surface oxidation, surface nitridation, and

sublimation. Thirty reactions are to be accounted for in this model; Table C.4 shows the

reactions and their associated reaction numbers. The coefficients for the forward reaction

rate equations 40:

218
  E f ,r 
 
k f ,r T   C f ,rT e
n  k BT 
, (C.2.1)

are given in Table C.5. The backward reaction rate is found in terms of the forward and

equilibrium rates76:

k f ,r T 
kb,r T   , (C.2.2)
K c ,r T 

where the equilibrium constant, Kc,r is given by 40:

 A 
K c ,r  exp 
 1E 
1
4
 T
 
 A2  A3 ln 1E 4  A4 1E 4  A5 1E 4
T T
    ,
2
(C.2.3)
 T 

with coefficients given in Table C.6.

219
Table C.3: Reaction Number Designations.
Dissociation
O 2  M  2O  M N 2  M  2N  M NO  M  O  N  M
M  r M  r M  r
O2 1 O2 10 O2 19
N2 2 N2 11 N2 20
O 3 O 12 O 21
N 4 N 13 N 22
NO 5 NO 14 NO 23
NO 
6 NO 
15 NO 24
CO 7 CO 16 CO 25
CN 8 CN 17 CN 26
C3 9 C3 18 C3 27
Exchange r
O  N 2  NO  N 28
NO  O  O 2  N 29
Ionization r
N  O  NO  e  30

Table C.4: Forward Reaction Rate Coefficients.


Reaction M r C f,r n f,r E f,r /k source
O 2  M  2O  M 40
molecules 1-2,5-9 2.00E+21 -1.5 59360
40
atoms 3,4 1.00E+22 -1.5 59360
N 2  M  2N  M 40
molecules 10-11,14-18 7.00E+21 -1.6 113200
40
atoms 12, 13 3.00E+22 -1.6 113200
NO  M  O  N  M all 19-27 7.95E+23 -2.0 75500 76

O  N 2  NO  N 40
28 5.70E+12 -0.42 42938
NO  O  O 2  N 40
29 8.40E+12 0 19400
N  O  NO   e  30 5.30E+12 0 31900 40

220
Table C.5: Equilibrium Constant Coefficients
Reaction A1 A2 A3 A4 A5 source
O 2  M  2O  M 40
1.578640 2.688744 4.215573 -8.091354 0.174260
N 2  M  2N  M 40
-3.293682 0.998998 -8.237028 -5.526183 -0.582174
NO  M  O  N  M 0.792000 -0.492000 -6.761000 0.091000 0.004000 76

O  N 2  NO  N 40
-3.032189 0.078468 -7.693047 1.411299 -0.517448
NO  O  O 2  N 40
1.840133 -1.768215 -4.759554 1.153812 -0.238985
N  O  NO   e  3.429239 -7.431449 6.012721 -8.276563 0.503539
40

C.3 Collision Integrals

In general, values of log10 srk ,k  T  at T = 2000K, and 4000K are given in the

literature, so interpolation is required, with a formula of 76:

log10 srk ,k  T   log10 srk ,k  2000


log10 srk ,k  4000  log10 srk ,k  2000 . (C.3.1)
 * ln T   ln 2000
ln 4000  ln 2000

These value of log10 srk ,k  T  are given in Table C.8. Most of the values come from 76,

with the exception of CO, CN, and C3 which come from 40.

221
Table C.6: Values of the Collision Integrals
s  O2 s  N2

log10 sr T 
1,1
 
log10 sr T 
2, 2 
 
log10 sr T 
1,1
  
log10 sr2, 2  T 
r 2000 K 4000 K 2000 K 4000 K r 2000 K 4000 K 2000 K 4000 K
O2  14.60  14.64  14.54  14.57 O 2  14.58  14.63  14.51  14.54
N2  14.58  14.63  14.51  14.54 N 2  14.56  14.65  14.50  14.58
O  14.69  14.76  14.62  14.69 O  14.63  14.72  14.55  14.66
N  14.66  14.74  14.59  14.66 N  14.67  14.75  14.59  14.66
NO  14.59  14.63  14.52  14.56 NO  14.57  14.64  14.51  14.54
NO  14.34  14.46  14.38  14.50 NO

 14.34  14.46  14.38  14.50
e-  15.52  15.39  15.52  15.39 e  15.11  15.02  15.11  15.02
-

CO  14.62  14.74  14.57  14.69 CO  14.59  14.72  14.54  14.66


CN  14.61  14.73  14.56  14.68 CN  14.58  14.70  14.53  14.65
C3  14.61  14.74  14.57  14.69 C3  14.59  14.71  14.54  14.66

sO sN

log10 sr1,1 T  
log10 sr2, 2  T   
log10 sr1,1 T   
log10 sr2, 2  T 
r 2000 K 4000 K 2000 K 4000 K r 2000 K 4000 K 2000 K 4000 K
O2  14.69  14.76  14.62  14.69 O2  14.66  14.74  14.59  14.66
N2  14.63  14.72  14.55  14.64 N2  14.67  14.75  14.59  14.66
O  14.11  14.14  14.71  14.79 O  14.76  14.86  14.69  14.80
N  14.76  14.86  14.69  14.80 N  14.08  14.11  14.74  14.82
NO  14.66  14.74  14.59  14.66 NO  14.66  14.75  14.67  14.66
NO  14.34  14.46  14.38  14.50 NO  14.34  14.46  14.38  14.50
e-  15.94  15.82  15.94  15.82 e-  15.30  15.30  15.30  15.30
CO  14.67  14.79  14.62  14.74 CO  14.65  14.78  14.60  14.72
CN  14.66  14.78  14.61  14.73 CN  14.64  14.76  14.69  14.80
C3  14.67  14.79  14.62  14.75 C3  14.65  14.77  14.60  14.72

222
Table C.6: Values of the Collision Integrals. cont.

s  NO s  NO

log10 sr T 
1,1
 
log10 sr T 
2, 2 
 
log10 sr T 
1,1
  
log10 sr2, 2  T 
r 2000 K 4000 K 2000 K 4000 K r 2000 K 4000 K 2000 K 4000 K
O2  14.59  14.63  14.52  14.56 O2  14.34  14.46  14.38  14.50
N2  14.57  14.64  14.51  14.56 N2  14.34  14.46  14.38  14.50
O  14.66  14.74  14.59  14.66 O  14.34  14.46  14.38  14.50
N  14.66  14.75  14.67  14.66 N  14.34  14.46  14.38  14.50
NO  14.58  14.64  14.52  14.56 NO  14.18  14.22  14.38  14.50
NO  14.18  14.22  14.38  14.50 NO  11.70  12.19  11.49  11.98
e-  15.30  15.08  15.30  15.08 e-  11.70  12.19  11.49  11.98
CO  14.61  14.73  14.56  14.68 CO  14.34  14.46  14.38  14.50
CN  14.60  14.72  14.55  14.67 CN  14.34  14.46  14.38  14.50
C3  14.60  14.73  14.56  14.68 C3  14.34  14.46  14.38  14.50

s  e s  CO

log10 sr T 
1,1
 
log10 sr T 
2, 2 
  
log10 sr1,1 T   
log10 sr2, 2  T 
r 2000 K 4000 K 2000 K 4000 K r 2000 K 4000 K 2000 K 4000 K
O2  15.52  15.39  15.52  15.39 O2  14.62  14.74  14.57  14.69
N2  15.11  15.02  15.11  15.02 N2  14.59  14.72  14.54  14.66
O  15.94  15.82  15.94  15.82 O  14.67  14.79  14.62  14.74
N  15.30  15.30  15.30  15.30 N  14.65  14.78  14.60  14.72
NO  15.30  15.08  15.30  15.08 NO  14.61  14.73  14.56  14.68
NO  11.70  12.19  11.49  11.98 NO  14.34  14.46  14.38  14.50
e-  11.70  12.19  11.49  11.98 e-  15.29  15.06  15.29  15.06
CO  15.29  15.06  15.29  15.06 CO  14.60  14.72  14.55  14.67
CN  15.29  15.06  15.29  15.06 CN  14.59  14.71  14.54  14.66
C3  15.29  15.06  15.29  15.06 C3  14.59  14.71  14.55  14.67

223
Table C.6: Values of the Collision Integrals. cont.

s  CN s  C3

log10 sr T 
1,1
 
log10 sr T 
2, 2 
 
log10 sr T 
1,1
  
log10 sr2, 2  T 
r 2000 K 4000 K 2000 K 4000 K r 2000 K 4000 K 2000 K 4000 K
O2  14.61  14.73  14.56  14.68 O2  14.61  14.74  14.57  14.69
N2  14.58  14.70  14.53  14.65 N2  14.59  14.71  14.54  14.66
O  14.66  14.78  14.61  14.73 O  14.67  14.79  14.62  14.75
N  14.64  14.76  14.59  14.71 N  14.65  14.77  14.60  14.72
NO  14.60  14.72  14.55  14.67 NO  14.60  14.73  14.56  14.68
NO  14.34  14.46  14.38  14.50 NO  14.34  14.46  14.38  14.50
e-  15.29  15.06  15.29  15.06 e-  15.29  15.06  15.29  15.06
CO  14.59  14.71  14.54  14.66 CO  14.59  14.71  14.55  14.67
CN  14.59  14.70  14.52  14.64 CN  14.58  14.70  14.53  14.65
C3  14.58  14.70  14.53  14.65 C3  14.58  14.71  14.53  14.66

C.4 Solid Thermal Conductivity

The thermal conductivity of graphite is given by 37:

ksolid T   exp  1.02 ln T   13.30 . (C.4.1)

224
APPENDIX D

NUMERICAL SCHEMES

This appendix contains descriptions of auxiliary numerical schemes and methods.

Section D.1 contains the spatial discretization schemes for primitive variables. Section

D.2 contains the description of the Gauss-Seidel method which is used extensively to

solve systems of equations.

D.1 Discretization Schemes for Primitive Variables

In the case where the finite volume contains only gas (i,j = 0) the difference

equations take the form of:

  gas i 1, j   gas i , j


 if  i 1, j  0
  gas   x1 i 1, j  x1 i , j
  
 x1 i 1 2, j 
 gas i 1 w, j   gas i , j
, (D.1.1a)
if  i 1, j  0
 x1    x 
  i 1  w , j 1 i , j

  gas i , j   gas i 1 2, j
 if  i 1, j  0
  gas   x1 i , j  x1 i 1 2, j
  
 x1 i 1 2, j 
 gas i , j   gas i 1 w, j
, (D.1.1b)
if  i 1, j  0
 x1   x1 
 i , j  i 1  w , j

225
  gas i , j 1   gas i , j
 if  i , j 1  0
  gas   x2 i , j 1  x2 i , j
  
 x2 i , j 1 2  
 gas i , j 1 w
  gas i , j
, (D.1.1c)
if  i , j 1  0
  x2   x2 i , j
 i , j 1 w

 gas i , j  gas i , j 1
 if  i1, j  0
 gas   x2 i , j  x2 i , j 1
  
 x2 i1 2, j    gas i, j1w
 gas i , j
, (D.1.1d)
if  i1, j  0
 x2 i , j  x2 i , j 1 w

 gas  1     gas  
    gas     , (D.1.1e)
 x1 i , j 2  x1 i 1 2, j  x1 i 1 2, j 

 gas  1     gas  
    gas     , (D.1.1f)
 x2 i , j 2  x2 i , j 1 2  x2 i , j 1 2 

 1   gas    gas  
       if  i 1, j  0
 2  x2 i 1, j  x2 i , j 
 

  gas   x1 i 1 w, j  x1 i 1 2, j   gas 
  
       , (D.1.1g)
 x2 i 1 2, j   x1  i 1  w , j
  x1 i , j
 x2 
    
 if  i 1, j  0
i 1 w , j

  
  1 i 1 2, j
x   x 
1 i, j
 
 gas  
   x   
 x1 i , j  x2 i , j
 
  1 i 1 w, j 

 1   gas    gas  
       if  i 1, j  0
 2  x2 i 1, j  x2 i , j 
 

  gas   x1 i 1 2, j  x1 i 1 w, j   gas 
  
       , (D.1.1h)
 x2 i 1 2, j   x1 i , j
  x1  i 1  w , j
 x2 
    
 if  i 1, j  0
i 1 w , j

 
  1 i , j
x    x 
1 i 1 2, j
 
 gas  
   x   x    x  
1 i 1 w, j  2 i , j
  1 i , j 

226
 1   gas    gas  
       if  i , j 1  0
 2  x1 i , j 1  x1 i , j 
 

  gas    x2 i , j 1 w  x2 i , j 1 2   gas 
 
       , (D.1.1i)

 1 i , j 1 2
x   x 2 i ,  j  1  w
  x 2  i , j
 x1 
 i , j 1 w 
  if  i , j 1  0

    x  
  x  
   

2 i , j 1 2 2 i , j gas
   x  
  x2 i , j  x1 i , j

  2 i , j 1 w 

 1   gas    gas  
       if  i , j 1  0
 2  x1 i , j 1  x1 i , j 
 

  gas    x2 i , j 1 2  x2 i , j 1 w   gas 
 
       , (D.1.1j)

 1 i , j 1 2
x   x 2 i , j
  x 2 i ,  j  1  w
 x1 
 i , j 1 w 
  if  i , j 1  0

    x    x  

   

2 i , j 2 i , j 1 2 gas
   x    x    
2 i , j 1 w  x1  i , j
  2 i , j 

and none of the derivatives are needed for the solid.

For the more general case when i,j is between zero and one, more terms must be

calculated. The following difference equations can be applied for terms at the faces

(i±½,j±½):

  gas i 1, j   gas i  w, j


 if  i 1  0
  gas   x1 i 1, j  x1 i  w, j
  
 x1 i 1 2, j  
 gas i 1 w, j
  gas i  w, j
, (D.1.2a)
if  i 1  0
 x1    x 
 i 1 w, j 1 i  w, j

  gas i  w, j   gas i 1 2, j
 if  i 1, j  0
  gas   x1 i  w, j  x1 i 1 2, j
  
 x1 i 1 2, j    gas i1 w, j
 gas i  w, j
, (D.1.2b)
if  i 1, j  0
 x1    x 
 i  w, j 1 i 1 w, j

227
  gas i , j 1   gas i , j  w
 if  i , j 1  0
  gas   x2 i , j 1  x2 i , j  w
  
 x2 i , j 1 2  
 gas i , j 1 w
  gas i , j  w
, (D.1.2c)
if  i , j 1  0
  x2   x2 i , j  w
 i , j 1 w

  gas i , j  w   gas i , j 1
 if  i 1, j  0
  gas   x2 i , j  w  x2 i , j 1
  
 x2 i 1 2, j    gas i, j 1 w
 gas i , j  w
, (D.1.2d)
if  i 1, j  0
  x2    x 
 i, j w 2 i , j 1 w

 gas  1     gas  
    gas     , (D.1.2e)
 x1 i , j 2  x1 i 1 2, j  x1 i 1 2, j 

 gas  1     gas  
    gas     , (D.1.2f)
 x2 i , j 2  x2 i , j 1 2  x2 i , j 1 2 

  x1    x1   
i 1 2, j   gas 
  i 1, j
  
   x1 i 1, j   x1 i  w, j , j  x2 i 1, j 
   if  i 1, j  0
    x1 i 1 2, j   x1 i  w, j   gas  
   


 
  gas      x1 i 1, j   x1 i  w, j , j  x2 i  w, j 
   , (D.1.2g)

 2 i 1 2, j
x   x1 i 1 w, j   x1 i 1 2, j   gas  
  x    
  1 i 1 w, j   x  
1 i  w, j , j   x 2  i 1 w, j 

   x   if  i 1, j  0

  1 i 1 2, j   x 
1 iw
  
 gas   

   x1 i 1 w, j   x1 i  w, j  x2  
   i  w, j 

228
   x1    x1 i 1, j   gas  
  i 1 2, j   
   x1 i  w, j   x1 i 1, j  x2 i 1, j 
   if  i 1, j  0
   x1 i  w, j   x1 i 1 2, j   gas  
   
  
    x1 i  w, j  x1 i 1, j  x2 i  w, j 
  gas  
   , (D.1.2h)
 x2 i 1 2, j   x1 i 1 2, j   x1 i 1 w, j   gas  
  x    
 
1 i  w, j   x1 i 1 w, j  x2  i 1 w, j 
 
   x   if  i 1, j  0

   1 i  w, j   x 
1 i 1 2, j
  
 gas   

   x1 i  w, j   x1 i 1 w, j  x2  
    i  w , j 

  x2    x2   
i , j 1 2   gas 
  i , j 1
  
   x2 i , j 1  x2 i , j  w  x1 i , j 1 
   if  i , j 1  0
    x2 i , j 1 2   x2 i , j  w   gas  
   
  
    x2 i , j 1   x2 i , j  w  x1 i , j  w 
  gas  
   , (D.1.2i)
 x1 i , j 1 2   
  2 i , j 1 w
x   x 
2 i , j 1 2
   
 gas   
  x      x1 

 2 i , j 1 w
x 2 i, jw  i , j 1 w 

  x   if  i , j 1  0

  
  x  
  gas 
 

2 i , j 1 2 2 i , j w

   x2 i , j 1 w   x2 i , j  w  x1   
   i , j 1 4 

and

  x2    x2 i , j 1   gas  
  i , j 1 2   
   x2 i , j  w   x2 i , j 1  x1 i , j 1 
   if  i , j 1  0
    x2 i , j  w   x2 i , j 1 2   gas  
   


 
  gas      x2 i , j  w   x2 i , j 1  x1 i , j  w 
   , (D.1.2j)

 1 i , j 1 2
x   x2 i , j 1 2   x2 i , j 1 w   gas  
 x    
  2 i, j w   x 
2 i , j 1 w 
 x 1  i , j 1 w 

   x   if  i , j 1  0

   2 i , j 1 4   x 
2 i , j 1 2
 
 gas   

   x2 i , j 1 4   x2 i , j 1 w  x1  
    i , j  w 

The following terms that must be calculated for the gas at the solid surface (face denoted

by i±w or j±w):
229
  gas i , j   gas i  w, j
  gas   if  1 i  w, j  1
    x1 i 1 2, j  x1 i  w, j , (D.1.3a)

 1  i  w, j 
x
 0 if  1 i  w, j  1

  gas i  w, j   gas i , j
  gas   if  1 i  w, j  1
    x1 i  w, j  x1 i 1 2, j , (D.1.3b)

 1  i  w, j 
x
 0 if  1 i  w, j  1

  gas i , j   gas i , j  w
  gas   if  2 i , j 1 4  1
    x1 i , j  x1 i , j  w , (D.1.3c)

 2 i , j  w 
x
 0 if  2 i , j 1 4  1

  gas i , j  w   gas i , j
  gas   if  2 i , j  w  1
    x1 i , j  w  x1 i , j , (D.1.3d)
 x2 i , j  w  0 if  2 i , j  w  1

and in general the derivative along the faces are zero:

 I          
    I    I    I  0. (D.1.3e)
 x2 i  w, j  x2 i  w, j  x1 i , j  w  x1 i , j  w

Finally for any case where the derivatives are needed within the solid:

 solid  solid i  w, j  solid i , j


   , (D.1.4a)
 x1 i  w, j x1 i  w, j  x1 i , j

 solid  solid i , j  solid i w, j


   , (D.1.4b)
 x1 i  w, j x1 i , j  x1 i w, j

 solid  solid i , j  w  solid i , j


   , (D.1.4c)
 x2 i , j  w x2 i , j  w,  x2 i , j

and

 solid  solid i , j  solid i , j w


   . (D.1.4d)
 x2 i , j  w x1 i , j  x1 i , j w
230
D.2 Gauss-Seidel Iteration

The Gauss-Seidel iteration scheme is used to solve a matrix equation of the form:

Ax  b , (D.2.1)

where A is a square matrix with n rows and n columns, x is an unknown column vector

with n rows, and b is the solution vector with n rows. An initial guess vector xk is

provided. The next iterated vector xk+1 is given by 74:

 i 1
k
  Ai , j x j   .
n
xi k 1  1
 i  i, j j
b  A x
k 1
  (D.2.2)
Ai ,i  j 1 j  i 1 

This equation is iterated until a convergence criterion is met, such as the relative

difference of each xi is less than some value. A sufficient condition to guarantee

convergence is that:

n
Ai ,i   Ai , j , (D.2.3)
j 1
j i

and for all i, and at least one i the left hand side is greater than the right hand side.

231

You might also like