You are on page 1of 245

Supersonic Injection and Mixing

in the Shock Wave Reactor

Robert G. Cerff

A thesis submitted in partial fulfillment of


the requirements for the degree of

Master of Science in Aeronautics and Astronautics

The University of Washington

2010

Program Authorized to Offer Degree: Aeronautics and Astronautics

The University of Washington


Graduate School

This is to certify that I have examined this copy of a masters thesis by

Robert G. Cerff

and have found that it is complete and satisfactory in all respects,


and that any and all revisions required by the final
examining committee have been made.

Committee Members:

Arthur T. Mattick
Carl Knowlen

Date:

In presenting this thesis in partial fulfillment of the requirements for a masters degree at
the University of Washington, I agree that the Library shall make its copies freely available
for inspection. I further agree that extensive copying of this thesis is allowable only for
scholarly purposes, consistent with fair use as prescribed in the U.S. Copyright Law. Any
other reproduction for any purpose or by any means shall not be allowed without my written
permission.

Signature

Date

The University of Washington


Abstract

Supersonic Injection and Mixing


in the Shock Wave Reactor
Robert G. Cerff

The shock wave reactor is a laboratory scale facility at the University of Washington for the investigation of the fundamental technical issues related to a new chemical
processing technique. It applies an innovative gas dynamic approach to pyrolysis of
hydrocarbons in olefin manufacture. The applicability and effectiveness of this concept
for high-volume production of ethylene is of interest to the petrochemical industry as a
substitute for traditional hydrocarbon cracking processes. Based on technologies developed in a prior C2 H6 feedstock shock wave reactor experiment, this work presents the
design and preliminary operation of a CH4 feedstock reactor system at laboratory scale
with commercial scale up capability. Combustion of H2 and O2 augment superheated
steam from a pebble bed heater thus generating temperatures of 2300 K for optimal
CH4 pyrolysis. Thin walled insulating liners heat up to 1500K during operation to
minimize flow heat loss. A novel fluid dynamical design has been evaluated using computational fluid dyanamics software ANSYS Fluent demonstrating effective supersonic
mixing of steam with CH4 , minimizing pressure and heat losses. Injection schemes
comprise an optimized axisymmetric supersonic nozzle and oblique CH4 injection ports
in a staggered configuration. A so called aeroramp enhances injectant penetration
and mixing. Preliminary testing of the system is currently underway.

TABLE OF CONTENTS

Page
List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Chapter 1:

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Chapter 2:
Background . . . . . . . . . . . . . . . . . . . . . . . .
2.1 Economic Importance of Olefins . . . . . . . . . . . . . . . . .
2.2 Industrial Production of Olefins . . . . . . . . . . . . . . . . .
2.2.1 Traditional Steam Cracker Operation . . . . . . . . .
2.2.2 Olefin Process Variables and Thermochemistry . . . .
2.2.3 Steam/Hydrocarbon Ratio . . . . . . . . . . . . . . .
2.2.4 Process Temperature . . . . . . . . . . . . . . . . . . .
2.2.5 Process Residence Time . . . . . . . . . . . . . . . . .
2.2.6 Millisecond Furnaces - Short Residence Time Cracking
2.2.7 Coking Considerations . . . . . . . . . . . . . . . . . .
2.2.8 Olefin Feedstock Selection . . . . . . . . . . . . . . . .
2.3 Scramjet Engines . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

. 3
. 3
. 4
. 4
. 6
. 6
. 6
. 8
. 9
. 9
. 9
. 10

Chapter 3:
The Shock Wave Reactor . . . . . . . . . . . . . . . . .
3.1 Short High-Temperature Pulse Concept . . . . . . . . . . . .
3.2 The First Shock Wave Reactor Project - Ethane Feedstock . .
3.2.1 Operating Conditions . . . . . . . . . . . . . . . . . . .
3.2.2 Fluid Mechanics Studies . . . . . . . . . . . . . . . . . .
3.2.3 Pyrolysis Modeling . . . . . . . . . . . . . . . . . . . . .
3.2.4 Hot Flow Facility . . . . . . . . . . . . . . . . . . . . . .
3.2.5 Results and Future Studies . . . . . . . . . . . . . . . .
3.3 Alternative Feedstocks: Conversion of Methane . . . . . . . . .
3.4 The Second Shock Wave Reactor Project - Methane Feedstock

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

13
13
14
14
16
17
19
20
23
24

3.4.1
3.4.2
3.4.3

Commercial Scale Methane Reactor . . . . . . . . . . . . . . . . . . . 24


Laboratory Scale Methane Reactor . . . . . . . . . . . . . . . . . . . . 27
Reactor Operation at Design Conditions . . . . . . . . . . . . . . . . . 27

Chapter 4:
Design and Layout of the Laboratory Scale Reactor
4.1 Thermal Liner Insulation and Run Duration . . . . . . . .
4.2 Combustor and H2 /O2 Injectors . . . . . . . . . . . . . .
4.3 Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.4 Hardware and Data Acquisition . . . . . . . . . . . . . . .
4.4.1 Signal Conditioning . . . . . . . . . . . . . . . . .
4.4.2 LabVIEW Software Development . . . . . . . . . .
4.4.3 Software Data Acquisition . . . . . . . . . . . . . .
4.4.4 Starting a Test Run . . . . . . . . . . . . . . . . .
4.4.5 Post-Processing . . . . . . . . . . . . . . . . . . . .
4.4.6 Sensor Calibration . . . . . . . . . . . . . . . . . .
4.4.7 Quench Water Mass Flow Rate Calculation . . . .

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

32
32
35
42
47
47
48
51
51
53
53
54

Chapter 5:
Operation of the Laboratory Scale Reactor . . . . . . . . . . . . . . . . 56
5.1 Reactor Startup Sequence and Valve Timing . . . . . . . . . . . . . . . . . . 56
5.2 Lab Safety and Safety Shutdown Logic . . . . . . . . . . . . . . . . . . . . . . 58
Chapter 6:
Testing of the Laboratory Scale Reactor . . . . . . . . . . . . . . . . .
6.1 Dump Tank Butterfly Valve Calibration . . . . . . . . . . . . . . . . . . . . .
6.2 Spray Bar Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.3 Heat Losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.4 Combustion Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.4.1 Successful Combustion Augemented Steam - Steam Combustion Test .
6.4.2 Final Combustion Test Results . . . . . . . . . . . . . . . . . . . . . .
6.4.3 Titanium Oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Chapter 7:

59
60
60
64
65
66
72
76

Design Methodology and Constraints . . . . . . . . . . . . . . . . . . . 79

Chapter 8:
Supersonic Nozzle . . . . . .
8.1 Design Criteria and Approach . . .
8.2 Rocket Nozzles . . . . . . . . . . .
8.2.1 Parabolic Approximation of
8.2.2 Method of Characteristics .

.
.
.
a
.
ii

. . . . . . .
. . . . . . .
. . . . . . .
Bell Nozzle
. . . . . . .

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

81
81
83
84
88

8.3

Nozzle
8.3.1
8.3.2
8.3.3
8.3.4
8.3.5
8.3.6

Analysis . . . . . . . . . . . . . . . . . . . .
Quasi 1-D flow field model. . . . . . . . . .
Computational Fluid Dynamical Simulation
Turbulent Compressible Boundary Layer . .
Thermal analysis . . . . . . . . . . . . . . .
ANSYS Heat Transfer Validation . . . . . .
Structural Pressure Loading . . . . . . . . .

. . . . . .
. . . . . .
in Fluent
. . . . . .
. . . . . .
. . . . . .
. . . . . .

Chapter 9:
Methane Injector Design . . . . . . . . . . . . .
9.1 Scramjet injector technology . . . . . . . . . . . . . . .
9.1.1 Mixing in a Supersonic Flow . . . . . . . . . .
9.1.2 Possible Injector Configurations. . . . . . . . .
9.1.3 Existing Models of Transverse Injection . . . .
9.2 Size and Number of Injector Ports . . . . . . . . . . .
9.3 Injector Penetration Height Enhancement . . . . . . .
9.3.1 Fin analysis . . . . . . . . . . . . . . . . . . . .
9.3.2 Dual Transverse Injection : Aeroramp . . . . .
9.3.3 Estimating Lateral Spread and Coverage Area

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

90
90
92
101
106
116
117

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

119
120
121
122
126
136
139
141
143
146

Chapter 10:
Injector Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
10.0.4 Injector Placement and Injection Angle . . . . . . . . . . . . . . . . . 150
10.0.5 Further Optimizations . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
Chapter 11:
Final CFD Simulation Results .
11.1 Computational Domain . . . . . . . .
11.2 Meshing and Grid Resolution . . . . .
11.3 Solver Sets and Boundary Conditions .
11.4 Results and Conclusions . . . . . . . .
Chapter 12:

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

161
161
161
165
167

Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Appendix A:

LabVIEW Block Diagrams . . . . . . . . . . . . . . . . . . . . . . . . . 179

Appendix B: MATLAB Code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183


B.1 post processor gui.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183

iii

B.2
B.3
B.4
B.5

get Line Info.m


calibrate.m . .
distPlot.m . . .
transientPlot.m

Appendix C:

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

190
191
194
195

Spray Bar Mass Flow Rate - Noise Reduction Techniques . . . . . . . 197

Appendix D:

MATLAB code for pseudo 1-D


Nozzle Heat Transfer . . . . . .
D.1 main.m . . . . . . . . . . . . . . . . .
D.2 calcCombustorPenetration.m . . . . .
D.3 calcNozzleFlowProperties.m . . . . . .
D.4 getArea.m . . . . . . . . . . . . . . . .
D.5 getCH4Viscosity.m . . . . . . . . . . .
D.6 getGasProperties.m . . . . . . . . . . .
D.7 getMach.m . . . . . . . . . . . . . . .
D.8 getNozzleMaterialConstants.m . . . . .
D.9 nozzleHeatTransfer.m . . . . . . . . .
D.10 plotCrossSectionalMixing.m . . . . . .

Appendix E:

.
.
.
.

model, Basic
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .

Feedstock
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .

Mixing and
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .

.
.
.
.
.
.
.
.
.
.
.

206
207
214
215
216
219
220
221
221
222
226

Nozzle and Injector Fabrication . . . . . . . . . . . . . . . . . . . . . . 227

iv

LIST OF FIGURES

Figure Number

Page

2.1
2.2
2.3

Traditional Cracking Reactor Layout. Ref.[42] . . .


An Olefin Francis Diagram Ref.[42] . . . . . . . . .
X-43, unmanned experimental hypersonic aircraft,
program Ref.[47] . . . . . . . . . . . . . . . . . . .

3.1
3.2
3.3

Temperature Recovery Due To A Shock Wave. Ref.(Ref.[13]) . . . . . . . . .


Schematic Of Shock Wave Reactor and Gas Temperature Profiles. (Ref.[13]) .
Ethylene Selectivity vs. Ethane Conversion For Isothermal, Isobaric Pyrolysis
At 1075 K And 1200 K , (Ref.[13]) . . . . . . . . . . . . . . . . . . . . . . . .
Schematic of hot-flow shock wave reactor facility. . . . . . . . . . . . . . . . .
Point Design Of A Shock Wave Reactor For Ethylene Production . . . . . . .
Measured and Predicted Yields of C2 H6 , C2 H4 , C2 H2 , and CH4 From the
Shock Wave Reactor. (Ref.[13]) . . . . . . . . . . . . . . . . . . . . . . . . . .
Measured Ethylene Yield vs. Ethane Conversion For The Shock Wave Reactor And For Conventional Reactors. Ref.(Ref.[13]) . . . . . . . . . . . . . . .
Methane Processing Design Example Ref.(Ref.[13]) . . . . . . . . . . . . . . .
Methane Processing Flow Diagram Ref.(Ref.[13]) . . . . . . . . . . . . . . . .
Current Shock Wave Reactor Layout Ref.(Ref.[13]) . . . . . . . . . . . . . . .
Combustor Section, Feedstock Injection Section and Mixing Sections . . . . .
Mixer Section, Transition Section, Reactor Section and Diagnostic Section . .
Water Quench Spray Bar System . . . . . . . . . . . . . . . . . . . . . . . . .
Laboratory Scale Shock Wave Reactor Flow Schematic . . . . . . . . . . . . .

3.4
3.5
3.6
3.7
3.8
3.9
3.10
3.11
3.12
3.13
3.14
4.1
4.2
4.3
4.4
4.5
4.6
4.7

. . . . . . . . . . . . . . . 5
. . . . . . . . . . . . . . . 7
from NASAs Hyper-X
. . . . . . . . . . . . . . . 12

Wall Insulation System . . . . . . . . . . . . . . . . . . . . . . . . . .


Optional caption for list of figures . . . . . . . . . . . . . . . . . . .
Optional caption for list of figures . . . . . . . . . . . . . . . . . . .
H2 O2 Injector Cut-away . . . . . . . . . . . . . . . . . . . . . . . .
H2 O2 Injector Cross-Section and Combustion Zone Schematic . . .
H2 O2 Injector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Equilibrium Combustion Products From Adiabatic H2 :O2 Injectants
: 1 Molar Stoichiometry . . . . . . . . . . . . . . . . . . . . . . . . .
v

. .
. .
. .
. .
. .
. .
At
. .

. .
. .
. .
. .
. .
. .
2.5
. .

.
.
.
.
.
.

13
15
17
18
19
21
23
25
26
28
29
30
30
31
33
34
36
37
38
39

. 40

4.8
4.9
4.10
4.11
4.12
4.13
4.14
4.15
4.16

2300 K Equilibrium Composition Of Combustion Products And Pebble Bed


Heater Steam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Sensor Port Layout in the Shock Wave Reactor . . . . . . . . . . . . . . . .
Optical Temperature Sensor Schematic . . . . . . . . . . . . . . . . . . . . .
UV And Visible Radiation Intensity Variation With Temperature . . . . . .
Data-Acquisition Organizational Chart . . . . . . . . . . . . . . . . . . . . .
LabVIEW Front Panel for Valve Control . . . . . . . . . . . . . . . . . . . .
LabVIEW Front Panel for Data Display . . . . . . . . . . . . . . . . . . . .
Software Algorithm Flow Chart of LabVIEW VI . . . . . . . . . . . . . . .
Noise in the water mass vs. time plot with trendline . . . . . . . . . . . . .

5.1
5.2

Five Phases Occuring During A Run . . . . . . . . . . . . . . . . . . . . . . . 56


Safety Control System Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

6.1
6.2
6.3
6.4
6.5
6.6

Dump Tank Butterfly Valve Calibration (at t = 10 s) . . . . . . . . . . . .


Dump Tank Pressure vs.Time . . . . . . . . . . . . . . . . . . . . . . . . . .
Spatial Pressure Distribution Controlled by Butterfly Valve Setting . . . . .
Water Mass Flow Rate Calibration . . . . . . . . . . . . . . . . . . . . . . .
Temperature at Three Wall Distances in the Diagnostic Section . . . . . . .
Temperature Drop vs. Time Across the Length of the Reactor (with 1400 K
Steam Only) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Successful Combustion Test: Combustion Pressure . . . . . . . . . . . . . .
Successful Combustion Test: Mass Spectrometer Species Data . . . . . . . .
Hot Combustion Test 2: Flow Temperature Immediately Downstream of
Combustor Section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Hot Combustion Test 2: Reactor Exit Temperature . . . . . . . . . . . . . .
Peak Combustion Temperature After 1 sec of Combustion . . . . . . . . . .
Hot Combustion Test 2: Combustor Liner Temperature . . . . . . . . . . .
Hot Combustion Test 2: Temperature Drop Along the Centerline Across the
Reactor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Titanium Fire: Combustor Exit Pressure (C3T) vs. Time . . . . . . . . . .
Titanium Fire: Reactor Exit Temperature (R5T) vs. Time . . . . . . . . .
Titanium Fire: Combustor Liner vs. Time . . . . . . . . . . . . . . . . . . .
Titanium Fire: Combustion temperature (at constant pressure) for titanium
with oxygen in post-shock flow as a function of shock Mach number. . . . .

6.7
6.8
6.9
6.10
6.11
6.12
6.13
6.14
6.15
6.16
6.17

.
.
.
.
.
.
.
.
.

.
.
.
.
.

41
43
46
47
48
49
50
52
55

61
61
62
63
64

. 65
. 67
. 68
.
.
.
.

69
70
71
72

.
.
.
.

73
74
75
75

. 77

7.1

Cross-Section View: Feedstock Injection Section

8.1

Conical And Bell Nozzle Contour (Ref.[24]) . . . . . . . . . . . . . . . . . . . 85


vi

. . . . . . . . . . . . . . . . 80

8.2
8.3
8.4
8.5
8.6
8.7
8.8
8.9
8.10
8.11
8.12
8.13
8.14
8.15
8.16
8.17
8.18
8.19
8.20
8.21
8.22
9.1
9.2
9.3
9.4
9.5

9.6

Parabolic Approximation Of Bell Nozzle Contour (Ref.[24]) . . . . . . . . . . 86


n and e as Function of Expansion Area Ratio (Ref.[24]) . . . . . . . . . . . 87
Approximation of a Minimum Length Bell nozzle . . . . . . . . . . . . . . . . 88
Comparison Between Different Nozzle Contours. . . . . . . . . . . . . . . . . 89
Optimal Nozzle Profile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
Optimal Nozzle: Mach Number vs. Distance . . . . . . . . . . . . . . . . . . . 92
Optimal Nozzle: Pressure vs. Distance . . . . . . . . . . . . . . . . . . . . . . 93
Optimal Nozzle: Temperature vs. Distance . . . . . . . . . . . . . . . . . . . 94
Optimal Nozzle: Density vs. Distance . . . . . . . . . . . . . . . . . . . . . . 95
Optimal Nozzle: Momentum Flux vs. Distance . . . . . . . . . . . . . . . . . 96
Comparison of Mach Profile In A Bell Model (top) and in the Optimized
Profile (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
Comparison Of Total Pressure Profile In A Bell Model (top) and in the
Optimized Profile (bottom). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Laminar and Turbulent Boundary Layer Growth in Flat Plates Ref.[42] . . . 105
Calculated Turbulent Boundary Layer Development in the Supersonic Nozzle. 107
CFD Results of Turbulent Boundary Layer Development in the Supersonic
Nozzle. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
Curvilinear Computational Grid for Simulating Nozzle Heat Transfer . . . . . 111
Temperature Distribution In Nozzle Walls. . . . . . . . . . . . . . . . . . . . . 113
Variation Of Temperature Distribution In Nozzle Walls During Experiment
Time. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Finite Element Tetrahedral Element Mesh. . . . . . . . . . . . . . . . . . . . 116
Variation of Temperature in Nozzle Walls from ANSYS. . . . . . . . . . . . . 117
Forces Acting on the Internal Nozzle Walls due to Steam Flow Pressures . . . 118
A Selection of Typical Injecting Techniques for Scramjets Ref.[42] . . . . . .
Sketch of a Flowfield in the Vicinity of A Transversely Injected and Underexpanded Jet. Ref.[40] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Spaids model of the effective shape of the injectant causing body interaction
in the free stream. Ref.[44] . . . . . . . . . . . . . . . . . . . . . . . . . . .
Billings New Model Of The Effective Shape Of The Injectant Causing Body
Interaction In The Free Stream. Ref.[10] . . . . . . . . . . . . . . . . . . . .
Comparison of calculated jet cross-sectional areas and measured jet concentration contours (on the left). Comparison of calculated values of the maximum ordinate of the Mach disk (on the right). (Ref.[10]) . . . . . . . . . . .
Flow Model Of Transverse Jet Interaction. (Ref.[40]) . . . . . . . . . . . . .

vii

. 123
. 124
. 128
. 130

. 131
. 132

9.7
9.8
9.9
9.10
9.11
9.12

9.13
9.14
9.15
9.16
9.17

Model Errors in Penetration versus Momentum Flux Ratio J. . . . . . . . . . 134


Jet penetration Development According To The 7 Models Described. . . . . . 136
Jet Penetration For 4 Injector Configuration. . . . . . . . . . . . . . . . . . . 138
Comparison between instantaneous images of the case of no fin (top figure)
and fin-guided fuel injection at various axial downstream distances. (Ref.[33]) 140
Fin Geometry Shown With Injection Port And Defined Axis System. . . . . . 140
Transverse Section Through Shock Wave Reactor Duct Showing Methane
Distribution At 12 Injector Port Diameters Downstream Of Injection. (approximately to scale). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
Thermal analysis for a fin made in titanium. After 15 sec it has exceeded its
melting temperature (1941 K). . . . . . . . . . . . . . . . . . . . . . . . . . . 143
Schematic View Of Mean Flow Field Of Dual Transverse Injection. Ref.(Ref.[28])144
Comparison of Penetration Height for Single And Dual Transverse Injection
Systems (Left) And Showing Mach Disc Position (Right). Ref.(Ref.[28]) . . . 145
Comparison Of Penetration Distances (Left) And Of Stagnation Pressure
Normalized By Stagnation Pressure Of The Cross Flow. Ref.(Ref.[28]) . . . . 146
Lateral Spread Obtained from Numerical Study (Ref [28]) . . . . . . . . . . 147

10.1
10.2
10.3
10.4

Mach contour for single transverse injection (top) and dual injection (bottom) .151
Comparison of mass fraction of methane for different jet angles. . . . . . . . . 152
Mach contour comparison between transverse and oblique secondary jet. . . . 154
Mach contour comparison between different distance from front and rear jet
in the aeroramp configuration. . . . . . . . . . . . . . . . . . . . . . . . . . . 155
10.5 Stagnation pressure comparison between different distance from front and
rear jet in the aeroramp configuration. . . . . . . . . . . . . . . . . . . . . . . 156
10.6 Scheme of the transverse injection 6 downstream the aeroramp injection point.158
10.7 Mach contour comparison between different distances from primary to secondary jets: 5 (top) vs 4 (bottom. . . . . . . . . . . . . . . . . . . . . . . . 159
11.1
11.2
11.3
11.4
11.5
11.6
11.7
11.8
11.9

Final configuration of nozzle, injector and mixing section. . . . . . . .


Mesh Refinement Zones using Spheres of Influence. . . . . . . . . . .
Injector Inlet Mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Boundary Layer Prismatic Mesh . . . . . . . . . . . . . . . . . . . . .
Mach Contour Lines. . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Sequence of Lambda Shocks And Section Geometry. . . . . . . . . . .
Static Temperature Contours at the Beginning of the Mixing Section.
Total Temperature Contours In Injector And Mixing Section. . . . . .
Stagnation pressure losses along nozzle-injector and mixing sections. .
viii

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

161
163
163
164
168
169
169
170
171

11.10Methane Mass Fraction Along The Duct And At The Exit Plane. . . . . . . . 172
11.11Jet Plume Development Over A Length Of 8 Downstream The First Injection Point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
11.12Jet Plume Development 8 Downstream The First Injection. . . . . . . . . . 173
A.1
A.2
A.3
A.4
A.5

Block
Block
Block
Block
Block

Diagram
Diagram
Diagram
Diagram
Diagram

1
2
3
4
5

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

179
179
180
180
181

C.1
C.2
C.3
C.4
C.5
C.6
C.7
C.8
C.9

Noise in the water mass vs. time plot . . . . . . . .


Noise in the water mass vs. time plot with trendline
Graphically determined mass flow rate . . . . . . . .
Mass flow rate vs. time before noise reduction . . . .
Mass flow rate vs. time: Lower Order Fits . . . . . .
Mass flow rate vs. time: High Order Fits . . . . . .
Mass flow rate vs. time: rlowess Fits . . . . . . . .
Water Tank Mass vs. time: Sliced rlowess Fits . . .
Water Tank Mass vs. time: Sliced rlowess Fits . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

197
198
199
200
201
201
203
204
205

E.1
E.2
E.3
E.4

CAD Model Of Injection Section. . . . . . . . . . . . . . . . . . . . . . . . .


Exploded Cutaway Of Injection Section CAD Model . . . . . . . . . . . . .
Cutaway of CAD Model Showing Injetor Ports . . . . . . . . . . . . . . . .
Photographs of fabricated parts: outside of the nozzle (left), inside nozzle
profile (mid), and nozzle with injection section showing aeroramp injector
ports (right) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

ix

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

. 227
. 228
. 228

. 229

LIST OF TABLES

Table Number

Page

2.1

Stability Series of Alkenes vs. Temperature Ref.[26]

. . . . . . . . . . . . . .

3.1

Chemical Yields vs. Station For The Shock Wave Reactor. Ref.(Ref.[13]) . . 22

4.1

A typical set of sensors in the reactor . . . . . . . . . . . . . . . . . . . . . . . 45

9.1

Injection Conditions in the Shock Wave Reactor . . . . . . . . . . . . . . . . 135

10.1 Comparison a 4 inch and 20 inch Long Conical Expander . . . . . . . . . . . 160


11.1 Total Temperature Losses Along Nozzle-Injector And Mixing Sections . . . . 171
A.1 Flow Valve Opening Sequence . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
C.1 Time domain slicing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 202

ACKNOWLEDGMENTS

The author wishes to express sincere appreciation to Professor Tom Mattick and Dr. Carl
Knowlen of the University of Washington Department of Aeronautics and Astronautics for
their help and encouragment, and their many contributions to the research which is the
subject of this thesis.
A great debt of gratitude is extended to Signorina Sofia Sartori for her equal part in the work
presented here. Her late nights and early mornings are appreciated. For her outstanding
fortitude, attention to detail and exhuberance the author is thankful and honoured! Grazie!
Thanks are given to colleages: Peter Gangar, Viggo Hansen, Alex Lacomb, Pedro O.
Ramirez, Alex Perez, Carlos Rufin and Isaac Statnekov at the Shock Wave Reactor Laboratory for their ingenuity and assistance.
The research work described in this thesis is supported by Universal Oil Products LLC, of
Honeywell, and thanks are extended to their staff in Chicago.

xi

DEDICATION

to my Father, who paved the way.

xii

Chapter 1
INTRODUCTION

The shock wave reactor is a laboratory scale facility at the University of Washington for
the investigation of the fundamental technical issues related to a new chemical processing
technique. It applies an innovative gas dynamic approach to pyrolysis of hydrocarbons in
olefin manufacture. The applicability and effectiveness of this concept for high-volume production of ethylene is of interest to the petrochemical industry as a substitute for traditional
hydrocarbon cracking processes.
Based on the potential of the shock wave reactor for improving yields in olefin manufacture, the first shock wave reactor research program was undertaken in 1990 to examine the
gasdynamic and chemical processes fundamental to the reactors operation. The goal was
to obtain the data needed to enable industry to make an assessment of this approach for
commercial application, as well as to make contributions to the scientific understanding of
these processes. The first project saw the construction of a shock wave reactor that realized
a 20% improvement in ethylene yield over that achieved in current industry practice. The
facility used ethane as a feedstock gas at pyrolysis temperatures of 1200 K. Though experimental studies allowed the fluid dynamics in the shock wave reactor to be well understood,
heat losses were prohibitive, high steam dilutions were necessary and industrial scale up
and economics required further investigation. [27].
Building on the techniques of the first project, the current second project in the program
is also exploratory and addresses the proposed use of methane as a feedstock at necessarily
higher pyrolysis temperatures to satisfy chemical kinetics requirements and reduce steam
dilutions. The wide availability of methane from natural gas at high pressure has considerable economic potential, and modeling indicates that shock wave reactors with methane
feedstock offer improvement in yields and conversion over conventional methane-to-ethylene

conversion techniques. The goals of the second project are to show enhanced yields of ethylene from methane on an economic and performance basis whilst overcoming technical design
challenges involved, particularly with respect to thermal management and fluid dynamics.
Careful fluid dynamic design in the shock wave reactor becomes necessary when effective
mixing of steam and methane is desired because pyrolysis chemistry can be adversely affected
by both non-homogeneous thermal and chemical mixing. In addition, ethylene yields have
long been known to be affected by control of pyrolysis residence time, and this, in turn is
controlled directly by the fluid dynamics in the shock wave reactor.
This thesis presents the second shock wave reaction project to date and the work of the
author and visting scholar to the University of Washington, Signorina S. Sartori of the
Politecnico di Torino, Turin, Italy. It presents the design and construction of the supersonic
mixing chamber for the effective injection and mixing of methane into supersonic steam cross
flow. Co-presentation of this work is given by Sartori 2010 [42] Goals for the chamber are the
precise control of pyrolysis temperature profiles to achieve enhanced yields of ethylene whilst
reducing heat loss from chamber walls and maintaining minimal injection pressure losses.
At the time of writing, construction of the chamber is complete but not experimentally
tested at full capacity.

Chapter 2
BACKGROUND
2.1

Economic Importance of Olefins

Among the most economically important and energy intensive enterprises in the petrochemical industry is the manufacture of olefins (alkenes).[27]. Olefins naturally occur in refinery
gasses but only to a minor extent. Their use in industry is so widespread today that they
are manufactured deliberately from raw materials such as crude oil fractions or natural gas.
The most useful member of the alkene series is ethylene [26], although polypropylene also
has major uses. World production of ethylene is around 110 million tons per annum, whilst
that of propylene is 65 million tons. Other series of organics hydrocarbons manufactured
from natural gas are aromatics (70 million tons per annum) and synthesis gas, making
ethylene the worlds largest petrochemical commodity.
The term petrochemicals refers to chemicals that are derived from crude oil and manufactured in large scale production processes. Ethylene must be produced in large volumes
because it is a building block in the manufacture of a wide variety everyday products and
industrial chemicals. The applications of ethylene are numerous and ethylene derivatives are
traded around the world. Ethylene is a petrochemical derived monomer whose predominant
use has been as an intermediary for polyethylene, a thermoplastic of wide utility. More generally, ethylene is used primarily as a feedstock in the production of plastics, synthetic fibers
and other organic chemicals that are ultimately consumed in the packaging, transportation
and construction industries and in a multitude of industrial and consumer markets (packaging uses make up more than half (55%) of ethylene derivative consumption worldwide).
Ethylene is used for different purposes: 80% is employed to create ethylene oxide, ethylene
dichloride, and polyethylene; in smaller quantities it is used as an anesthetic agent, to hasten fruit ripening, as well as a welding gas. Polyethylenes of various densities and melt flow
indices account for more than 50% of world ethylene demand. The primary use of polyethy-

lene is in film applications for packaging, carrier bags and trash liners. Other applications
include injection molding, pipe extrusion, wire and cable sheathing and insulation, as well
as extrusion coating of paper and cardboard.
2.2

Industrial Production of Olefins

Up until the later half of the 20th century, ethylene had been produced by hydrogenation
of acetylene or dehydrogenation of ethanol over either alumina or phosphoric acidcoke catalysts [26]. Though these older processes are in use today in parts of Africa, Asia and
South America, industrialized countries make use of steam to dehydrogenate hydrocarbon
feedstocks in a process of thermal dehydrogenation called cracking.
Although most petrochemical manufacturing processes employ catalysts where reactions are
carried out selectively at modest temperatures, the highly endothermic and refractory nature
of the reactions of saturated hydrocarbons to produce olefins have not so far been amenable
to this approach; the current production method uses steam cracking of hydrocarbons in
furnace tubes. This approach, developed with continual, incremental improvements over the
last 50 years, simply exposes a petrochemical feedstock (ranging from light alkanes such as
ethane or propane to liquids such as naphtha), mixed with steam, to temperatures ranging
up to 1200 K as the mixture passes through tubes in a furnace.[41]
2.2.1

Traditional Steam Cracker Operation

Typical arrangement of such a cracking reactor is shown in Figure 2.1. A hydrocarbon


stream is heated by heat exchange against gas flow in the convection section, mixed with
steam, and further heated to incipient cracking temperature 500680 C depending on the
feedstock. The stream then enters a fired tubular reactor (radiant tube or radiant coil)
where, under controlled residence time, temperature profile, and partial pressure, is heated
from 500650 C to 750875 C for 0.10.5s. During this short reaction time hydrocarbons
that are in the feedstock are cracked into smaller molecules: ethylene, other olefins, and
diolefins are the major products. Since conversion of saturated hydrocarbons to olefins in a
radiant tube is highly endothermic, high energy input rates are needed. Reaction products
leaving the radiant tube at 800850 C are cooled to 550660 C within 0.020.1 s to prevent

Figure 2.1: Traditional Cracking Reactor Layout. Ref.[42]

degradation of highly reactive products by secondary reactions. Resulting product mixtures,


which can vary widely, depending on feedstock and severity of the cracking operation,
are then separated into desired products by using a complex sequence of separation and
chemical-treatment steps. The cooling of the cracked gas in the transfer line exchanger
is carried out by vaporization of high-pressure boiler feed water (612 MPa), which is
separated in the steam drum and subsequently superheated in the convection section to
high-pressure superheated steam (612 MPa).

2.2.2

Olefin Process Variables and Thermochemistry

The most important process variables in these steam crackers are steam/hydrocarbon ratio,
reactor temperature and residence time. Feedstock characteristics are also considered, since
they influence the process severity. Shock wave reactors offer benefits over traditional steam
crackers by virtue of their higher temperature operation, method of reaction enthalpy supply
and their shortened and precisely controlled feedstock residence time. To understand why
this is the case, olefin pyrolysis mechanisms must first be understood.
2.2.3

Steam/Hydrocarbon Ratio

Superheated steam is included as a diluent in the traditional ethylene manufacturing process


for two key reasons. Since olefin formation is favored for lower feedstock partial pressures,
the first of these reasons is to reduce the partial pressure of ethane feedstock. The second
is that superheated steam reduces carbon deposits that are formed by the pyrolysis of
hydrocarbons at high temperatures.
2.2.4

Process Temperature

Optimal cracking processes hinge on the control of alkane and alkene thermochemistry. This
is due to the fact that steam cracking reactions are highly endothermic. Increasing temperature favors the formation of olefins, high molecular weight olefins and aromatics. Optimum
temperatures are selected to maximize olefin production and minimize the formation of
carbon (coke/soot) and its deposition (coking).
For example, Ethylene is chemically much more reactive than is ethane at room temperature,
and has much greater free energy of formation. The free energy of formation is, however,
temperature-dependent as is shown on the Francis diagram in Figure 2.2.
. Thus, above about 1125 K, i.e. about 850 C, ethylene is more stable than is ethane
and can be prepared from it by cracking. Moreover, as the temperature is raised further,
acetylene is also formed: acetylene has a lower free energy of formation than ethane above
about 1250 K (975 C deg C) and than ethylene above about 1400 K (1225 C deg C). The
stability series in Table 2.1 thus reads:

Figure 2.2: An Olefin Francis Diagram Ref.[42]

Temperature Range

Stable C2 Hydrocarbon

Below 850 C

Ethane

850 to 1225 C

Ethylene

Above 1225 C

Acetylene

Table 2.1: Stability Series of Alkenes vs. Temperature Ref.[26]

At any one temperature in the 1000 C region, significant amounts of all three hydrocarbons
will occur. Over 820 K, all hydrocarbons are less stable than their components, carbon and
hydrogen, and these comprise the mixture equilibrium which is reached rapidly. Obtaining
ethylene from ethane thus requires temperatures exceeding 850 C. Propylene and the higher
alkenes are favored over the corresponding alkanes at progressively lower temperatures.
Thermal cracking therefore provides easy access to the alkenes from alkanes, which in turn
are readily obtained from the refining of crude oil or from natural gas rich in ethane and
propane (particularly in the U.S. Gulf coast).
2.2.5

Process Residence Time

Many secondary reactions occur in parallel with these primary alkane-to-olefin reactions
and reaction kinetics are highly temperature and time dependent. What mixture is actually
isolated after cracking temperature is reached thus depends on the residence time. Very
rapid chilling (quenching) will freeze the chemistry at a stage where the fastest kinetics
are underway. This is used when acetylene is the desired product. On slower quenching,
however, the kinetics take over and the acetylene combines with the hydrogen; a mixture
of ethylene and ethane results containing relatively little acetylene. [26]. Lack of quenching
results in other deleterious hydrocarbons being produced as well as carbon and hydrogen
formation. Residence time is also a compromise between reaction temperature and other
variables.
In general, when ethane and light hydrocarbon gasses are used a feeds, shorter residence
times are used to maximize olefin production and minimize benzene, toluene, xylene (BTX)
and liquid yields. Residence times of 0.5-1.2 s are typical. [31] There is a strong economic
incentive for considering short residence time cracking processes. Not only do such processes
increase the yields of the more valuable liquid and gaseous products, but more compact
designs would also decrease capital costs. Careful control of the vapor residence times
appears to be crucial in order to prevent secondary cracking and yet allow for maximum
cracking of the feedstock. Rapid and thorough mixing of the feedstock with the heat source,
not just creating a uniform dispersion, is also a key design aspect to consider.[23]

2.2.6

Millisecond Furnaces - Short Residence Time Cracking

A fairly new development in cracking liquid feeds that improves ethylene yield is the millisecond furnace. Typical Kellogg millisecond furnaces operates with a residence time
between 0.030.1 seconds.[31] The millisecond furnace probably represents the last step
that can be taken with respect to this critical variable because contact times below 0.01
seconds range lead to production of acetylenes in large quantities.[31]Shock wave reactor
residence times are on the same order as millisecond furnaces.
2.2.7

Coking Considerations

Due mainly to the high temperatures, coke (essentially solid carbon) is always formed in
hydrocarbon pyrolysis; it is a highly undesirable by-product. The coke formation and the
collection on the coil surfaces increase the heat-transfer resistances and the furnace has
to be shutdown at intervals for decoking. Both complete shutdown of the furnace, or the
so-called on-line decoking, reduce the ethylene production, and increase production cost.
[49] In the steam cracking of hydrocarbons, a small portion of the hydrocarbon feed gases
decomposes to produce coke that accumulates on the interior walls of the coils in the radiant
zone and on the inner surfaces of the transferline exchanger (TLX). Albright et al.[4, 48, 5]
identified three mechanisms for coke formation. Mechanism 1 involves metal-catalyzed
reactions in which metal carbides are intermediate compounds and for which iron and nickel
are catalysts. Mechanism 2 results in the formation of tar droplets in the gas phase, often
from aromatics. These aromatics are often produced by trimerization and other reactions
involving acetylene. Mechanism 3 first involves the reactions of gaseous microspecies with
the free radicals on the coke surfaces. These microspecies include acetylene, ethylene or
other olefins, butadienes, and free radicals such as methyl, ethyl, or benzyl radicals; second,
C-H bonds located on the surface break to produce coke. [49]
2.2.8

Olefin Feedstock Selection

The choices of feedstocks for the production of ethylene and other olefins are of major importance. Modern ethylene units are often designed so that different hydrocarbon feedstocks

10

can be used. A modern plant often costs between US 500 million and 2 billion depending on the production capacity, the feedstocks, and the products obtained. Countries with
considerable natural gas ethane and propane prefer using them as feedstocks; such areas
include the United States, Canada, and the Middle Eastern countries. Europe and Japan,
however, have and prefer liquid feedstocks (naphthas and gas oils). Ethylene operators
typically employ sophisticated computer models and optimization techniques in choosing
feedstocks.[49]
Use of methane as a feedstock is less common than heavier hydrocarbons. Direct conversion
of methane to other useful products is one of the most challenging subjects to be studied
in heterogeneous catalysis. Methane activation is difficult because of its thermodynamical
stability with a noble gas-like electric configuration. The strong tetrahedral C-H bonds
(435 kJ/mol) make it less reactive than ethane and propane and nearly all its conversion
products.[43]
2.3

Scramjet Engines

An analogy exists between the shock wave reactor and the design of scramjets in the
aerospace industry. The way in which the shock wave reactor mixes steam and feedstock
holds many similarities with the way scramjet combustion chambers mix air and fuel. The
analogy is more clearly understood when the background and operation of a scramjet are
more familiar to the reader.
Air-breathing engines are conventionally used on civil and military air aircraft: they use
a compressor to squeeze air into the engine, then spray fuel into the compressed air
and ignite it to produce thrust by funnelling it through the back. Most of these engine
are turbojets whose performance remains limited to a range of Mach numbers around 34. Ramjets are distinct from this in that they do not require a compressor and uses the
speed of the aircraft to compress the air, so very few moving parts are needed to operate
it. In particular there is no high-speed turbine, as in a turbofan or turbojet engine, that
is expensive to produce and can be a major point of failure. In ramjets, combustion takes
place at subsonic velocities.
A scramjet (supersonic combustion ramjet) is an air-breathing propulsion engine designed

11

for hypersonic flight (Mach numbers 5) and it is a variation of a ramjet distinguished


by supersonic combustion. To leave Earths atmosphere, a rocket has to reach and escape
velocity and thus a Mach regime of 20 to 25. For satellites and space shuttle launches, thrust
is provided by large rockets that are very heavy, non-reusable, have low manoeuvrability
and a vertical take off is required. Turbojets cannot be used for this purposes because
they are limited by the turbine blade velocity.In addition, as aircraft velocity increases, air
stagnation temperature rises and so there is a risk of melting in the engine.
To allow flight at high Mach numbers while avoiding problems associated with compressorturbine velocity and thermal loads, engineers have developed a new engine: the ramjet.
Ramjets essentially consists of a constricted tube through which inlet air is compressed
by the high speed of the vehicle, a combustion chamber where fuel is combusted, and a
nozzle through which the exhaust jet leaves at higher speed than the inlet air. Ramjets are
capable of flight up to Mach 6 but at these speeds, when the intake air is slowed to subsonic
velocities to enter in the combustion chamber, temperatures rise to around 3500 K leading
to chemical dissociation. Thus, when the combustion of H2 fuel starts, instead of producing
water (that will be followed by substantial pressure rise and thrust) the combustion reaction
produces free radicals at lower pressure and with lower thrust. To avoid this loss in thrust
efficiency, the intakes of ramjets have been modified to slow down the air flow at the inlet
less than before in order to have a supersonic, and not subsonic, combustion. This is the
principle of scramjet operation.
The history of these engines dates back to World War II, when a lot of effort was put in
research towards high speed jets and rocket aircraft as military weapons. In particularly the
U.S. Navys needed an anti air weapon that could defeat aircraft threats using small, highspeed, radar-guided anti-surface missiles. This request led to rocket- and ramjet-powered
flight vehicles.
When the war was over this new technology had been adopted by civil industries. However, the main goal of civilian air transport has been reducing operating cost, rather than
increasing flight speeds. Because supersonic flight using conventional jet engines requires
significant amounts of fuel, airlines have favoured subsonic jumbo jets rather than supersonic
transports.

12

In the United States, from 1986-1993, a reasonably serious attempt to develop a single
stage to orbit reusable space plane using scramjet engines was made, but the Rockwell X30 (NASP) program was discontinued before it yielded any working hardware.Hypersonic
flight concepts have not disappeared, however, and low-level investigations have continued
over the past few decades. Intitially, the attention was focused on ramjets though scramjet
offered several improvements like :
- More efficient operation at off-design conditions;
- Less dissociation of combustion products and lower static pressure, which results in
less heat transfer as well as structural load. This can reduce aircraft weight. The key
is that supersonic combustion occurs at lower static temperature in the combustion
chamber .
Nowadays scramjet technology is seen as one with high potential, there are many applications that could be developed. Currently space launches are considered a viable application,
though the topic is still open with advantages and disadvantages.The scramjet is seen as
an attractive alternative to rocket propulsion; where a rocket carries its oxygen and fuel
on board, the scramjet captures its oxygen from the atmosphere thus making the vehicle
lighter. Potential applications for scramjet include satellite launch vehicles and hypersonic
cruise missiles. Few scramjet test beds exist, though the most famous is likely the NASA
Hyper-X programs X-43 experimental aircraft seen in Figure 2.3

Figure 2.3: X-43, unmanned experimental hypersonic aircraft, from NASAs Hyper-X program Ref.[47]

13

Chapter 3
THE SHOCK WAVE REACTOR
3.1

Short High-Temperature Pulse Concept

It has long been known that the olefin yield from hydrocarbon pyrolysis increases with the
use of high processing temperatures and short reaction durations. Much of the technology
development in steam cracking over the past several decades has been aimed at achieving
such a temperature profile.[41, 36] However, the rate of the temperature rise of the feedstock
is established by heat transfer from tube walls, and is fundamentally limited by the properties of the tube materials. Despite impressive yield gains resulting from use of improved
alloys and optimal tube geometries, much improved yields could result from application of
a short high-temperature pulse to the feed as illustrated in Figure 3.1. In gasdynamics,
such a short timescale pulse is observed in the passage of a shock wave which capability to
instantaneously and volumetrically heat a gas.

htpb]
Figure 3.1: Temperature Recovery Due To A Shock Wave. Ref.(Ref.[13])

14

3.2

The First Shock Wave Reactor Project - Ethane Feedstock

The first shock wave reactor, conceived in 1990, is a fundamentally new approach to pyrolysis that employs a gasdynamic shock wave heating, rather than convection, to control the
temperature history of a reactant.[7, 1] This approach decouples the processes of heat addition and pyrolysis, and enables one to fine-tune the temperature history to reach optimum
yields of a desired product. Tube alloy properties are no longer a controlling factor in this
process. Development of the first shock wave reactor idea took place in 1992. The basic
idea was threefold.
1. A carrier gas superheated steam in this case is superheated to a high enthalpy
state and cooled by adiabatically acceleration to supersonic speeds (M = 2.0). Enough
enthalpy is supplied for the entire pyrolysis reaction to occur.
2. Hydrocarbon feedstocks are then mixed, at supersonic speeds, with cooler high-temperature
steam.
3. The mixed flow is suddenly decelerated and heated volumetrically in passage through
a shock wave to initiate pyrolysis. Reactions proceed adiabatically using the enthalpy
in the carrier flow.
Experimental design of the first shock wave reactor process was based on three significant
development efforts, namely a fluid mechanics study involving the construction of a cold flow
facility, a numerical pyrolysis modeling code and finally the construction of the first shock
wave reactor as a laboratory scale hot flow facility. The conditions of shock wave reactor
operation are presented first to introduce the process. Subsequently, the ideas behind the
fluid dynamical design and chemical kinetics modeling of this process are detailed. Finally,
relevant details of the hot flow facility are described.
3.2.1

Operating Conditions

The hot flow facility was designed as a laboratory-scale shock wave reactor in which flow
conditions (temperatures, pressures, mass flux density) would be comparable to those an-

15

ticipated for a commercial scale reactor. A generalized shock wave reactor that covers both
reactor scales as illustrated in Figure 3.2. It uses an inert carrier gas (steam) to provide

Figure 3.2: Schematic Of Shock Wave Reactor and Gas Temperature Profiles. (Ref.[13])

the reaction enthalpy. This carrier is heated to a temperature Tco , exceeding a target pyrolysis temperature T3 , and expanded to supersonic speed in an array of convergent-divergent
nozzles. This lowers the carriers static temperature, although (for adiabatic expansion)
its enthalpy is retained. The feedstock is preheated, to a temperature Tf o below pyrolysis
levels, and is expanded in supersonic streams parallel to carrier streams via nozzles interleaved with the carrier nozzles. The streams mix, and then pass through a stationary shock
wave whose placement is controlled by throttling the flow appropriately downstream. The

16

supersonic flow is abruptly decelerated at the shock, raising the temperature on a microsecond time scale to T3 , at which pyrolysis proceeds rapidly. The pyrolysis is adiabatic, so
the temperature decreases as the endothermic reactions proceed, to an extent dependent on
the ratio of carrier to feedstock (dilution). When the desired conversion of feedstock is
reached the flow is rapidly cooled to quench the reactions; conventional transferline heat
exchangers (TLE) can be used for this purpose.[32]
3.2.2

Fluid Mechanics Studies

The control over reaction temperature history in the first shock wave reactor relies on the use
of an inert carrier fluid to provide reaction enthalpy. Mixing of this carrier with feedstock
in supersonic flow is a key feature of the reactors operation. While previous studies had
been carried out to determine mixing properties for simple flows (e.g., unconfined flow
of two adjacent streams), little work had been done on the types of interleaved nozzle
supersonic mixing flows characteristic of the first shock wave reactor. For this reason, a
cold-flow facility having geometry similar to that of the hot-flow design was constructed
and operated to measure the degree of mixing of carrier and feedstock versus length. This
work was complemented by analytical mixing and CFD studies[13]. Because of the finite
time and distance required for mixing the carrier and feedstock, a boundary layer develops
during this process. This boundary layer interacts with the shock in a manner that produces
a spread-out region of recompression, which is actually comprised of multiple shocks as well
as a flow separation region[29]. Experiments verified previous correlations between the
extent of this recompression zone and the thickness of the boundary layer with respect to
the channel diameter. For the laboratory scale cold-flow and hot-flow reactors used, the
recompression length amounted to a distance of 30-50 cm, and resulted in a substantial
decrease in total pressure though the studies indicated a rapid degree of mixing occurring
just downstream of the shock system[35]. For commercial-scale reactors, the recompression
is expected to take place by a nearly normal shock, with little additional total pressure loss.

17

3.2.3

Pyrolysis Modeling

Early simulations of ethane pyrolysis using the first shock wave reactor demonstrated that
yields of ethylene could reach levels significantly above those of conventional reactors, as
illustrated Figure 3.3 . This figure plots yield vs. conversion at two reaction temperatures,

Figure 3.3: Ethylene Selectivity vs. Ethane Conversion For Isothermal, Isobaric Pyrolysis
At 1075 K And 1200 K , (Ref.[13])

1200 K, representative of an average temperature in a first shock wave reactor, and 1075 K,
representative of an average temperature in a conventional reactor. The first shock wave
reactor enables the use of substantially higher conversions, with much less decrease in yield.
In conventional reactors, the necessarily longer processing times needed to achieve high yield
are accompanied by degradation of the olefin product, and generation of deleterious products
such as coke and tars which can foul the downstream components. Figure 3.5 shows a point
design of the first shock wave reactor idea for industrial scale ethane cracking, indicating

18

Figure 3.4: Schematic of hot-flow shock wave reactor facility.

operating temperatures and pressures. Despite the fact that the first shock wave reactor
requires much higher steam dilutions than those used in conventional reactors (5:1 vs. 0.3:1
for ethane cracking), if efficient energy recovery from the product stream is implemented, the
energy consumption per kg of ethylene may be less than that for current technology.[8] The
first shock wave reactor project saw the development of a kinetics modeling code to predict
ethylene conversion and selectivity. A mechanistic reaction model for pyrolysis of light
alkanes was assembled by making a critical review of available rate constants. The original
reaction set comprised about 500 reactions and 65 species; a sensitivity study was carried
out to reduce this set to about 230 reactions. Correspondingly, a set of thermochemical
properties was developed using the latest published values for specific heats, enthalpies and
entropies of formation. This is important for determining the reverse reaction rates via

19

detailed balance. Two separate reacting flow codes were developed for modeling the first
shock wave reactor, one based on CHEMKIN and one developed from scratch. The results
from each code were checked to insure consistency. In addition, the codes were tested by
comparing predictions with available yield data for alkane pyrolysis. The codes and reaction
mechanism comprise a very useful tool for evaluation of the effects of parameter variation
on yields for pyrolysis using the first shock wave reactor.
3.2.4

Hot Flow Facility

The hot flow facility, being a laboratory scale facility has certain inherent facility limitations, whilst this is not the case in similar commercial scale reactors. Steady flow conditions
are not achievable in the hot flow facility as it is not capable of the providing the required
power input (2 MW) or advanced species extraction, steam recycling and hydrocarbon storage technologies that can readily complete commercial scale reactor cycles. Therefore, the
hot flow facility, in so far as possible, recreates reacting flow conditions of a commercial scale
in a transient manner using the configuration shown in Figure 3.4. A steam accumulator

Figure 3.5: Point Design Of A Shock Wave Reactor For Ethylene Production

20

generates saturated steam at a nominal 2.5 bar. The steams passage through a Pebble Bed
Heater (containing heated alumina pebbles) superheats the flow to a temperature of 1400 K
with capacity for up to 35 s of continuous flow. Superheated steam behaves as an ideal gas
to a very good approximation and is choked in the series of interleaved nozzles, restricting
mass flow rates to nominal 0.4 kg/s. Ethane is preheated and injected within the nozzle
block, with mixing occurring at conditions below pyrolysis temperature. Mixed flow passing through a series of downstream recompression shock waves is heated in a microsecond
timescale, initiating pyrolysis almost instantly. Reacted products and remaining feedstock
and steam collect in an evacuated 4 m3 dump tank containing arrays of aluminum sheet
that act as heat absorbers to condense for steam. The dump tank is the primary limitation
in experimental run duration. Although steam is condensed out of the flow, the tank fills
and back pressure eventually rises to a point that unchokes the flow. Run times of 15 s are
typical. Argon is used as a tracer for species recognition and inert nitrogen is as a blanket
gas to flush the system of explosive reactants and products.
3.2.5

Results and Future Studies

A plot of yields of the major products is shown in Figure 3.6 , along with theoretical
predictions from our reacting flow code, which includes heat transfer to channel walls. It
is seen that experiment and theory agree very closely for C2 H6 and C2 H4 . Satisfactory
agreement is also seen for acetylene and methane. Most of the ethylene is produced within
first 2 meters of the reaction zone (15-20 msec); again, heat transfer and the resultant
temperature decrease slows the pyrolysis toward the end of the reactor. Operating conditions
and measured yields of principal species for Run 58 are listed in Table 3.1. Ethane conversion
and ethylene yield have reached nearly terminal values by station 13 (15-20 msec reaction
duration). The ultimate conversion from the receiver tank sample of 81.4% is well beyond
that used in conventional ethane cracking, and the ethylene yield of 68.7% by weight is
likewise beyond that attainable in current industry.
Figure 3.7 shows results from a number of runs, with the molar selectivity to ethylene
plotted vs. conversion. These results include runs made without any argon dilution of

21

Figure 3.6: Measured and Predicted Yields of C2 H6 , C2 H4 , C2 H2 , and CH4 From the Shock
Wave Reactor. (Ref.[13])

ethane (overall dilution, via steam only, of 8.5:1) in which the maximum conversion observed
was 60%, and for 2:1 dilution by argon, with conversions running beyond 80%. The data
labeled unstart refer to runs where the backpressure was too high for the nozzles to choke
the flow; the entire flow in these runs was subsonic and the ethane and steam were not
mixed prior to pyrolysis. The lowered molar efficiency under these conditions demonstrates
that good mixing of the streams is needed to use the first shock wave reactor to best effect.
Of interest is the lack of substantial variation in yield at a given conversion under different
operating conditions. The high ethylene yield persists out to very high conversions, in
contrast to experience with conventional reactors. An envelope of yield vs. conversion seen
in current industry demonstrates that substantial gains in pyrolysis efficiency results from
the short-duration, high-temperature pyrolysis profile of the shock wave reactor.

22

Table 3.1: Chemical Yields vs. Station For The Shock Wave Reactor. Ref.(Ref.[13])

The high yields of ethylene measured when pyrolyzing ethane in the first shock wave reactor
indicate that there is an opportunity to improve the economics of olefin manufacture using
this technology. Much of the information on the construction and operating costs of olefin
plants is proprietary, so it was not possible for the University of Washington shock wave
reactor research program to undertake a realistic assessment of the economic benefits of the
use of this reactor. However, based on the continuous efforts made by the petrochemical
industry to improve yields (at far smaller increments than offered by the first shock wave
reactor) it does appear that the first shock wave reactor may be highly advantageous. Based
on the results of the first shock wave reactor, pursuing several interesting areas of study
may be beneficial: E.g. improved nozzle design, control of boundary layers, use of other
feedstocks and carriers, and use of very-high-temperature steam to enable a reduction in
the carrier dilution.[30]

23

Figure 3.7: Measured Ethylene Yield vs. Ethane Conversion For The Shock Wave Reactor
And For Conventional Reactors. Ref.(Ref.[13])

3.3

Alternative Feedstocks: Conversion of Methane

Industrial interest in further shock wave reactor processes using ethane (C2 H6 ) and propane
(C3 H8 ) hydrocarbon feedstocks prompted trade studies. However little economic advantage
was indentified. Subsequently, kinetic simulations of methane (CH4 ) conversion to C2 hydrocarbon intermediaries proved favorable when used in conjunction with a very high enthalpy
carrier, typically 3000 K in the configuration of the first shock wave reactor. Such extreme
carrier operating temperatures can be generated via a clean hydrogen-oxygen combustion
process in this case.
Choice of feedstock in a commercial cracking reactor is based on feedstock availability,
logistics, economics, regional laws amongst others. Ethylene operators typically employ

24

sophisticated computer models and optimization techniques in choosing feedstocks. In the


USA, development of ethylene manufacture is largely based on ethane and propane. Often,
these gasses can be regionally expensive or become costly to transport when remote facilities
are not in the vicinity of wet natural gas reserves so alternative raw materials becomes
necessary. In fact, any fractions of crude oil distillation, up to and including gas oil, can be
cracked to ethylene, with technical difficulties increasing for increased molecular weights.
Use of gas oil or naptha is been popular elsewhere such as in Western Europe and Japan.
Methane is generally not an economical replacement for ethane, propane, naptha or gas
oil.Its inherent chemical stability means that it requires a significant supply of energy for
conversion into useful intermediaries like ethylene. Yet methane accounts for 90% of natural
gas making it widely available. In remote locations, the excess methane from natural gas
cannot be used economically is disposed of, often by flaring. Making use of the the C1
routes (CH4 has one carbon) is considered the holy grail of the chemical industry due to
the potential to make use of large volumes of excess methane in remote locations. [39]
The predominant use of methane is in the manufacture of synthesis gas, the precursor for
ammonia and methanol production. Nonetheless, it has found limited use as a feedstock for
ethylene.
3.4

The Second Shock Wave Reactor Project - Methane Feedstock

In 2009, work began on a laboratory scale test bed for the second shock wave reactor. It
utilizes methane as the feedstock where its predecessor, the first shockwave reactor, used
the market standard, ethane. The first shock wave reactor design has been adapted to the
needs of the second shock wave reactor, having been tailored for the optimal conversion of
methane feedstock to ethylene.
3.4.1

Commercial Scale Methane Reactor

An industrial scale shock wave reactor using methane as a feedstock is examined first,
indicating required flow conditions for operation. In Figure 3.8, the point design of a
commercial scale CH4 shock wave reactor is presented to exemplify the methane processing
conditions. This commercial scale reactor design differs from the first shock wave reactor

25

Figure 3.8: Methane Processing Design Example Ref.(Ref.[13])

design in the following ways. Note that the commercial scale reactors are intended to be
steady state units.

Combustion of hydrogen (H ) and oxygen (O ) produce a higher operating tempera2

ture carrier steam at a temperature of 3100 K.

A single supersonic De Laval nozzle accelerates steam to supersonic flow velocities.


Methane feedstock is injected from the side walls. Intrusive injectors like the array of
interleaved nozzles in the first shock wave reactor require active cooling if they are to
survive the 3100 K steam temperature and thus affecting an adverse heat loss increase.
Sidewall injection is versatile in that injectors can easily be scaled for varying reactor
sizes whilst maintaining adequate chemical and thermal mixing.

Sophisticated fluid dynamical design gradually heats feedstock-carrier mixture to ini-

26

tiate pyrolysis. Gradual heating limits hydrocarbon-steam peak temperatures and


reduces over-cracking of CH4 feedstock or excessive soot production. A supersonic
diffuser or series of oblique shocks, are the two primary means of gradual heating.
During pyrolysis, the endothermic reactions consume mixture enthalpy, lowering the
mixture temperature to 1800K.

Reacting gasses are quenched by heat transfer to preheat methane, oxygen and combustion supply water.

Separation of H

and CH4 for reuse and of C2 hydrocarbons for storage. O2 supply

is anticipated to be from an onsite air distillation plant or similar.


A flow diagram is presented in Figure 3.9 showing relative mass fractions of species and
their corresponding economics on a per ton CH4 input basis.

Figure 3.9: Methane Processing Flow Diagram Ref.(Ref.[13])

27

3.4.2

Laboratory Scale Methane Reactor

Recreating the steady state commercial scale process at laboratory scale is not feasible
mainly due to project cost, space and material constraints. A representative process operating at 2200K has been created that builds on the existing hardware and technical knowledge
from the first shock wave reactor. Key aspects of this test bed are listed below:
1. Added thermal management strategies to better cope with increased operating temperature. These introduce the use of refractory materials and the addition of insulating
liners to combat unappreciated heat losses observed in the first shock wave reactor.
2. Commercial scale temperatures of 3100 K are not viable from a cost and materials
standpoint. A nominal 2200 K carrier steam is more feasible at the laboratory scale.
However, the existing pebble bed heater capacity is not sufficient to generate the
nominal 2200 K steam. Achieving high peak temperatures requires augmentation by
the use of clean H2-O2 combustion. Inherently, the ignition source is 1400 K pebble
bed heater steam.
3. Methane injection was accomplished in a versatile and scalable sidewall injection
scheme analogous to scramjet fuel injection systems. Interleaved nozzle arrays, such
as those used in the first shock wave reactor, were found to be prohibitive in that
they create significant pressure loss and cannot be easily scaled up to handle the
temperature extremes (3100 K) necessary at commercial scale.
4. Temperature extremes and residence time control necessitate that the reacted gasses
be quenched with a water spray bar system before entering the dump tank (evacuated
to 15kPa initially).
3.4.3

Reactor Operation at Design Conditions

Previous hardware from the first shock wave reactor, such as the steam accumulator tank
and pebble bed heater remain in use to supply nominal 1400 K and 2.25 bar steam to the
system. Combustion augemented carrier steam travels through five sections before being

28

discharged into a dump tank.In order, these are the combustor section, feedstock injection
section, mixing section, reactor section and diagnostic section, Constant mass flow of 0.283
kg/s is maintained by a steam nozzle that accelerates the flow to supersonic velocities, hence
choking the flow, and decelerates it again through a normal shock slightly downstream of
the nozzle throat. A total pressure loss of 0.20.3 bar is incurred. The location of this shock
is controlled by the downstream back pressure and the search for optimal design conditions
relating to shock location is ongoing.
The shock wave reactor configuration downstream of the pebble bed heater is shown in
Figure 3.10 . Steam at 1400 K from the pebble bed heater enters a 32 inch long combustor

Figure 3.10: Current Shock Wave Reactor Layout Ref.(Ref.[13])

section. Four co-annular injectors inject H2 and O2 in a molar ratio of 2.5 : 1 that enters
the 4 inch diameter combustor channel transversely. Steam to O2 molar ratio is 6:1. The
H2 rich reactant mixture ensures minimal surplus oxygen is present downstream as oxygen
is detrimental to pyrolysis chemistry and poses and explosion hazard in the presence of
volatiles like CH4 . The high reactivity of nascent oxygen can also be damaging to reactor
internals from oxidation of metal components, especially that of titanium. Ignition occurs
immediately upon contact with 1400 K temperature steam. Complete combustion occurs
over a finite distance, but this was found to be small with respect to the allotted 32 inch
length. The mixing of combustion steam and pebble bed heater steam produce a 2200 K
bath of superheated steam.

29

Following the combustor section is the feedstock injection section.It begins with a nozzle
section containing an axisymmetric supersonic nozzle designed to accelerate the main flow
to a Mach number to 2.2.This is followed by a methane injection section designed to mix
feedstock and carrier at supersonic velocities. The nozzle demonstrates a compromise between a shorter length to avoid heat losses and a longer length to produce a wave free
flow field with minimal pressure loss. Methane is preheated to 850 K before entering sonic
oblique injectors; being side wall injectors they introduce no flow physical obstructions. The
nozzle and methane injector sections each represent significant development efforts and the
details of these are presented in Sections 8, 9 and 11 respectively.
As shown in Figure 3.11, a 30 inch long, 4 inch diameter mixing section subsequently
allows feedstock and carrier stream mixing to occur and reach a homogenously mixed condition. Oblique shocks are present in this section and their presence promotes more rapid

Figure 3.11: Combustor Section, Feedstock Injection Section and Mixing Sections

mixing, reduces overall pressure loss and gradually increases the temperature of the reactants to prevent over-cracking. A weak terminating normal shock situated in the conical
transition (recovery) section represents the final pressure and temperature recovery mechanism. Flow then proceeds through the 10 inch diameter reaction section subsonically and
at high temperature. Pyrolysis starts and ends across the the 60 inch reaction section seen
in Figure 3.12 and is quenched by a spray-bar system (Figure 3.13) to cool flow and freeze
chemistry as it enters the 6 inch diameter diagnostic section where a temperature sensitive
soot sensor is installed. Bellows greatly improve accessibility to the sealed reactor, thus
greatly simplifying the removal of individual sections. A high temperature butterfly valve
leading to the evacuated dump tank is the primary controller of back pressure in the system.

30

Figure 3.12: Mixer Section, Transition Section, Reactor Section and Diagnostic Section

Figure 3.13: Water Quench Spray Bar System

31

Flow is choked across this variable area orifice valve, and at design conditions controls the
pressure drop between the 1.2 bar diagnostic section and 10-40 kPa dump tank. A 9.4 kW
vacuum pump is used to produce the initial 10 kPa dump tank vacuum and operates continuously over the test run duration. Valve locations in the system, flow regulators, tanks
and additional hardware are illustrated in Figure 3.14. Note that the figure is divided in

Figure 3.14: Laboratory Scale Shock Wave Reactor Flow Schematic

two part with the arrow representing the location of the split as well as the flow direction.

32

Chapter 4
DESIGN AND LAYOUT OF THE LABORATORY SCALE REACTOR

The design of the laboratory-scale second shock wave reactor (henceforth referred to as
the reactor) has been an ongoing process even during construction and testing. Regular
design reviews incrementally improve the reactor operation and experimental setup, and
due to this side-by-side development, have become inseparable. With this in mind, the
experimental setup is most clearly presented with the associated design elements presented
alongside it. The use and importance of thermal liners is examined first, followed by the
combustor design. Flow measurement capabilities, reactor layout, sensor locations and
data acquisition methodologies are then described. Feedstock injection sections, being such
a large portion of the design effort, are given their own sections. Chapter 8 presents the
supersonic nozzle design, 4 the injector design, and Chapter 11 the computational fluid
dynamics simulation of flow inside the feedstock injection section.
4.1

Thermal Liner Insulation and Run Duration

Compared with the second shock wave reactor, the first shock wave reactor temperatures
were fairly moderate. Nonetheless, heat transfer effects in the first shock wave reactor were
under appreciated and hence heat loss in the second shock wave reactor was anticipated to
be far greater if left unchecked. Heat loss mechanisms are primarily driven by convection
to reactor walls where their high thermal mass acts as a sponge, soaking up this heat
in large quantities. In transient operation, reactor walls absorb heat but do not heat up
appreciably, keeping temperature gradients and associated heat losses prohibitively high.
The idea of lining the reactor walls with thin sheet-metal tubes came about by noting that
heat loss is slowed significantly if reactor walls are hot. Thin liners have low thermal mass,
heating up quickly and considerably reducing the steep temperature gradient between hot
steam and liner metal. Liners are vented and thus the relatively stagnant pocket of gas

33

(steam) between them and the reactor walls is an effective insulator from conduction and
convection. Radiation remains dominant and calls for the use of double liners in certain high
temperature combustion and mixing sections, as shown schematically in Figure 4.1. Since

Figure 4.1: Wall Insulation System

radiative power, modeled by Plancks blackbody distribution, scales as the fourth power of
temperatures, net radiation that is dominant from inner to outer liners becomes negligible
between the outer liner and the reactor wall. Note that reactor walls alone provide support
against flow pressures and thus liners do not experience significant structural loads.
Initially, exotic materials were explored for liner material; material having low specific heat
capacity (Cp ) and average thermal conductivity (k) was sought. Titanium, with a melting
temperature Tmelt = 1950 K, was initially chosen as the primary liner material; it was
to come into direct contact with moving flow. A series of combustion tests (described
in Section 6.4) later revealed that titaniums reactivity with oxygen at high temperatures
made it unsuitable as a liner material. Nonetheless, during the design stages, titanium was
selected to be used for single liners in the conical transition and 10 inch diameter reactor
sections. In the case of double liners, an inner titanium and outer stainless steel 304 (Tmelt
= 1650K) were utilized, separated by stainless steel ring spacers. In all cases, titanium is
in first contact with the core flow, chosen particularly for its inertness and non-catalytic
action in the presence of hydrocarbon pyrolysis.

34

Run durations were carefully selected to heat up the liners up to high temperature and
within safety limits. A heat transfer analysis of single and double liners concluded that
optimal run durations were as follows. First, 10 seconds of 1400 K steam from the pebble
bed heater and subsequently 5 seconds of combustion augmented 2300 K steam resulted in
suitable liner temperature vs. time profiles. Shown in Figures 4.2(a)and 4.2(b). are the

(a) 3.5 inch Combustor, Dummy and Mixer Liners

(b) 10 inch Reactor Section Liner

Figure 4.2: Simulated Temperature History of Liners

profiles for the 3.5 inch sections with double liners and the 10 inch reactor section with a
single liner.
Liners exposed to such large cyclic temperatures, between 400 K up to 1850 K, are prone to
significant thermal expansion. The thermal coefficient of linear expansion (t ) for titanium
varies linearly with temperature over the range 0.86x105 C

-1

at 298 C to a maximum

of 1.3x105 C-1 near melting. This can be approximated by a first order polynomial fit
t (T ) = (7.755 + 0.002T )x106 C-1 . Hence, the percentage change in any liners is given
by Equation 4.1.1:
L
=
L0

T1 =300K

t (T )dT = 1.53
T2 =1850K

(4.1.1)

35

For a 1.524 m (60 inch) long liner, such as that in the reactor section, this is a non-trivial
change in length of 23.4 mm (0.92 inch). Similarly, the inner titanium liner diameter, 91
mm (3.6), increases by 1.4 mm. The reactor walls and outer liner are constructed from
stainless steel with thermal coefficient of linear expansion of 1.7x105 C-1 . Reactor walls
experience minimal temperature changes, and hence thermal expansion is negligibly small.
Stainless steel inner liners are predicted to reach temperatures of 950 K corresponding to
an expansion of 0.6% or 0.56 mm.
Adjacent liner sections make allowances for thermal expansion by the design of their overlapping joints.
4.2

Combustor and H2 /O2 Injectors

The pebble bed heater is limited to producing 0.283 kg/s steam at 1400 K. This flow rate
corresponds to the choked flow in the steam nozzle driven by a stagnation pressure of
P0 = 2.25 kPa. However, this steam has neither the sufficient reaction enthalpy nor the
high temperature necessary for methane pyrolysis. Combustion augmentation is required
to overcome this limit by heating up pebble bed heater steam flow to a nominal 2300
K, ignoring heat losses. The required enthalpy change in the gas can be estimated using
GasEq [34], a chemical equilibrium code for perfect gasses. Since steam, at temperature
and pressure of 1400 K and 2.25 kPa respectively, is far from its critical points, it is well
approximated by perfect gas theory. Figures 4.3(a) and 4.3(b). show the steam states at
temperatures of 1400 K and 2300 K respectively; enthalpies are -11007 kJ/kg and -8174
kJ/kg respectively, giving an enthalpy change of 2833 kJ/kg. For 0.283kg/s mass flow rate,
this corresponds to an energy rate of 801 kJ/s. The latter is the energy rate required from
combustion augmentation.
Combustion is achieved by a simple hydrogen and oxygen injector configuration based on
injectors available off-the-shelf. Maxon Corporation developed a simple non-premixed combustor that relies on co-annular injection of gaseous fuel and oxidizer for effectively mixing
and burning of natural gas. Mass flow rates are controlled by varying fuel and oxidizer
supply pressure that feed the choked orifices. The shock wave reactor uses similar injection
methodology: Four co-annular steady state injectors shown in Figure 4.4 mix and inject

36

(a) 3.5 inch Channels Liners

(b) Reactor Section Liner

Figure 4.3: Thermodynamic States of Steam at 1400 K and 2300 K (a) and (b)

37

hydrogen and oxygen transversely from ports in the reactor walls. Construction of injectors

Figure 4.4: H2 O2 Injector Cut-away

makes heavy use of off-the-shelf Swagelok fittings as shown in Figure 4.5. Hydrogen and
oxygen emerge downstream of the choked orifices as underexpanded jets with different supersonic flow velocities. Hydrogen orifices are 0.08 inches in diameter. Oxygen is injected
via annular 0.25 inch inner diameter orifices with a 0.01 inch gap. As the injected jets
impinge on one another, shear layer mixing occurs, and subsequently the reactant mixture
enters the flame front where steady state combustion occurs. This flow field is sketched in
cross section in Fig 4.6 (not to scale).
Subsonic steam cross flow at a total temperatures of 1400 K provides the initial ignition
source for combustion, though the details of the ignition process have not been explored
in this work. Subsequent to initial ignition, the combustion is self-sustaining. A key point
of interest was that the rate of heat release during self-sustained combustion would be
sufficiently high so as to allow complete combustion to occur prior to feedstock injection.
This can be thought of as an induction and turbulent mixing time delay, and depends on

38

Figure 4.5: H2 O2 Injector Cross-Section and Combustion Zone Schematic

many factors such as H2 /O2 /H2 Ofeed mass flow rates, Mach numbers, free stream flow
back pressure to mention only a few. Separate from heat release rate considerations, an
ignition delay time also exists for pre-mixed hydrogen-oxygen; the delay is measured from
when the mixture reaches a temperature at which spontaneous combustion occurs until the
subsequent pressure rise associated with detonation occurs. The delay times are predicted
poorly by theoretical predictions at higher temperatures.[12] such as those found in the
shock wave reactor.
Few options for theoretical prediction are available given the highly complex flow field created by the hydrogen - oxygen injectors. Nonetheless, some ignition characteristics are
known. Ignition time scales for pre-mixed hydrogen and oxygen are on the order of 4ms
whilst auto-ignition temperature is around 900 K. An order of magnitude estimate of the

39

Figure 4.6: H2 O2 Injector

cumulative effect of these delays (induction and ignition) is on the order of 1-10 ms. Combustor section length is chosen to be 32 inches nominally and ongoing combustion tests have
yet to confirm that complete combustion occurs to occur before flow enters the feedstock
injection section.From combustion currently acquired data (as described in Section 6.4),
Mach numbers of 0.24 have been calculated and hence the flow passage through the 32 inch
combustor section is approximately 4 ms.
In the immediate combustion zone, reactions are adiabatic and the effect of turbulent mixing with carrier steam is less significant. Assuming that the process is close to equilibrium,
combustion products can be estimated with GasEq. Figure 4.7 shows the combustion products expected in this zone produced by the reaction of 2.5 moles of hydrogen and 1 mole of
oxygen, when each is preheated to approximately 600 K by electric heating wires wrapped
around manifold inlet lines. As previously stated, to avoid oxygen from being present in

40

Figure 4.7: Equilibrium Combustion Products From Adiabatic H2 :O2 Injectants At 2.5 : 1
Molar Stoichiometry

combustion products, a fuel rich stoichiometry is used. The adiabatic flame temperature is
3193 K.
A total mass flow rate from the combustors of 0.081 kg/s is selected to produce the nominal
2300 K mixed bath temperature. Here, the bath refers to the mixed result of two
streams, the carrier steam 2 kPa steam produced from combustion at 3193 K and steam
from the pebble bed heater at 1400 K. The calculation to determine the aforementioned flow
rate proceeds as follows: Beginning with the product set in Figure 4.7 above, equilibrium
results can be used to estimate the heat transfer rate required to bring this mixture to

41

nominal 2300 K. Figure 4.8 shows the results of this procedure performed in GasEq.

Figure 4.8: 2300 K Equilibrium Composition Of Combustion Products And Pebble Bed
Heater Steam

Note that the change in specific enthalpy of the steam is


h = 7340kJ/kg 845kJ/kg = 6495kJ/kg.
Thus the total energy release rate by completely combusted hydrogen and oxygen is:
H = m
h = 0.081kg/s 6495kJ/kg = 801kJ/s
This corresponds exactly with the aforementioned 801 kJ/s required to heat carrier steam
from 1400 K to 2300 K. Figure 4.8 indicates that O2 mass fraction is only 1.2% whilst
trace amounts of dissociated oxygen on the order of 0.026% are present. This is a result of

42

running fuel rich. However, referring back to Fig 4.7 , it is noted that for pure combustion
products close to their adiabatic flame temperature of 3193 K, molecular oxygen levels
remain relatively unchanged whilst dissociated oxygen levels are one hundred fold higher,
at 2.1%.
4.3

Sensors

Refer to Figure 4.9 showing the layout of the sensor ports in the reactor. It illustrates
the locations of 1/8 diagnostic ports in the reactor used to sample both flow, wall and
liner conditions during a run. Varieties of type C and K thermocouples, a custom-built
optical temperature sensor and typically 16 Barksdale pressure transducers are used for
this purpose. These are sealed with Swagelok tube or NPT fittings. Additional type K and
R thermocouples and Omegadyne pressure transducers (not shown in Figure 4.9) monitor
supply and manifold flow conditions, water spray bar supply tank mass, and dump tank
vacuum pressure. A mass spectrometer and gas chromatograph both sample flow at the
end of the reactor for the purpose of analyzing pyrolysis yields. Since the experiments
are performed with steam, all section and manifolds and sensor flow paths are equipped
with an electrical heating system to avoid condensation. As additional precaution, pressure
transducers are filled with a small quantity of Fomblin polyfluoropolyether lubricant to
prevent condensation and provide a measure of thermal insulation for the sensors in zones
of high gas temperature.
Thermocouple type selection, assembly and installation require careful consideration in the
adverse conditions present in the reactor. Only type C thermocouples (operating up to
2300 C) are capable of measuring the extremes of temperature present downstream of the
combustion zone. Though pure oxygen is injected in the combustion zone, the sensitivity of
type C thermocouples to oxygen at high temperatures is not a concern since the combustion
fuel mixture is hydrogen rich. Post-combustion, oxygen is present only in trace amounts
downstream of the injection point, as shown previously in Figure 4.7.
Both Type C and K thermocouples typically measure flow temperature over the length of
the reactor. However, the time constant (or e-folding time) for the type K thermocouples
used in the reactor is far longer than that for type C, thus implying a longer response

43

Note: Figure has been divided to fit, as indicated by arrows.


Arrows also show flow direction.

Figure 4.9: Sensor Port Layout in the Shock Wave Reactor

44

time. By considering a simple first order heat transfer model of thermocouple the constants
and physical quantities driving the time constant are evident. Conservation of energy requires that the internal energy change (Equation 4.3.1) in the thermocouple hot junction be
balanced by the convective heat flux (Equation 4.3.2) from the surrounding hot fluid flow:
ZZZ

de
dT
= mCp
dt
dt

(4.3.1)

ZZ
qdA

= hA(T T )

(4.3.2)

Hot junction properties are denoted by density (), internal energy (e), volume (V ), mass
(m), temperature (T ), specific heat capacity (Cp ), convection surface area (A), and surface
convective heat flux (q).

Flow properties are denoted by the convection coefficient (h)


and free stream flow temperature (T ) The independent variable t represents time. The
equation of energy conservation and be reformulated as Equation 4.3.3:

dT
+ T = T
dt

where =

mCp
hA

(4.3.3)

Solving this first order ordinary differential equation yields Equation 4.3.4:

T (t) = T (1 e

(4.3.4)

Note that the quantity ,the time constant for a thermocouple, is proportional to mass. The
type K thermocouples used in the reactor have hot junctions that are larger in diameter
than those of type C. This additional mass of the type K thermocouples increases the time
constant, and explains why type K thermocouples show significantly slower response to
transient dynamics of flow temperature than do type C thermocouples.
A table of sensors typically present in the system is summarized in Table 4.1 listing sensor
type, manufacturer, calibration, and expected. Sensor location is determined by cross referencing with Figure 4.9. Note that this table changes from run to run as the limited number
of sensors are rearranged to probe flow at stations of interest.
Thermocouples are limited in their ability to measure liner temperatures. Generally, thermocouple hot junctions must be directly attached to liner walls (i.e. via welds) and make

45

Line
0
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64

Multiplexer
Channel Name
0-0
0-1
0-2
0-3
0-4
0-5
0-6
0-7
0-8
0-9
0-10
0-11
0-12
0-13
0-14
0-15
1-0
1-1
1-2
1-3
1-4
1-5
1-6
1-7
1-8
1-9
1-10
1-11
1-12
1-13
1-14
1-15
2-0
2-1
2-2
2-3
2-4
2-5
2-6
2-7
2-8
2-9
2-10
2-11
2-12
2-13
2-14
2-15
3-0
3-1
3-2
3-3
3-4
3-5
3-6
3-7
3-8
3-9
3-10
3-11
3-12
3-13
3-14
3-15

Reactor
Location
H2MP
02MP
H2TP
02TP
MTHP1
MTHP2
MTHP3
C1L
C2L
C3L
M1L
M2L
M3L
M4L
M5L
T1L
T2L
T3L
R1L
R2L
R3L
D1L
N1L
DT1
PBHBP
PBHTP
OTS
WTM
PBHTTC
PBHBTC
PBHMTC
PBH
H2MTC
D1T
C1T
D1B
N1R
M5T
R1R
D1R
M3R
02MTC
C3T
R2T
R2R
R3T
N1T
MFR

Function
H2 Manifold Pressure
O2 Manifold Pressure
H2 Tank Pressure
O2 Tank Pressure
CH4 Pressure 1 - out
CH4 Pressure 2 - out
CH4 Pressure 3 - out
0-7 NOT ASSIGNED
Combustor Pressure 1
Combustor Pressure 2
Combustor Pressure 3
Mixer Pressure 1
Mixer Pressure 2
Mixer Pressure 3
Mixer Pressure 4
Mixer Pressure 5
Transition Pressure 1
Transition Pressure 2
Transition Pressure 3
Reactor Pressure 1
Reactor Pressure 2
Reactor Pressure 3
Diagnostic Pressure 1
Dummy Section Pressure 1
Dump Tank Pressure
1-9 NOT ASSIGNED
1-10 NOT ASSIGNED
1-11 NOT ASSIGNED
1-12 NOT ASSIGNED
1-13 NOT ASSIGNED
1-14 NOT ASSIGNED
1-15 NOT ASSIGNED
PBH Bottom Pressure
PBH Top Pressure
Optical Temperature Sensor
Water Tank Mass
2-4 NOT ASSIGNED
2-5 NOT ASSIGNED
2-6 NOT ASSIGNED
2-7 NOT ASSIGNED
PBH Bed Top (Bow) Temp - out
PBH Bed Bottom (Bow) Temp - out
PBH Bed Mid (Starboard) Temp - out
2-11 NOT ASSIGNED
PBH Head Temperature
Hydrogen Manifold Temperature
Diagnostic Flow Temp
Combustor Entrance Flow Temp
Diagnostic Temp Bottom
Dummy Liner Temp
Mixer Flow Temp Exit
Reactor Flow Temp Entrance
Diagnostic Center Flow Temp
3-5 NOT ASSIGNED
Mixer Liner Temp
Oxygen Manifold Temperature
Combustor Flow Temp Exit
Reactor Flow Temp Mid
Reactor Flow Temp Mid
Reactor Flow Temp Exit
Dummy Flow Temp
3-13 NOT ASSIGNED
3-14 NOT ASSIGNED
3-15 NOT ASSIGNED
Mass Flow Rate

Probe Distance
to Centerline
(inchs)
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.0
2.75
1
1
Liner
0.5
2
1
Liner
0
3
3
0
0
-

Calibration
0-5V: 0-1000psia
0-5V: 0-1000psia
0-5V: 0-3000psia
0-5V: 0-3000psia
0-5V: 0-3000psia
0-5V: 0-3000psia
0-5V: 0-3000psia
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
0-5V: 1V/bar
-2V to +2V
0 to +10V
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned
0-5V Conditioned

Sensor
PX319 Omegadyne
PX319 Omegadyne
PX319 Omegadyne
PX319 Omegadyne
PX319 Omegadyne
PX319 Omegadyne
PX319 Omegadyne
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Barksdale
Opt Sensor
Drum scale output
Type R Thermocouple
Type R Thermocouple
Type R Thermocouple
Type R Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type K Thermocouple
Type C Thermcouple
Type C Thermcouple
Type C Thermcouple
Type C Thermcouple
Type C Thermcouple
Type C Thermcouple
Type C Thermcouple
Type C Thermcouple

Table 4.1: A typical set of sensors in the reactor

Source
Pr Box 1
Pr Box 1
Pr Box 1
Pr Box 1
Pr Box 1
Pr Box 1
Pr Box 1
Pr Box 2
Pr Box 2
Pr Box 2
Pr Box 2
Pr Box 2
Pr Box 2
Pr Box 2
Pr Box 2
Pr Box 2
Pr Box 3
Pr Box 3
Pr Box 3
Pr Box 3
Pr Box 3
Pr Box 3
Pr Box 3
Pr Box 4
BNC1
BNC2
BNC3
BNC4
Temp Box R & K
Temp Box R & K
Temp Box R & K
Temp Box R & K
Temp Box R & K
Temp Box R & K
Temp Box R & K
Temp Box R & K
Temp Box K
Temp Box K
Temp Box K
Temp Box K
Temp Box K
Temp Box K
Temp Box K
Temp Box K
Temp Box C
Temp Box C
Temp Box C
Temp Box C
Temp Box C
Temp Box C
Temp Box C
Temp Box C

Units
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Pressure (kPa)
Temperature (C)
Water Mass (kg)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Temperature (C)
Mass Flow Rate (kg/s)

46

excellent thermal contact over the wide range of operating temperatures (300 to 1800 K)
to measure consistent data. Thermal expansion can often cause detachments over a several
run cycles. An optical sensor was found to be superior to thermocouples in their ability to
measure temperature unobtrusively (no physical attachment) in extreme temperature conditions that would otherwise damage some thermocouple types. The sensor, constructed
in-house, comprised an optical collector (sapphire rod) capturing incoming radiation, optical
fiber radiation path with transmission efficiency of 60%, and a pair of photodiodes sensitive
to the visible and UV spectrum. The sensor is diagrammed in Figure 4.10. The pair of

Figure 4.10: Optical Temperature Sensor Schematic

photodiodes each capture the grey-body emission spectrum from the hot liners. A greybodys emission spectrum varies due to different temperatures, as sketched in Figure 4.11.
Raw voltages from the photodiodes enter signal converter box which electronically takes
the log of the ratio between the two signals. The log-ratio voltage output is non-linear
with temperature, though results are reproducible with minimal hysteresis. A calibration
curve was obtained by heating a specimen of titanium and measuring corresponding surface
temperature via a directly attached thermocouple.It is noted that the optical sensor is only
accurate above 600 C due to its limited sensitivity to low radation levels emmitted by

47

Figure 4.11: UV And Visible Radiation Intensity Variation With Temperature

samples below this temperature.


4.4
4.4.1

Hardware and Data Acquisition


Signal Conditioning

In general, sensors outputs are low voltage signals on the order of millivolts. When unshielded, high levels of electromagnetic interference generated by high power laboratory
equipment such as the vacuum pump, pebble bed heater were noted to cause spurious
voltage excursions from expected voltage levels. Hence, voltage signals are electronically
conditioned in custom built circuitry that is shielded within enclosed aluminum boxes. All
pressure transducers are individually calibrated and zero-offset to ensure measured signals
are accurate to within manufactured precision. Calibrations are applied later at a software level. Raw pressure transducer voltages are gained electronically and thermocouples
receive electronic cold junction compensation, filtering and gain before reaching the data
acquisition boards. These conditioned signals are consolidated in a custom built 4 channel
multiplexer allowing 16 input signal lines per channel. A multiplexer is necessary due to

48

the high number of input signals (up to 60) and limited analog inputs (32) on the computer
data acquisition card. An organizational chart showing the data acquisition system is shown
in Figure 4.12

Figure 4.12: Data-Acquisition Organizational Chart

4.4.2

LabVIEW Software Development

LabVIEW version 8 software is the development environment used for software data acquisition, real-time data display and automated test run control. Used in conjunction with a
National Instruments NI PCI-6229 data acquisition card, a LabVIEW virtual instrument
(VI) provides a front-end graphical user interface (GUI) for the real-time monitoring and
saving of voltage signal data during a test run, real-time valve state monitoring, safety monitoring and controlling test start times (screen captures are shown in Figures 4.13 and 4.14).

The back-end block diagram in the VI (a full set of screen captures are presented in Appendix
A: Figures A.1 to A.5) performs three tasks in parallel, as follows:

49

Figure 4.13: LabVIEW Front Panel for Valve Control

50

Figure 4.14: LabVIEW Front Panel for Data Display

Data acquisition (signal input, display and saving) with multiplexer control,
Sequenced valve opening and closing with precision timing
Safety monitoring
At the time of writing, the VI remains under development, though core functionality has
been implemented.

51

4.4.3

Software Data Acquisition

When the VI is run, software data acquisition of raw signal voltages begins immediately
and are displayed in real-time. Note that data saving and valve control sequences do not
begin immediately but wait until a user presses the Start button to begin a test run. In
the GUI, data acquisition signals are partitioned into tabs by their physical origin (box), as
in Figure 4.14, and this natural layout makes troubleshooting sensors easier. Conditioned
signals from sensors are displayed in real-time charts as raw voltage, and their corresponding
calibrated values for pressure, temperature and mass are displayed as numbers for monitoring purposes. The VI controls the multiplexer switching by directing each channels 16
input voltage signals to their respective charts in the GUI. A binary signal is produced by
the VI specifying channels 1 through 4 as 0b00 0b01 0b10 and 0b11 respectively and is sent
to the multiplexer at a switching frequency of 28-32Hz (depending on real-time computer
performance). The switching frequency can be user specified and determines sampling rate.
Typically the delay between data samples is 120-140ms on average.
4.4.4

Starting a Test Run

When a run is started using the Start button, all proceeding activities are automatic.
Users need only set initial values like data sampling rate and fail-safe temperature thresholds before pressing Start. The aforementioned three parallel activities begin and their
algorithms are shown in Figure 4.15. Test time begins at time t = -5.0 s when the main
timer and safety timers are started. Five seconds of buffer data are stored before the valves
are opened and the test run begins at time t = 0.0 s. At this point, the valve timer starts
and the experiment begins. Incoming data raw signal voltages and calibrated signal voltages
are written to formatted text files as ASCII with 6 significant digits of precision.
Valves are automatically opened in a pre-determined sequence with precise timing shown in
Appendix A: Table 3.4.7.5 A.1. The order in which valves open is dependent on a variety
of physical and practical considerations and are further discussed in Section 3.4.8.

52

Figure 4.15: Software Algorithm Flow Chart of LabVIEW VI

53

4.4.5

Post-Processing

Data from the LabVIEW VI are stored in eight ASCII (text) files, four of which are raw
voltage data, with the remaining four being calibrated data files that contain pressures,
temperatures and masses. The four calibrated files are used only for purposes of ensuring
that real-time data displayed during a run is consistent with ongoing re-calibration activities
when sensors require adjustment. Raw voltage data is read by a MATLAB script and
GUI developed for the purpose of data-post processing, simple plotting and visualization.
MATLAB code is presented in Appendix B. The code reads data files, consolidates data
into a tabular array for exporting purposes, calibrates data and plots the temporal and
spatial profiles of pressure and temperature inside the reactor.
4.4.6

Sensor Calibration

Calibrations are carried out via mathematical formulae in the aforementioned MATLAB
post-processing script and this section describes the origin of the formulae. Please refer
to Table 4.1 in section 4.3 listing sensor types. Barksdale pressure transducers receive
individual electronic calibration and conditioning in order that they produce 1V/bar of
pressure as a raw signal. The electronic conditioning of the each transducer voltage occurs
in the electronic boxes where a simple operational amplifier circuit and filter circuit are
individually adjusted.True pressure is confirmed by a precision calibrated, standardized
transducer. On each transducer in the reactor, a two step process is used: First, a hard
vacuum is drawn on each transducer and the zero offset is adjusted to read the same as the
standardized transducer. Second, transducers are exposed atmospheric pressure conditions,
confirmed with both the meteorological standard and the standardized transducer. As
seen in Appendix 1: calibrate.m, conditioned raw voltages from these sensors need only
be multiplied by a factor of 100 to produce pressure in kilopascals. Omegadyne pressure
transducers do not use the same procedure since existing three-point calibration curves
are available that are linearly interpolated, as seen in Appendix B.3. Thermocouples are
cold junction compensated and their polynomial calibration curves are obtained from the
National Institute of Standards and Technology reference database for types K and R and

54

from the Omega reference library for type C


4.4.7

Quench Water Mass Flow Rate Calculation

Quench water entering the reactor during a run is driven by pressure from an inert gas (N2
or Ar) and sprays into the reactor section through an array of 33 by 0.94 mm-diameter
ports bored into a 5/8 stainless steel tube. Water mass inside the quench water ballast
tank is measured with a 0-100 kg electronic scale and calibrated according to the following
tare formula: m = 21.991V + 1.06082 where m denotes water mass and V denotes measured
voltage.
During a run, on the order of 2-5 kg of water is sprayed into the test section over 35 s,
representing a very small change in mass when compared with the range of the electronic
scale. Hence, signal noise is found to be high, generally about

0.5 kg peak-to-peak. This

is exemplified in Appendix C: Figure C.1 which illustrates measurements acquired during a


run with two phases, both low and high water flow rate. Note that, in this example, water
injection timing is not the same as that used in nominal reactor conditions; the low flow
rate phase runs from 1.0 to 15.0 seconds and from 20.0 to 30.0 seconds whilst the high flow
rate phase runs from 15.0 to 20.0 seconds.
Noise reduction techniques are addressed in Appendix C and results are shown in Appendix C: Figures C.3 and C.9.

55

Figure 4.16: Noise in the water mass vs. time plot with trendline

56

Chapter 5
OPERATION OF THE LABORATORY SCALE REACTOR
5.1

Reactor Startup Sequence and Valve Timing

The following section describes the anticipated startup sequence for a reactor run to successfully reach design conditions, avoid unstart of choked orifices and to heat up liners
appropriately. Before a test run begins, the steam accumulator tank contains 2.25 bar saturated steam and dump tank brought to vacuum, containing ambient air at 10 kPa. The
Start button in the LabVIEW VI is pressed, initiating a run. The order in which valves
open is dependent on a variety of physical and practical considerations, and the sequence is
divided into five phases, as shown in Figure 5.1. Timing is shown in detail in Appendix A:
Table A.1.

Figure 5.1: Five Phases Occuring During A Run

The Pre-Start phase occurs during the first 5 seconds to prepare the reactor before the
run starts. In general, valves are set to their initial open or closed positions. The heater
isolation valve is opened to evacuate the feedstock heater chamber and subsequently closed

57

at t = -0.5 s to prevent steam from entering the chamber before feedstock injection begins.
Following this, the Steam phase sees the opening of the pebble bed heater and dump tank
valve, initiating steam flow for 10 seconds. The start of this phase marks the beginning of
what is referred to as a run. The pebble bed heater valve is fitted with a metering device
for control of the speed at which the valve opens. The valve is nominally set to open in
1.75 s, primarily to avoid sharp pressure gradients in the pebble bed heater that run the
risk of entraining alumina pebbles into the reactor system. In Appendix A: Table A.1 the,
two dump tank valve settings are described, namely Partial-Open and Fully-Open. The
latter indicates that the dump tank valve is partially opened a predetermined amount whilst
the former indicates that the dump tank valve is held fully open. The calibration of this
valve is explained in greater detail in Section 6.1. Quench water injection starts at t = 1.0
s instead of t = 0 s to compensate for the delay in pebble bed heater valve opening. Water
enters the reactor via the spray bars to cool steam flow entering the diagnostic sections
and dump tank. At this stage, the quench water mass flow rate is low to ensure that all
quench water is vaporized and no liquid water remains in the diagnostic section in the form
of droplets or pools. These can be damaging to instrumentation due to erosion or even
simply detrimental to thermocouple measurements if droplets impinge on a hot junction.
The liners heat up to about 1200 K during the Steam phase and the system is purged of
any blanket gasses by the steam flow.
Next, hydrogen and oxygen injection begin during the Combustion phase, starting at t
= 10.0 s for a total duration of 5.0 seconds. Liners heat up further and reach about 1700 K
in preparation for methane injection. Water flow rate is increased for additional cooling of
this higher temperature flow.
During the Feedstock Injection phase that starts at t = 13.0 s, methane isolation valves
open for 2.0 s to initiate pyrolysis. Liners begin at peak design temperature of 1700 K and
continue to heat up in the combustor section whilst cooling off downstream of feedstock
injectors due to injection of 850K methane. . The methane and steam mixture is cooler
than liners in terms of total and static temperature thus causing a reversal of heat transfer
direction to one in which heat flows from liners to core flow for a short period.
The subsequent Shutdown phase terminates the run, ending injection of hydrogen, oxygen

58

and methane by shutting their isolation valves. Pebble Bed Heater steam continues to flow,
thus purging the reactor sections of volatiles. Flow of inert gasses (argon or nitrogen) begin;
they are used to purge the manifold system of volatiles when purge ballast isolation valves
are opened at t = 15.0 s. Data continues to be acquired for 20 s after the run has ended.
5.2

Lab Safety and Safety Shutdown Logic

In situations where unsafe temperatures or pressures are detected in the data acquisition
system, certain safety thresholds are present that can trigger a controlled shutdown of the
reactor. Safety shutdown logic is shown in Figure 5.2.

Figure 5.2: Safety Control System Logic

Upon completion of a run, the dump tank is filled with volitiles. Leftover hydrogen from
combustion is present.Both reactant and product gasses from pyrolysis such as methane,
ethylene and acetylene are present. The dump tank used in the experiment is sized sufficiently to withstand explosion from any of these chemicals if an anomalously high level of
oxygen allows combustion to occur.

59

Chapter 6
TESTING OF THE LABORATORY SCALE REACTOR

After reactor hardware had been assembled in untested fashion, it was deemed prudent
to individually test each of the sub-systems in the reactor, primarily the combustor and
quench water injectors, before running them all simultaneously at design conditions. The
complexity of the entire apparatus, data acquisition, control system, piping layouts and
heating system made a strong case for separate testing to reveal any bugs, sensor issues,
leaks or design flaws. This was a safety concern as well as a good means of isolating these
problems to a particular sub-system. Note that sub-system testing began prior to fabrication
of the feedstock injection section and thus, a dummy section is currently substituted in
its place. The dummy section is simply a cylindrical channel with liners, similar to the
combustor and mixer sections.
At the time of writing, sub-systems are being tested in conjunction with one another using
an incremental procedure. That is, by adding each sub-system to the reactor in turn. In
particular, combustion tests and water spray bar operation are evaluated below in Sections
6.4 and 6.4. The final stage of testing remains to be conducted: to remove the dummy
section, insert the actual feedstock injection section, and test the reactor with all sub-systems
in place. For safety reasons, inert gas will initially be used in the feedstock injection section
instead of volatile methane to ensure that leak free operation under nominal conditions is
realized.
In addition to sub-system testing, several calibration tests were conducted to distinguish
actual flow conditions realized in the reactor from those predicted by theoretical design
calculations. In particular, the dump tank butterfly valve and H2 / O2 injectors were key
mass flow rate calibration areas that required in-place testing. Their orifices are choked at
design conditions and determination of accurate mass flow rates was necessary for chemical
mass balances for the evaluation of C2 yields.

60

6.1

Dump Tank Butterfly Valve Calibration

As described in Section 5.1, the dump tank is connected to the reactor by a high temperature
butterfly valve. It is pressure actuated and can be set to two positions, either Partial
Open or Full Open during a run. A pair of pressure regulators and a pair of solenoids
actuate 791 kPa (100 psig) laboratory supply air to achieve these two valve states. The
valve opening sequence in Appendix A: Table A.1 previously indicated that the dump tank
valve was partially open during the steam phase (see Figure 5.1) though the degree to
which it is opened was not specified. This Partial Open valve setting is controlled by
regulated actuating pressure. When an actuating pressure of one atmosphere or 101.3 kPa
(0 psig) is used the valve is shut whilst when 197 kPa (14 psig) is used the valve is fully
open. Intermediate actuating pressures result in a partially open valve position. The orifice
area corresponding to a Partial Open setting varies non-linearly with actuating pressure.
Because it is choked, the orifice area determines steam mass flow rate whilst also affecting
the pressure in the reactor upstream of the dump tank valve. Effectively, this upstream
pressure is the backpressure seen by the steam nozzle during the steam phase. Hence,
reactor pressure (in kPa) can be seen to be a function of actuating pressure (in psig). This
relationship was measured and is plotted in Figure 6.1 together with a 6th order polynomial
fit. The equation for the curve fit is given by:
Preactor = 0.010 p6 + 0.588 p5 12.95 p4 + 147.6 p3 909.9 p2 + 2804 p 3128.
Note that for a given Partial Open setting the reactor pressure remains constant due to
the choking of the steam flow, despite the small rise in dump tank pressure over the course of
the 10 second steam phase, as shown in Figure 6.2. The resulting pressure profile when the
butterfly valve is set to Partial Open is exemplified in Figure 6.3.Note that this illustrates
the spatial distribution of pressure measured at various stations during the quasi-steady 10
seconds steam phase.
6.2

Spray Bar Tests

Spray bar tests were conducted to determine quench water mass flow rate realized during
a run and their effectiveness at cooling the flow. The mass flow rate of water during a

61

Figure 6.1: Dump Tank Butterfly Valve Calibration (at t = 10 s)

Figure 6.2: Dump Tank Pressure vs.Time

62

Figure 6.3: Spatial Pressure Distribution Controlled by Butterfly Valve Setting

run was calculated from measured water ballast tank mass (following the procedure in
Appendix C) . Shown in Figure 6.4). is a rough calibration curve for the variation of
flow rate with injection pressure. Steam flow temperatures downstream of the spray bar
(in the diagnostic section) were measured to verify the cooling potential of the spray bar
system. Flow temperatures are to remain low in the downstream diagnostic section in order
to prevent damage to the temperature sensitive soot sensor and to prevent melting of the
array of aluminum condensing plates within the dump tank. Inside the diagnostic section,
temperatures were measured in three circumferential locations: Center line, halfway between
wall and centerline measuring 50.8mm (2 inches) the reactor wall inner diameter, and finally

6.35 mm ( inch) from the reactor wall inner diameter. Example profiles of temperature vs.
time at these locations are shown in Figure 6.5 for a combustion test that used 50% of

63

Figure 6.4: Water Mass Flow Rate Calibration

nominal combustor mass flow.


During 1400 K steam flow between 1 and 15 seconds, heat transfer effects are evident as
temperature decrease from the centerline flow toward the reactor wall. During the combustion phase, from 15 to 18 seconds, superheated steam flow leaves the reactor section and
enters the diagnostic section at a peak temperature of 1168 C. Quench water reduces this
temperature to 937 C at the center line, 734C at 50.8 mm from the wall and approximately
110 C at 6.35 mm from the wall. In the latter case this indicates the presence of saturated
steam and incomplete quench water vaporization. Center line flow temperatures are higher
due to heat losses and a hotter combustion core. A new spray bar has been designed to
compensate for higher center line temperatures via a distribution of spray holes that is more
concentrated in the middle, thus allowing a higher mass flow rates at the core of the flow.
The effect of water mass addition and vaporization on pressure was found to be small.

64

Figure 6.5: Temperature at Three Wall Distances in the Diagnostic Section

6.3

Heat Losses

Heat losses in the reactor increase with increasing internal flow temperature. Thus, peak
heat loss is observed during combustion. Note that flow on the reactor centerline during
combustion is expected to be hotter than outer flow due to incomplete thermal mixing of
combustion steam and pebble bed heater steam. Heat loss during combustion has not been
measured directly but a lower bound on the temperature drop due to heat losses can be
estimated based on tests using only 1400 K pebble bed heater steam. Figure 6.6 plots this
temperature difference taken between the start of the combustion section and the end of the
reactor section. Note that the temperature drop is expected to be greater than this during
combustion. At t = 10 s, t = 15 s and t = 18 s the temperature drops are 195 C, 158 C and
146 C respectively. Heat loss decreases with time as liners approach operating temperature.

65

Figure 6.6: Temperature Drop vs. Time Across the Length of the Reactor (with 1400 K
Steam Only)

6.4

Combustion Tests

Cold combustor tests were initially performed using unheated inert nitrogen instead of hydrogen and oxygen combustibles. These allowed flow rates of the hydrogen/oxygen injectors
as well as the correct operation of all sub-systems and sensors to be safely confirmed. Nominal operation was observed.
Three hot combustor tests have been performed to date using hydrogen and oxygen. The
first used 25% of the nominal combustor mass flow for 1 second. The second and third test
used 50% and 75% nominal combustor mass flow, respectively, for 3 seconds. The second
combustion test was most successful in demonstrating combustion augmentation for high
temperature steam production. These results and discussion are presented first, followed
by issues that were encountered during the third combustion test.

66

6.4.1

Successful Combustion Augemented Steam - Steam Combustion Test

Three pertinent questions arise when examining results from the combustion tests. Ultimately it is not currently possible to address each in a comprehensive manner given the
limited data that has been acquired. However, examining the data available does take a
step toward answering each question (listed below):
1. Complete Combustion: With few avenues for theoretical combustion modeling, mentioned previously in Section 4.1, is complete combustion and thermal mixing achieved
by the end of the nominal 32 inches combustor section?
2. Peak Pyrolysis Temperature: Are combustion augmented steam temperatures sufficiently high for methane pyrolysis?
3. Liner Temperatures: How closely do liner temperatures follow those predicted by
theory and are liner temperatures in the correct range when methane is to be injected
at t = 18 s.
Complete Combustion and Thermal Mixing
Data of interest in the second combustor test are mass spectrometer data (species partial
pressure vs. time) and temperature vs. time plots in the dummy section, at end of the
reactor, and for the liners in the combustion section. Combustion pressure at the exit
of the combustor section is also included; this is shown in Figure 6.7 and illustrates the
characteristic rise in pressure when combustion is initiated at t = 15 s.
Unaltered mass spectrometer data measured in the reactor section is shown in Figure 6.8.
Note that the time scale does not correspond with the standard run times in this document,
and hence, run start and end times have been marked, together with combustion start and
end times. As the run begins, air is swept from the system as seen by the rise and fall of
nitrogen and oxygen in their well known 78% to 21% ratio. At the onset of combustion,
H2 Ocontent rises substantially as expected. Oxygen and hydrogen levels do not rise significantly from the level observed from t = 1 to 15 s during the 1400 K steam phase level,

67

Figure 6.7: Successful Combustion Test: Combustion Pressure

indicating that hydrogen and oxygen react to completion by the end of the reactor section
and combustion is complete.
Refer to Figure 6.9 where steam flow temperature in the dummy section, immediately
downstream of the combustor section, is plotted.
Combustion begins at t = 15 s, and a sharp transient temperature rise is observed. At first,
this transient may appear to be as a result of the startup transient behavior of hydrogen and
oxygen gasses as their flows are initiated but this is more likely to be due to the response
time of the type C thermocouples. Peak temperature of 1674 C is reached at t = 18 s
To better evaluate thermal mixing, a theoretical prediction of the final mixed temperature
of combustion augmented steam can be obtained. This analysis can be made using a simple
first law balance between the stream of pebble bed heater steam at 1400 K and the stream
of combusted hydrogen and oxygen at 3192 K.
The energy release rate required from combustion in order to raise 1127C (1400 K) pebble
bed heater steam up to nominal 2026 C (2300 K) can be obtained from GasEq software,

68

Figure 6.8: Successful Combustion Test: Mass Spectrometer Species Data

where the process is assumed to be quasi-steady state:

E in = m
steam hsteam = (0.283kg/s) ((11008kJ/kg K) (9515)) = 432.8kJ/s
(6.4.1)
By a simple first law balance, neglecting heat losses, the energy into the pebble bed heater
steam is equal to the energy out of the combustion process. That is:
E in = E out = m
comb hcomb = 432.8kJ/s

(6.4.2)

For complete combustion at 50% mass flow of hydrogen and oxygen mixture (0.50.081kg/s =
0.0405kg/s), this corresponds to a change in specific enthalpy of:
hcomb =

E out
432.8kJ/s
=
= 10686kJ/kg
m
comb
0.0405kg/s

(6.4.3)

69

Figure 6.9: Hot Combustion Test 2: Flow Temperature Immediately Downstream of Combustor Section

Since the initial enthalpy for hydrogen and oxygen reactants in a mixture is h1,comb =
843.4kJ/kg the final combusted and mixed enthalpy state can be determined from the
change in specific enthalpy from Equation 6.4.3:
h2,comb = h2,comb + hcomb = (843kJ/kg) + (10686kJ/kg) = 9842kJ/kg

(6.4.4)

The corresponding thermally mixed temperature for this enthalpy state is 1545.2 C (1818
K). This is lower than the measured temperature (1674 C) on the center at the line end
of the combustor. It is not possible to draw firm conclusions about complete combustion,
though it is clear that if the steam were to be thermally mixed, one would expect 1545.2 C
to be measured with no heat losses. With a rough heat loss compensation of Tloss = 146 K
based on the minimum heat losses estimated prior in Section 6.3, this number drops to 1399
C.

Because the centerline combustor temperature of 1674 C is higher than the theoretical

thermally mixed temperature of 1399.2 C, it is then possible to infer that the thermal

70

mixing is not complete by the exit of the combustor. A likely cause of this inhomogeneity is
the presence of a hotter core flow as established by combustion jets impinging and releasing
heat at the centerline whilst combustion heat lease occurs less toward the reactor walls.
The flow at the end of the reactor is plotted in Figure 6.10. The peak temperature at t = 18
s is 1160 C on the centerline. This is 240 C below the theoretically predicted thermallymixed temperature of 1399 C. Two possible causes of this are that heat losses are severely
underpredicted in Figure 6.6 or combustion products did not release enough heat to raise
the temperature to 1399 C.

Figure 6.10: Hot Combustion Test 2: Reactor Exit Temperature

Peak Pyrolysis Temperature


To ensure that the peak pyrolysis temperature is nominal, the peak combustion augmented
steam temperature must also be nominally around 2000 C. An estimate for the latter can
be obtained by extrapolating peak temperatures acquired using lower combustion mass flow

71

rates. The peak temperature will occur in the dummy section after 3 seconds of combustion, but with available data, an assessement can only be made for the peak temperature
after 1 second of combustion. Figure 6.11 shows a graph of combustor mass flow rates vs.
peak temperature at a time which is one second after combustion begins. An exponential
decay trendline has been added to allow rough extrapolation. This gives an expected peak
temperature of 1810

70 C at 95% confidence level.

Examination of a previous combus-

tion plot Figure 6.9 from hot test 2, flow temperature is seen to rise 50 C per second or
more during combustion and thus 1910

70 C.

Nominal conditions for the combustion

augemented steam temperature fall within these error bounds, giving confidence that peak
pyrolysis temperatures will be sufficient.

Figure 6.11: Peak Combustion Temperature After 1 sec of Combustion

72

Liner Temperature
The combustor section inner titanium liner temperature was measured by the optical temperature sensor and is presented in Figure 6.12. As mentioned in Section 4.3, the sensor
only displays accurate results above 600 C. Peak liner temperature of 1267 C (1540 K) in
the plot occurs at t = 18 s after 3 s of combustion. At full combustor mass flow rate, a tem-

Figure 6.12: Hot Combustion Test 2: Combustor Liner Temperature

perature of 1700 K is expected after 3 s of combustion, as seen previously in Figure 4.2(a).


At half mass flow rate, a temperature of 1540 K is measured. The corresponding centerline
temperature across the reactor across the reactor is shown in Figure 6.13 indicating an
upper bound for the temperature drop due to heat losses is at most 500 C.
6.4.2

Final Combustion Test Results

During the final combustion test, unforeseen oxidation of titanium liners occurred, causing
an internal titanium fire to burn for over five seconds.Many temperature sensors situated

73

Figure 6.13: Hot Combustion Test 2: Temperature Drop Along the Centerline Across the
Reactor

directly in the flow path were damaged during the run and did not record data, though all
pressure transducers survived unscathed.Several sensors were already damaged in the prior
combustion run.
Pressure vs. time data is shown in Figure 6.14. The pressure profile follows the trend
seen in the prior successful combustor test (Figure 6.7) with one exception. When H2 /
O2 combustion is terminated at t = 18s,a pressure drop is expected because heat release
ends. In fact, the pressure begins to drop and subsequently rises up to the temperature
achieved during combustion.With no combustion after t = 18 s and no other work input
mechanisms available, only a significant source of heat release can cause such a pressure
rise. The pressure rise is sustained until the pebble bed heater steam is shut off at t = 25 s.

74

Figure 6.14: Titanium Fire: Combustor Exit Pressure (C3T) vs. Time

Some thermocouples acquired data during the early part of the run.Reactor exit temperature
is shown in Figure 6.15. As with the pressure profile, the temperature profile follows that of
the prior successful combustor test and is slightly hotter due the increased mass flow rate
of H2 / O2 . However, at t = 17.3 s, the sensor is damaged and data beyond this point can
be ignored.
Combustor liner temperature is shown in Figure 6.16 as measured by the optical temperature
sensor.A peak temperature of 1565 C is reached at t = 20.2 s followed by a gradual drop off
in temperature. Large sections of titanium and stainless steel liners were incinerated during
the run and areas of the liners were later found to be peeled away or torn off. Damage to
the sensors is believed to be due to molten debris entrained into the steam flow.

75

Figure 6.15: Titanium Fire: Reactor Exit Temperature (R5T) vs. Time

Figure 6.16: Titanium Fire: Combustor Liner vs. Time

76

6.4.3

Titanium Oxidation

Despite being uncommon, fires in bulk metals can occur when rapid surface heating occurs. Slow heating of a bulk metal typically results in the reaction being quenched by
large conductive heat losses.Higgins et al.[46] have examined bulk titanium burning in oxidative environments by making use of shock tube heating.Combustion is sustained by the
exothermic nature of the Ti and O2 reaction. Burning is distinguished from ablation by the
sustained luminosity on the surface of metal samples even when the flow temperatures drop
with time.
Figure 6.17 shows a comparison they obtained between the temperature of the shock heated
oxidizer (pure oxygen) and several the titanium characteristics such as melting point, boiling
point and adiabatic flame temperature. Combustion in titanium was shown to start occuring
when flow Mach numbers reached 6 and above, corresponding to a flow static temperature
above 2000 K.The melting point of titanium shown in Figure 6.17 is lower than this, only
1866 K.The reactivity of titanium was found to be the highest of the other metals examined
in the study. Its reactivity in the presence of oxygen is such that burning does not require
that titanium be volatilized. It experiences heterogeneous surface reaction when heated
beyond its melting point.The mechanism for this behavior is the low dissociation pressure
of titanium, which is 1016 atm at 1873 K (1600 C).[45]It explains titaniums affinity for
oxygen because it will decompose if the oxygen partial pressure is lower than the dissociation
pressure.
The titanium liners in the shock wave reactor were exposed to steam at a minimum of
1400K and Mach 0.24.Heat transfer rates are orders of magnitude lower than in the case
of Higgins et al. due to the significantly lower flow velocity. Nonetheless, titanium liners
rose in temperature, in this case up to a peak of 1564 C (1838 K) as seen in Figure 6.16,
meaning that the aforementioned reaction quenching by conduction into the cooler bulk
material was not in effect here. This measured liner temperature is effectively the melting
temperature of titanium to within measurement error. In the presence of pure oxygen, liner
burning appears highly probable but oxygen levels were kept low in the reactor by design.
Levels of oxygen available for oxidation of titanium were expected to be minimal in a

77

Figure 6.17: Titanium Fire: Combustion temperature (at constant pressure) for titanium
with oxygen in post-shock flow as a function of shock Mach number.

mixed bath of steam at nominal temperature of 2026 C (2300 K). As seen previously in
Figure 4.3(b), on a mass basis, expected levels are approximately 0.025% for dissociated
oxygen and 1.2% for molecular oxygen.In the absence of sufficient mixing as may occur
locally around each H2 / O2 injector and with locally high temperatures close to the adiabatic
flame temperature of H2 / O2 , the levels of oxygen may be higher.Near the adiabatic flame
temperature, levels of dissociated and molecular oxygen are estimated to be approximately
2.1% and 1.9% respectively.The reaction proceeds as:

2H2 O

4H + 2O

78

T i + 2O

T iO

As a consequence of the titanium liner fire, new liner materials were sought. a single ceramic
liner made of 2.54mm (0.1 inch) thick mullite (Al6 Si2 O3 ) is to replace double liners in the
combustion, dummy and mixing sections. Haynes 214

alloy is to replace single titanium

liners in the transition and mixing sections. Run durations are to remain unchanged. With
the heat capacity of mullite (1250 J/kg K) being far greater than that of titanium (400
kJ/kg K) it is anticipated that greater heat losses will be incurred. In addition, mullite
must be operated at a lower peak temperature (expected 1400C) than titanium, further
increasing heat losses.

79

Chapter 7
DESIGN METHODOLOGY AND CONSTRAINTS

The design of the supersonic nozzle and methane feedstock injectors is a complex fluid
dynamical problem and represents a significant portion of the analysis in this work. The
overall design is novel to the petrochemical industry, which is not a field accustomed to
compressible flow regimes. The design is also novel in the approach taken for mixing of
carrier steam with feedstock within this project. Whilst the first shock wave reactor project
utilized an array of interleaved two-dimenional nozzles to produce side-by-side acceleration
of both steam and feedstock followed by shear layer mixing, the new approach in the second
shock wave reactor project is contrained by the extreme operating temperatures of 3100 K
at commercial scale. Interleaved nozzles were shown to produce significant total pressure
losses in the first shock wave reactor and those that are capable of surviving such a 3100
K flow temperature would inherently cause intolerable heat losses. An abundance of design
analogies exist in the aerospace industry for this kind of high temperature supersonic nozzle
design, injection and mixing, especially in combustion chambers and rocket nozzles.These
have allowed the author to draw on a wealth of prior art for precursory design methodologies.
What follows is the set of specifications that governed the design process and steered it
toward a practical solution.
The shock wave reactor concept relies on the decoupling of two key processes, the first being
the supply of enthalpy to carrier steam and mixing with feedstock, and the second being the
intitiation of pyrolysis reactions. For pyrolysis to be delayed, steam and methane are mixed
at supersonic velocities in which the flow static temperature has dropped below pyrolysis
temperatures. Further, carrier steam is brought to supersonic speeds before feedstock injection takes place and this acceleration is achieved by a carefully selected De Laval nozzle
described in the Section 8. Also note that a reference to the feedstock injection section
refers to both the supersonic nozzle and the feedstock injectors. Flanges represent the start

80

and end of the section, as seen in cross section in Figure 7.1.

Figure 7.1: Cross-Section View: Feedstock Injection Section

The design of the supersonic nozzle is not entirely decoupled from the injection of feedstock
and the two sections have overlapping design criteria. A list of overarching design goals or
specifications that apply to both are presented below.
1. The total temperature of mixed steam and feedstock before the onset of pyrolysis is
to be 1500-1700 K;
2. Pressure losses are to be minimized to maintain supersonic flow and avoid premature
shock waves;
3. Throat material should remain under within structural and thermal limits and not
exceed 1800 K;
4. Heat loss is to be minimized;
5. Mixing : Thorough mixing CH4 and H2 0 is critical in order for sufficient carrier enthalpy to be available for pyrolysis and to ensure high temperature is reached for
effective pyrolysis of methane. The mixture is to be as thermally uniform and chemically homogeneous as possible.

81

Chapter 8
SUPERSONIC NOZZLE

In this chapter, rocket nozzle and supersonic wind tunnel design techniques are explained
followed by their use in determining the optimal supersonic nozzle shape. Then a set of
structural, thermal and fluid dynamical analyses are presented that confirm the suitability
of the selected design. Appendix E includes images of the final CAD design and fabricated
parts.
8.1

Design Criteria and Approach

The following incoming flow properties are assumed. Refer to Figure 4.3(b) in Section 4.1
for intensive flow properties from the upstream combustor section. A mass flow rate of
m
= 0.36kg/s is expected at a flow total temperature and total pressure of 2100 K and 2.1
bar respectively. These values are lower than those from the pebble bed heater on account
of a rough correction for thermal and total pressure losses occuring during combustion.
The correction is based on prior experience of the extent of the heat losses from the first
shock wave reactor project. Incoming flow is to have a Mach number of 0.24 in the 3.5 inch
diameter channel.
A calculated nozzle throat diameter of 2.1 inches chokes the 0.36 kg/s of incoming steam.
This length is obtained using the classic isentropic mass flow equation for perfect gasses
flowing in variable area ducts. The diameter,D is solved for implicitly:
p0
m
=
RT0 (1 +

M
+1

1
2 2(1)
2 M )

D2
4

Mt = 1; Mach number at the throat


= 1.19; Ratio of specific heats for steam at 2100 K
R = Cp ( 1
) = 467 J/Kg/K; gas constant for steam at 2100 K
p0 = 2.1; bar stagnation pressure upstream

(8.1.1)

82

T0 = 2100 K; stagnation temperature upstream

Outgoing flow properties are not known beforehand since they are influenced by the subsequent feedstock injector design requirements. However, total temperature and pressure
remain unchanged under the isentropic assumption. From available experimental hardware
capability, a known mass flow rate of CH4 of 0.081 kg/s can be supplied by the feedstock
injectors at total temperaturea and total pressure of 850 K and 1 MPa, respectively. Using
this and, from the analysis of Section 9, it was determined that a waveless axially uniform
supersonic flow at a Mach of 2.1 would be an acceptable outgoing flow for the supersonic
nozzle. In addition the nozzle exit diameter was chosen to be 3.0 inches rather than 3.5
inches to improve penetration of methane injected into carrier steam cross flow. Their
expansion ratio () is the ratio between the throat area and the exit area.
The design of a nozzle with specified incoming and outgoing boundary conditions is neccessarily an implicit or iterative one. Fluid flow analysis is, in general, a process of applying
boundary conditions and flow properties to a set of solvable governing equations or simplifications thereof. That is, boundary conditions and flow properties must be completely
specified first in order for solutions to be found. To the authors knowlege, no analytical
and only few numerical techniques exist that take only a partial set of flow boundary conditions as an input and parametrically alter the remainder of the set to achieve the desired
flow. Those numerical codes that exist are computationally expensive and their use here is
not feasible. The approach taken is therefore to design and analyze the supersonic nozzle
by keeping general supersonic flow principles in mind and iteratively test various nozzle
contours. These principles can be summed up by the following:

Long nozzles have a larger area in contact with the flow and have higher heat losses.
Their length allows gentle turning of supersonic flow and is able to produce an axial
waveless flow (that is, one without shock waves that cause pressure losses)

Shorter nozzles generate fewer heat losses but in turn, a waveless flow cannot be
produced.

83

Boundary layers grow as nozzle length increases, potentially leading to flow separation
and losses.

High enthalpy flows induce a peak temperature at the throat of the nozzle. A rounded
throat profile is desired as this peak temperature rises with a sharp edged throat.

Numerical flow analyses (computational fluid dynamics or CFD) are computationally expensive and thus time intensive. Analytical approaches are preferred to reduce
the number of iterations required.

Cost limitations dictate that ease of manufacturing be crucial.


When a satisfactory shape is found using these principles, design of the physical hardware
is performed to ensure the practicality of the solution. But first, a starting point for the
shape is required.
8.2

Rocket Nozzles

An aerospace designer turns to rocket nozzle design literature when seeking the shortest
possible axisymmetric nozzle that operates at high temperatures. Most rocket nozzles have
a converging - diverging De Laval type layout.Since gas velocity in converging section of
rocket nozzle is relatively low and subsonic, any smooth and well-rounded convergent nozzle
section will have very low losses. In contrast, the contour of the diverging section is crucial to performance, because of very high flow velocities involved. Selection of an optimum
rocket nozzle shape, for a given expansion area ratio, designers desire the following:
- Uniform, parallel, axial flow at the nozzle exit for maximum axial momentum, and
hence, thrust;
- No separation and turbulence losses within the nozzle ;
- The shortest possible nozzle length for minimum space envelope, weight, wall friction
losses and cooling requirements ;

84

- Ease of manufacturing .
The most basic contour that can be considered for a nozzle is a truncated cone. Conical
nozzle such as these yield close to uniform exit velocity but with a variation in flow angle
at the exit plane.i.e. there exists a radial component of flow velocity. Axial flow is desired
for maximum thrust, and thus a curved contour is better in that it is able to turn the flow
and minimize the radial component.
The most commonly utilized curved contour is the so called minimum length or bell nozzle.
Its profile was developed analytically by G.V.R. Rao at Rocketdyne to produce waveless
flow inside the divergent section of a rocket nozzle and axial exit flow with the minimum
possible length. An ideal nozzle, designed with the method of characteristics, produces the
same flow field but all properties are uniform across the exit plane. This is described in
Section 8.2.2 and often used in supersonic wind tunnel design. However, an ideal nozzle
is unnecessarily long for a rocket and simply adds extra weight. In contrast, a bell nozzle
is far shorter, but does not have uniform axial flow at the exit. Rather, the flow on the
centerline is faster and flow properties vary over the exit plane. In rockets, this variation is
of no concern as thrust is affected only by momentum leaving the nozzle at the exit plane.
Downstream of the exit plane, shock formation is induced due to this variation and it is
common to observe rocket plumes with shock-filled flow fields. In the shock wave reactor,
these shocks represent a pressure loss penalty, though weak shock waves posing minimal
losses can be tolerated if thermal losses are significantly reduced.
Bell nozzles are specified by their relative length to a conical nozzle with a 15 degree divergence half angle and the same expansion ratio. A fractional length Lf is given to specify
the equivalence. A bell nozzle and conical nozzle are shown in Figure 8.1

8.2.1

Parabolic Approximation of a Bell Nozzle

G.V.R Rao introduced an parabolic approximation for creating a bell nozzle of a given
length. Bell nozzles shorter than Lf = 0.8 do not contribute significantly to performance

85

Figure 8.1: Conical And Bell Nozzle Contour (Ref.[24])

in rockets. The Lf = 0.85 bell nozzle has been chosen, which corresponds to a length of
1.51 inches.The contour immediately upstream of the throat is a circular arc with a diameter of 1.5 Dt , where Dt is the throat diameter of 2.1 inches.This gives the nozzle a total
length of 2.58 inches. In the analytical bell nozzle design the divergent section begins with
a sharp edged expansion. However, it is well known that rocket nozzle throats experience
the greatest thermal stresses and peak temperatures and sharp curvatures are known to exacerbate this issue. Raos parabolic approximation replaces the share edged expansion with
a practical rounded and smooth circular arc posessing a diameter of 0.382 Dt and running
from throat T to the point N.The remaining curvature of the nozzle is approximated by a
parabola between point N and the exit E (fig. 8.2).

Design of the diverging parabolic section of a bell nozzle requires following input parameters:
throat diameter Dt , axial length of nozzle from throat to exit plane (Ln ) expansion area
ratio (ratio of exit area to throat area), initial wall angle of parabola n and nozzle-exit
wall angle e . Wall angles n and e have been chosen based on the empirical curves in
Fig. 8.3 where they are plotted as a function of expansion area ratio and fractional nozzle
length Lf .
The equations for the parabolic approximation are given below. Equation ( 8.2.1) is rear-

86

Figure 8.2: Parabolic Approximation Of Bell Nozzle Contour (Ref.[24])

ranged to specify the parabolic curve. Specifying initial and final angles as coefficients and
matching the first derivatives to ensure a smooth profile a set of three Equations ( 8.2.2,
8.2.3 and 8.2.4) are obtained:
p
b2 4a(c z)
z = aR + bR + c R =
2a
1
b2 4a(c ze ) =
tan2 e
1
b2 4a(c zn ) =
tan2 n
b

ze zn = LP

(8.2.1)
(8.2.2)
(8.2.3)
(8.2.4)

This set must be solved in 5 unknowns (a, b, c, zn and ze ). To complete the system, two
more equations are required. Beginning with the expression for R at zn and at ze , replace
R with 8.2.1 and substitute the quantity under square root with 8.2.3 and 8.2.2.

p
b +
R(ze ) = 5.67 108 Js1 m2 K 4 R =

R(zn ) = R =

b +

q
p
1
b
+
2
b 4a(c ze )
tan2 e
=
2a
2a
(8.2.5)
q

p
b + tan12 n
b2 4a(c zn )
=
2a
2a

(8.2.6)

87

Rearranging equation 8.2.5 and 8.2.6 and denoting kn


8

5.67 10

Js

1
tan2 n

and ke

1
:
tan2 e

( ke b)2
R =
4a2

( kn b)2
2
(R ) =
4a2

(8.2.7)
(8.2.8)

With give equations ( 8.2.2, 8.2.3, 8.2.4, 8.2.7 and 8.2.8) with five unknowns (a, b, c, zn
and ze ), MATLAB is used to solve the system. The results are that n = 19 and e = 12 .
As mentioned, the total length is only 3 inches, an excellent result when heat losses are a
concern. The final nozzle profile is shown in Figure 8.4

Figure 8.3: n and e as Function of Expansion Area Ratio (Ref.[24])

88

Figure 8.4: Approximation of a Minimum Length Bell nozzle

8.2.2

Method of Characteristics

A second nozzle shape can be obtained with method of characteristics. This profile is long
but the flow exit conditions are entirely uniform and axisymmetric. The method of characteristics is a technique to solve partial differential equations. Considering the first order
linear partial differential equation in two variables (x,t), the goal of the method, when applied to this equation, is to change coordinates from (x,y) to a new coordinate system (x0 ,s)
in which the PDE becomes an ordinary differential equation (ODE) along certain curves in
the x-y plane. Such curves along which the solution of the PDE reduces to an ODE, are

89

called the characteristic curves or just the characteristics. The new variable s will vary, and
the new variable x0 will be constant along the characteristics.

The application of the method to axisymmetric nozzle geometries results in a set of differential equations to be solved along characteristic lines, instead of the simpler algebraic
equations for a planar geometry. With the development time involved in writing a numerical solver code is substantial. To avoid this, a graphical approach has been taken. Since a
starting point for the design process is all that is required, the gas dynamics reference by
Zucrow and Hoffman [51] provides a set of ideal nozzle curves for various area ratios that
can be traced graphically. For the given expansion area ratio of  = 2.04, the length of the
an ideal nozzle diverging section is 8 inches.
The corresponding ideal nozzle and bell nozzle configurations were imported into Microsoft
Excel to act as bounds for the design. Ten intermediate curves were generated by trial and
error (some are shown in Figure 8.5). Their shapes were based on the idea of an expansion
with initially high positive curvature generating expansion fans followed by an inflection
point and then a gradual negative curvature, inducing compression waves that cancel the
expansions exactly. These designs were analyzed in a computational fluid dynamics software
package called FLUENT as described in Section 8.3.2

Figure 8.5: Comparison Between Different Nozzle Contours.

90

8.3

Nozzle Analysis

In this chapter, several analyses are presented in which the nozzle contour is optimized for
Mach 2.1 flow at the exit plane, minimal shock waves, thermal loads, heat losses and finally
structural loads.
8.3.1

Quasi 1-D flow field model.

Each of the proceeding analyses require knowlege of the flow field internal to the nozzle.
This can be modeled using computational fluid dynamics (CFD), though computational
expense is high. Instead, a quasi-one-dimensional algebraic model has the capability to
better simulate the flow in a convergent-divergent nozzle when rapid testing and interation
is required.
Code was developed in MATLAB for this purpose (see Appendix D) and assumes isentropic
flow of a perfect gas through the nozzle. The inputs for this program were the nozzle
contour (radius vs. distance) and incoming steam flow properties like (m,
T0 , P0 , Mach, Cp ,
) that were obtained from GasEq in Section4.1. Static entrance conditions are obtained
via the isentropic flow equations based on Mach number and stagnation quanities. The
laws of mixing in calorically perfect gasses allow the calculation of the ratio of specific heat
capacities, .

pi+1 = pi (

(1 +

Ti+1 = Ti (

i+1 = i (

1
2

2 Mi
) 1
1
2
2 Mi+1 )

(8.3.1)

1
2
2 Mi
)
1
2 )
M
i+1
2

(8.3.2)

1
2
1
2 Mi
) 1
1
2
2 Mi+1 )

(8.3.3)

1+

1+
(1 +
1+

(1 +

Equation 8.1.1,is used to estimate the Mach number at the nozzle inlet (Minlet = 0.22 ).
Combustion tests in Section 6.4 revealed that the measured Mach number is in fact 0.24.
From a known geometry, the ratio

Ai
At

is known where Ai is the area at an arbirtrary location

and At is the throat area. Imposing mass conservation in combination with Equation 8.1.1

91

and Mt = 1, the Mach number at an arbitrary location can be solved for from the resulting
Equation 8.3.4:
+1

Mi =

(1 +

1
2 2(1)
2 Mi )

Ai /At

(8.3.4)

As an example, the optimal nozzle contour mentioned in Section 8.2.2 is analysed using the
quasi-one-dimensional code. The nozzle is a total of 6 inches long and its curve is shown in
Figure 8.6.

Figure 8.6: Optimal Nozzle Profile

Trends of Mach number, pressure, temperature, density and momentump flux are plotted
in Figures 8.7, 8.8, 8.9 8.10 and 8.11 respectively.

92

Figure 8.7: Optimal Nozzle: Mach Number vs. Distance

8.3.2

Computational Fluid Dynamical Simulation in Fluent

To best choose the optimal nozzle profile a computational fluid dynamics (CFD) model was
created in ANSYS Fluent software. As mentioned previously in Section 8.2.2 , ten shapes
were tested.The focus of the analysis was the nozzle exit plane flow. A Mach 2.1 flow on the
centerline was a firm requirement, though marginally lower off center flow Mach numbers
were accepted. Resonably uniform flow properties were acceptable, though these are directly
related to the flow Mach number. Of critical importance was to reduce any unnecessary
flow losses incurred in the presense of shock waves or strong compression waves.
Meshing Solid Works solid modeling computer aided design software was used to construct

93

Figure 8.8: Optimal Nozzle: Pressure vs. Distance

the nozzle geometry for import into ANSYS Fluent. Subsequent meshing was performed
in Fluents built-in meshing package. Computer hardware was limited to grid sizes of
below 600 000 nodes. An coarse unstructured grid with spherical refined areas (called
spheres of influence in Fluent) was placed in the area of the expected inviscid flow field.
Refinements were placed in areas where high pressure gradients were expected. A finer
structured boundary layer mesh was added to model the viscous region.
Solver Models Two different treatments of the continuity, momentum and energy equations
are available, a pressure or a density based solver. Density-based solvers give superior results

94

Figure 8.9: Optimal Nozzle: Temperature vs. Distance

in high velocity and compressible flows without significant low velocity regions so this is
chosen.
The presence of turbulent flow in the supersonic nozzle is certain due to the upstream
turbulent flow from H2 / O2 combustion as well as the development of a relatively thick
turbulent boundary layer. Flow Reynolds numbers (Rex ) are expected to be at least 3.8
105 and since the transition from laminar to turbulent flow in flat plates occurs between
105 and106 , this transition is expected.
Turbulence is not a well understood phenomenon, though it could be thought of as the onset
of instability in laminar flow that occurs at high Reynolds numbers (Re). Fluid viscosity is

95

Figure 8.10: Optimal Nozzle: Density vs. Distance

able to damp out such instability at low Re, but at a critical Reynolds number, a transition
occurs, instabilities grow and flow becomes turbulent. The reference to such instabilities is
simply a way of describing the interactions between non-linear inertial terms and viscous
terms in within the Navier Stokes equations. Increasing instability suggests that inertial
terms dominate viscous terms.
Turbulence is characterized by vorticity in a flow field, the transport of eddies, the transport, diffusion and dissipation of turbulent kinetic energy and transfer of energy from one
form to another. The presence of turbulence generates entropy and represents a mechanism for flow pressure losses. Turbulent eddies are rotational, fully time-dependent and

96

Figure 8.11: Optimal Nozzle: Momentum Flux vs. Distance

fully three-dimensional. Furthermore turbulence is thought of as random process in time.


Therefore no deterministic approach is possible. Certain properties could be learned about
turbulence using statistical methods. These introduce certain correlation functions among
flow variables. However it is impossible to determine these correlations in advance.
Another important feature of a turbulent flow is that vortex structures are convected along
with the flow. Their lifetime is usually very long. Hence certain turbulent quantities can
not be specified as local. This simply means that upstream history of the flow is also
important of great importance. Nowadays turbulent flows may be computed using several
different approaches. Either by solving the Reynolds-averaged Navier-Stokes equations with

97

suitable models for turbulent quantities or by computing them directly. The objective
of the turbulence models for the RANS equations is to compute the Reynolds stresses,
which can be done by three main categories of RANS-based turbulence models: linear eddy
viscosity models, non-linear eddy viscosity models and Reynolds stress model (RSM). The
linear eddy viscosity approach involves turbulence models in which the Reynolds stresses,
as obtained from a Reynolds averaging of the Navier-Stokes equations, are modelled by a
linear constitutive relationship.
There are several subcategories for the linear eddy-viscosity models, depending on the number of (transport) equations solved for to compute the eddy viscosity coefficient: algebraic
models, one equation models and two equation models.
Zero Equation Models Algebraic turbulence models or zero-equation turbulence models are
models that do not require the solution of any additional differential equations, and are
calculated directly from the flow variables. As a consequence, zero equation models may
not be able to properly account for history effects on the turbulence, such as convection
and diffusion of turbulent energy. In general, these models are inaccurate but can be quite
useful for a simpler class of flow geometries or in start-up situations.
One Equation Models One equation turbulence models solve one turbulent transport equation, usually for the turbulent kinetic energy. Three have become popular, namely: The
Prandtl, Baldwin-Barth and Spalart-Allmaras models. Their solution is faster than two
equation models and provide a level of accuracy that most diligent designers may find inadequate.
Two Equation Models The two equation models provide a better degree of accuracy and are
the type most commonly used in CFD solvers. The likes of the k-model and the k-omega
model have become industry standard models and are typically used for in general classes
of engineering problems. Two equation turbulence models are also very much still an active
area of research and new refined two-equation models are still being developed.
By definition, two equation models include two extra transport equations to represent the
turbulent properties of the flow. This allows a two equation model to account for history
effects like convection and diffusion of turbulent energy.
Most often one of the transported variables is the turbulent kinetic energy, k. The second

98

transported variable varies depending on what type of two-equation model it is. Common
choices are the turbulent dissipation, , or the specific dissipation, . The second variable
can be thought of as the variable that determines the scale of the turbulence (length-scale
or time-scale), whereas the first variable, k, determines the energy in the turbulence.
There are two major formulations of K-models: standard k-model, realisable k-model
and RNG k-model. In the Fluent analysis of the supersonic nozzle profiles, a standard
k-model was selected due to its better treatement of high Reynolds number flows such as
those expected inside each nozzle. Default values for kinetic energy (k) and turbulent dissipation rate (5.67 108 Js1 m2 K 4 ) were kept because, despite changing their values,
no noticable changes in flow were noted

In fact, the simplest complete models of turbulence are the two-equation models in which
the solution of two separate transport equations allows the turbulent velocity and length
scales to be independently determined. The standard k- 5.67 108 Js1 m2 K 4 model in
ANSYS FLUENT falls within this class of models and has become the workhorse of practical engineering flow calculations in the time since it was proposed by Launder and Spalding
[9]. Robustness, economy, and reasonable accuracy for a wide range of turbulent flows explain its popularity in industrial flow and heat transfer simulations. It is a semi-empirical
model, and the derivation of the model equations relies on phenomenological considerations and empiricism. This semi-empirical model is based on model transport equations
for the turbulence kinetic energy (k) and its dissipation rate (5.67 108 Js1 m2 K 4 ).
The model transport equation for k is derived from the exact equation, while the model
transport equation for 5.67 108 Js1 m2 K 4 was obtained using physical reasoning and
bears little resemblance to its mathematically exact counterpart.

A fluid material model that is appropriate for modeling the state changes within high temperature supersonic steam was selected. The specific heat (CP ) is allowed to change with
temperature according to two part polynomial fits that use internal Fluent gas dynamic
databases. Pressure, temperature and density are related by the ideal gas law. Following Sutherlands law, viscosity changes with temperature. Sutherlands viscosity law was

99

derived directly from kinetic theory by Sutherland (1893) using an idealized intermolecularforce potential. The formula in Fluent is specified using either two or three coefficients,
though the two coefficient equation (below) was selected
3

C1 T 2
T + C2

Boundary Conditions Typical boundary conditions for internal compressible flows are pressure inlet and pressure outlet. It should be noted that at the time of simulation, the
stagnation pressure of steam entering the nozzle was known to be influenced by the capabilities of the pebble bed heater, which was to run at maximum capacity. This supply
pressure was expected to be 2.5 bar and hence, this value was used at the pressure inlet.
The pressure outlet required a pressure condition that would drive the flow to supersonic
and result in Mach 2.1 flow. Since total pressure was expected to remain fairly constant
(from prior discussion), this static pressure corresponds to 28.4 kPa using an isentropic
assumption. Pressure outlet conditions are loosely defined and do not pin the solution to
the value supplied, but rather drive the solution such that the averaged boundary pressure
is close.
Pressure Inlet Boundary Condition :
- Stagnation pressure 250 kPa,
- Static pressure 244180 Pa (evaluated by the quasi-one-dimensional code for a known
Mach of 0.22),
- Total Temperature 2100 K .
Pressure Outlet Boundary Condition :
- Static pressure 28.4 kPa,
- Total back flow temperature 1500 K.

100

Solver Type Even though the conditions inside the reactor are transient, it is the design
condition that is of interest and this is quasi-steady state. A steady state explicit solver
is thus employed. Explicit solvers demonstrated better stability in this case. Between the
discretization schemes available the first order upwind method proved stable and sufficiently
accurate for the solution of the momentum equation.
The final consideration is choice of Courant number since in Fluent, a transient term is
included in the coupled sovler even for steady state problems. For ANSYS FLUENTs
density-based solver, the main control over the iterative stepping scheme is the Courant
number, the step is proportional to the Courant number as follow.

t =

2( Courant number )V
P max
Af
f f

is the maximum of the local eigenvalues


V is the cell volume, Af is the face area and max
f
defined in the Fluent user manual.
Linear stability theory determines a range of permissible values for the Courant number (
the range of values for which a given numerical scheme will remain stable). In general, taking
larger time steps leads to faster convergence, so it is advantageous to set the Courant number
as large as possible (within the permissible range). Linear stability analysis shows that the
maximum allowable Courant number for the multi-stage scheme used in the density-based
explicit formulation will depend on the number of stages used and how often the dissipation
and viscous terms are updated . But in general, you can assume that the multi-stage scheme
is stable for Courant numbers up to 2.5. This stability limit is often lower in practice
because of non-linearities in the governing equations. The stability limits of the densitybased implicit and explicit formulations are significantly different. The explicit formulation
has a more limited range and requires lower Courant number settings than does the densitybased implicit formulation.Since an explicit formulation is used, the Courant number is set
equal to 0.95
With these settings the simulation ran until the convergence; in terms of energy and continuity residuals of below 106 were achieved .
A simulation was also run on the bell nozzle in order to verify that with our optimized

101

contour shape, oblique shocks at the exit are going to be weaker and less total pressure
losses will be present than for the bell contour. Results are shown in fig. 8.12 and fig. 8.13
for comparison.
Looking at results we noticed that, as predicted, pretty strong oblique shocks were coming
from the diverging section of the bell model while, in the optimized shape one, much more
weaker shocks develop . This happens because our shape has not the ideal length, a shorter
shape led inevitably to a curvature at the exit thats a bit still too sharp.

8.3.3

Turbulent Compressible Boundary Layer

The growth of a boundary layer inside rocket nozzles and the supersonic nozzle in the shock
wave reactor is a direct result of friction (or viscous) created by the nozzle wall on the gas flow
within them. These effects are very closely modelled by the governing differential equations
of fluid flow, the Navier-Stokes equations, though their solution is not known in general.
A great many mathematical techniques for their solution rely on simplifying assumptions,
especially, that of assuming that the fluid, being a gas, has a very small viscosity that can
be neglected. Ludwig Prandtl showed that this is valid in free stream conditions far from
fluid boundaries, but even for a vanishingly small viscosity, there always exists a thin layer
of fluid in which viscous effects cannot be neglected.
This is known as the boundary layer and its effects in rocket nozzles and have been closely
studied due to their negative impact on performance and especially their role in heat transfer. Heat transfer effects and analysis receive further treatment in Section 8.3.4. This
seciton adresses performance effects from the point of view of a rocket nozzle designer and
how these can adversely affect the shock wave reactor nozzle.
Boundary layers inside rocket nozzles affect performance in in three primary ways:

The first, the presence of the boundary layer slightly alters the free stream characteristics.
By definition, the boundary layer displacement thickness is that thickness of the free stream
flow that is apprently lost owing to the velocity deficit within the boundary layer. Hence

102

Figure 8.12: Comparison of Mach Profile In A Bell Model (top) and in the Optimized Profile
(bottom).

103

Figure 8.13: Comparison Of Total Pressure Profile In A Bell Model (top) and in the Optimized Profile (bottom).

104

the free stream is in effect displaced from the wall by this thickness. At the throat the result
is a slight reduction of the throat area At and hence of mass flow rate. The displacement
thickness will also affect the distribution of A/At through the nozzle, and this in turn will
affect the pressure distribution and rocket thrust. For large-diameter nozzles, this effect
is typically small. During rocket nozzle analysis, so called boundary layer corrections
are necessary to account for displacement of the largely inviscid free stream and the corresponding area changes. This is applied in the shock wave reactor nozzle analysis that
follows below.
Second, the boundary layer affects the thrust of the rocket through shear stress or skin
friction on the nozzle wall. This effect is not a concern in the shock wave reactor since
thrust is of no interest.
The third influence of the boundary layer concerns the interactions between it and shock
waves. The nozzle in the shock wave reactor is designed to have no shock waves inside its
diverging (supersonic) section, and so this influence is also less of a concern inside the nozzle.
Downstream of the nozzle, flow is expected to contain a great many three dimensional shock
structures due to feedstock injection. In addition, studies by Masse et al. [29] undertaken
in conjunction with the first shock wave reactor project focussed on these effects to better
understand how they induce pressure recovery and add to shock losses. These effects will
not be considered in this section however.
The aforementioned boundary layer corrections will now be developed using a semi-empirical
turbulent compressible boundary layer treatment for axi-symmetric flow geometries. Turbulent treatments apply for high Reynolds number flows, where Rex = f racU x, U is fluid
free stream velocity, x is a characteristic length scale and is kinematic viscosity. Typically
in flat plates, turbulent boundary layer growth is predicted by Equation 8.3.5. Here x is
the distance from the edge of the plate.


U x
= 0.37x

 1
5

(8.3.5)

The derivation makes use of the classic observation that turbulent boundary layer velocity
 1
7
profiles are proportional to to y , where y is the normal distance from plate surface and

105

is boundary layer thickness. Note that turbulent boundarys grow faster than the laminar
boundary layers modeled with the Blasius solution (see Figure 8.14).

Figure 8.14: Laminar and Turbulent Boundary Layer Growth in Flat Plates Ref.[42]

The simplicity of Equation 8.3.5 is undermined by its underprediction of boundary layer


development, and moreover, the equation can verily be used only as a first approximation in
axisymmetic flows. Turning to a literature review [9] that has compared several well known
turbulent compressible boundary layer modeling approaches, an appropriate formula with
a greater reliability is found. For Reynolds numbers of the order of 106 this approximation
has been made using Spences method [15] for insulated walls.

X = P 1

P dx
0

106

P = M 4 (1 +

M 2 3.343
)
(1 + 0.128M 2 )0.822
5

For axisymmetric geometry, P is replaced by P r where r is the radial distance from the
axis and is parameter set equal to

RX =

5
4

when ReX is about 106 .

UX
=




M 2 3
a0
XM 1 +
0
5

(8.3.6)

1
2
= 0.37XRX

(8.3.7)

Where X is an equivalent flat plane length, P is a function of local Mach number (M) as
before, and suffix 0 refers to stagnation conditions. The quantity is the exponent in the
viscosity-temperature relation and T , with reasonable value for being 0.75, excepting
at very high temperatures where = 0.5 is more accurate.

In fact, the results from both flate plate model or the axisymmetric model are in fair
agreement as in Figure 8.15. The former model (Equation 8.3.5) is traced in green and
latter model (Equation 8.3.7) in red.

The boundary thickness at the exit plane of the nozzle is found to be 0.189 inches, so this
represents a 12.6 % reduction in exit area taken up by the boundary layer. Subsequent CFD
simulations confirmed the accuracy of this semi-empirical result as shown in Figure 8.16.
The results of the simulation give an exit plane thickness of 0.192 inches.
8.3.4

Thermal analysis

Since extremely high temperature steam flow is expected in the nozzle and because heat
losses are of prime concern, a thermal analysis was conducted. This type of analysis is
common in rocket nozzles and, aside from highly simplified models, analytical treatments
of the problem are not tractable. The approach taken in this work was numerical and saw
the development of a Matlab code able to estimate heat losses through the walls and using
a commercial ANSYS finite element software package.

107

Figure 8.15: Calculated Turbulent Boundary Layer Development in the Supersonic Nozzle.

The results of the analysis were used in an interative design process to determine two
competing design parameters. These were nozzle wall thickness and the material used
in construction. Thinner nozzle walls have low capacity for storing heat, and this low
thermal mass is a direct consequence of their lower mass. Low thermal mass materials heat
up quicker during transient heat transfer and this is desirable in the shock wave reactor
because a hotter nozzle wall causes less heat loss. On the other hand, thinner nozzle walls
are structurally weaker and when subjected to aerodynamic loading from steam flow they
may be susciptible to failure. A structural analysis is later presented in Section 8.3.6.
Material choice had a vast effect on the heat transfer rates, peak temperatures and thresholds
in the nozzle design. Refractory metals like titanium, Haynes 214 and stainless steel as well
as ceramics such as alumina, zirconia and mullite were considered. Material properties of

108

Figure 8.16: CFD Results of Turbulent Boundary Layer Development in the Supersonic
Nozzle.

interest were primarily the melting temperature (Tmelt ), the thermal conductivity (k), the
specific heat capacity at constant pressure (Cp ) a combination of the two called the thermal
diffusivity ( =

k
Cp ).

The quantity Cp can be thought of as a volumetric heat capacity for

a material with density . The thermal diffusivity is a measure of the ability of the material
to conduct heat relative to its ability to store it and hence its response time to changes in
temperature at its boundaries.
Materials with high melting temperatures are desirable in that a higher operating temperature can be achieved without structural failure. Low thermal condutivity is beneficial as the
rate of heat transfer is lower and heat losses are reduced. A material with low specific heat
capacity is also beneficial since such a material will heat up faster. Thermal diffusivities
are a secondary consideration. Sufficiently high thermal diffusivity is required to heat the
nozzle up to operating temperature, though not too high as a sharper temperature rise will
occur and has the possibilty of inducing material warping or melting. Ease of manufacture, strength and price were also factors of influence in the design and often trumped the

109

aforementioned material properties in cases where they proved overly restrictive.


In terms of multi-layered materials, a thermal barrier coating may also be beneficial. These
form a layer of thermal protection because have low thermal diffusivities but tolerate temperature extremes beyond those of the underlying structural material. Ceramic spray coatings
are commonly used to keep turbine blades cool in air breathing turbojet or turbofan engines. In nozzle design, the throat region is known to see peak temperatures due to sonic
flow, small radius of curvature and hence higher convective heat transfer. This region would
benefit most from such a barrier coating but, in the final analysis, a coating was not deemed
necessary.
The numerical modeling technique developed in MATLAB is presented first, followed by
the results of the design process. The heat equation in polar coordinates (radial r, axial z)
is solved on a curvilinear grid. The governing differential equation represents conservation
of energy for an axisymmetric solid geometry and are shown below in Equation 8.3.8.




1
T

T
T
kr
+
k
= cP
r r
r
z
z
t

(8.3.8)

Assuming a homogeneous material with a constant k simplifies this to:


1 T
2T
2T
1 T
+ 2 + 2 =
r r
r
z
t
Again, is the thermal diffusivity and is equal to

k
CP

(8.3.9)

A finite difference approximation applied to this equation is obtained for the first and second
partial derivatives in Equation 8.3.9.
Time discretization is a second order Adams-Bashforth scheme. Fourth order Runge-Kutta
methods are more accurate, and allow larger time steps to be taken for the same number of
iterations, but in this case the explicit scheme can easily become unstable for large time steps
so little gain was made from using this scheme. Additionally Runge-Kutta time stepping
was avoided due to the added complexity of evaluating at half time steps.:
n+1
n
Ti,j
Ti,j
1 T n T n1
(3

)=
2
t
t
t

(8.3.10)

110

Spatial discretizations are second order accurate central differences.


Ti+1,j Ti1,j
T
=
r
2r

(8.3.11)

Ti1,j 2Ti,j + Ti+1,j


2T
=
2
r
(r)2

(8.3.12)

Ti,j1 2Ti,j + Ti,j+1


2T
=
z 2
(z)2

(8.3.13)

Substitution of Equations 8.3.10, 8.3.11, 8.3.12 and 8.3.13 into Equation 8.3.9 and rearranging then reduces to an explicit time stepping scheme.
In general, a curvilinear grid requires a coordinate transformation to correctly approximate
derivatives in r and z. However, the grid is selected such that curvilinear distortion is only
present in the axial direction, since grid points are arranged with a constant r. Only 2nd
derivatives in z require treatment.
The transformation required is:
2T
2T
=
z 2
r2

r
z

2
+

T 2 r
r z 2

(8.3.14)

Two more discretizations related to the grid are thus required:


ri,j+1 ri,j1
r
=
z
2z

(8.3.15)

ri,j1 2ri,j + ri,j+1


2r
=
z 2
(z)2

(8.3.16)

Grid
The grid is shown in Figure 8.17. Grid spacing is constant in the axial direction, with
z = 1.3mm. Grid spacing in the radial direction varies as the grid warps in space, though
the maximum rmax = 6.0mm.
Boundary Conditions
Four computational grid boundaries exist, two curvilinear and two cartesian. The primary
mechanism for transferring heat into the nozzle is from the main flow by convection with a
negligiblly small amount of radiation. This heat enters through the lower curvilinear wall

111

Figure 8.17: Curvilinear Computational Grid for Simulating Nozzle Heat Transfer

seen in Figure 8.17. On the opposite curvilinear wall, only radiation to the atmosphere at
298 K is significant.At the two straight end walls, an adiabatic condition is imposed.
We first consider conduction at an arbitrary point in any of the solid nozzle walls. The rate
of conductive heat transfer through a differential area dAr at radius (r) is given by Fouriers
law ( 8.3.17):
kT
qr
=
dAr
r

(8.3.17)

The area dAr = 2rdl in a zeroth order approximation. More accurately, a curved wall
segement, when discretized by a computational grid, is effectively the segment of a cone.
Consider a cone placed with its axis of revolution collinear with the axial z axis. A sliced
cone segement represents a small area of the nozzle wall. It starts at axial distance zi and
end at zi+1 . The outer edge of the cone begins at radius Ri and ends at Ri+1 . A straight
line joining the two points (zi ,Ri ) and (zi+1 ,Ri+1 ) and has an equation R(z) = mz + c.

112

Integration of the cone area in the segment yields:

1
dAr = 2( m(z22 z12 ) + c(z2 z1 ))and
2

m
c

qr = 2krdl

z1 1
z2 1

R1
R2

T
r

(8.3.18)

(8.3.19)

Second, consider convection through a generic wall i .

qi = 2ri hg (Tw1 T,g )

(8.3.20)

- Tw1 is wall temperature


- T,g is main flow bulk temperature
- hg is the heat transfer coefficient

Bartz developed Equation ( 8.3.21) to estimate the heat transfer coefficient hg in supersonic
compressible flow at high temperatures. The application was rocket nozzle cooling. The
equation can be used here by making use of the flow properties estimated by the pseudo
1-D MATLAB code from Section 8.3.1
 0.8  0.1  0.9

p0
Dt
0.026 0.2 Cp
At

hg =
0.2
0.6

Pr
c
rc
A
Dt


Where subscript

represent stagnation conditions and t denotes throat dimensions:

- Pr is Prandtl number
- c =

p 0 At gc
,

(m)

(8.3.21)

the characteristic velocity

- rc is throat radius of curvature in a plane which contains the nozzle axis.

i = 


1 Twh
2 T0g 1 +

- Twh = Tw1
- Mi local Mach number
- local ratio of specific heats

1
2
2 Mi

1
0.80.2 
+ 12
1+

1
2
2 Mi

0.2

113

- = constant T Viscosity-Temperature Relation

Third, using the Stefan-Boltzmann law for a grey body, that calculate the total energy
radiated per unit surface area of a black body in unit time we have:

qrad =

qr
= Tw4 Ts4
dAr

(8.3.22)

Where is Stefan-Boltzmann constant and is equal to 5.67 108 Js1 m2 K 4 , material


emissivity of the grey body, is absorptivity of the wall and Ts is room temperature .

Results
In Figure 8.18 below, a transient simulation running with 10 seconds of 1400 K steam and
5 seconds of combustion augmented steam at 2100 K is shown. Walls are made of stainless
steel 316 and a minimum of 0.9 inches thick

Figure 8.18: Temperature Distribution In Nozzle Walls.

114

As expected, peak temperature is found at near the throat and is in the region of 1287 K
(1014 C), and well below Tmelt . Other simulations with ceramics materials such as the
derivatives of zirconia or alumina were also performed. Figure 8.19 demonstrates the ability
of refractory ceramics (if practical) to reach higher peak temperatures than titanium or
stainless steel.

Figure 8.19: Variation Of Temperature Distribution In Nozzle Walls During Experiment


Time.

Despite this excellent temperature rise and ability also to withstand this extreme condition
without failure, most ceramics are prohibitively expensive to manufacture complex shapes
for internal flow devices like nozzles. The key benefit of a refractory ceramic material is its
high melting temperature varying bewteen 1950 C for more common alumina and 2400 C
for zirconia. However, manufacture often requires casting to produce a complex nozzle

115

shape and the result is a porous surface. This can be eroded by steam and makes their use
less practical. Non-porous Mullite can be acquired but a disadvantage is the nozzle shape
in this project is not readily machinable. The use of steam is also a concern as it has more
tendency to damage porous walls and the high thermal conductivity of steam compared to
gasses like air can induce very sudden thermal shock in a ceramic nozzle wall that results
in thermal stresses great enough for cracks to form.
Titanium, on the other hand, does not suffer from the brittleness of ceramics and it is
not susceptible to thermal shock. It is far more readily machined into nozzle shapes than
ceramics, though this is by no means an easy task. Raw titanium bar stock used to fabricate
such a nozzle is relatively expensive compared with less exotic alternatives like stainless steel.
Its performance is clearly better since it reaches a higher operating temperature and is able
to withstand temperatures up to 1950 K before melting.
The concern with stainless steel was the low melting temperature of 1780K. To keep temperatures below this limit, a thick nozzle wall is required which decreases the rate at which the
nozzle wall temperature rises and is detrimental to heat losses. After several simulations,
however, the sensitivity of heat losses to this temperature was were found to be lower than
anticipated, primarily due to the short length.
Despite the lower peak operating temperatures achievable in stainless steel, losses are affected by only 10 to 20 K degrees when compared with titanium. The workability and ease
of machining then became significant factors in its choice. The supersonic nozzle has been
fabricated on a CNC lathe and is shown in Appendix E: Figure E.4.

116

8.3.5

ANSYS Heat Transfer Validation

To validate results from the MATLAB thermal simulation above, a commercial finite element
software package, ANSYS, was utilized. Geometry was created in Solid Works and imported
into ANSYS for meshing with tetrahedral solid elements (Fig. 8.20) resulting in a total of
93448 elements.

Figure 8.20: Finite Element Tetrahedral Element Mesh.

A thermal analysis was performed using stainless steel material properties and identical
boundary conditions to those used in the MATLAB code. Although the convective heat
transfer coefficient hg is set equal to a constant 1300

W
m2 K

to be compatible with the software,

this value is known to vary along the length of the nozzle. Its dependence on flow properties
such as Mach number can be obtained from Bartzs equation ( 8.3.21) and it is seen to vary
from 600

W
m2 K

up to 1450

W
.
m2 K

The results of the analysis show that the temperature at the throat of the nozzle will be
1002 C (Figure 8.21) . A difference of 1.4% exists between the results of MATLAB and

117

ANSYS due to the different in the parameters used in each analysis

Figure 8.21: Variation of Temperature in Nozzle Walls from ANSYS.

8.3.6

Structural Pressure Loading

A structural analysis was performed to determine aerodynamic loading by the steam flow
at the nozzle wall. The loading is primarily due to pressure forces, and hence shear forces
are neglected. Pressure predictions along the length of the nozzle are obtained from the
quasi one-dimensional code described previously in Section 8.3.1. Integrating these pressures along the surface of the nozzle reveals the distribution of forces on the internal wall.
Figure 8.9 is a force vector plot of this result.
In Figure 8.9 the longest arrow represent a force of 257N while the shortest is 28N . These
forces are trivial for the chosen solid stainless steel 304 nozzle.

118

Figure 8.22: Forces Acting on the Internal Nozzle Walls due to Steam Flow Pressures

119

Chapter 9
METHANE INJECTOR DESIGN

A critical area of design in the second shock wave reactor project has been that of feedstock
injection because injection of any kind of mass flow into a supersonic steam cross flow results
in a severe formation of shock waves, separation zones, high temperature stagnation regions.
Due to the temperatures involved, placing stirring devices in the flow is not practical and
thus this injection also needs to provide the primary mechanism for mixing. Keeping the
high operating temperature goals of pyrolysis in mind, the injector design must also be such
that:
- Feedstock must penetrate the cross flow and at least reach the duct centerline ;
- The jet cloud must cover the maximum cross sectional area in the duct ;
- To achieve complete thermal and molecular mixing in the shortest length (reducing
heat losses) and within millisecond timescales;
- Maintain supersonic flow at the end of injection section (M 1.5) to delay pyrolysis.
Stated another way, this involves minimizing stagnation pressure losses since stagnation pressure can be thought of as a driving force behind the flow ;
- Produce a mixed static temperature around 1450-1550 K such that sufficient reaction
enthalpy is availble ;
Some geometric and facility restrictions are:
- Injection into a diameter of 3.5 ;
- Feedstock injection pressure maximum of 150 psi;

120

Several of the goals of are quite restrictive from a design perspective. The short feedstock
injection section length is required to minimize heat losses and provide the required exit
temperature. A system of liners were installed in other sections to reduces heat losses, but
because of the high complexity of the flow field and temperature extremes in the supersonicnozzle and injection regions, liners cannot be utilized. The short feedstock injection section
length also imposes the constraint that mixing be achieved in a very short distance. Here,
the supersonic flow velocities are in the region of 1300 m/s and flow must then be fully
mixed in the time it takes for the mixture of steam and feedstock to traverse the feedstock
injection section. A restriction is then that mixing must take place in a timescale on the
order of 0.1 to 3.0 ms depending on the allowed length.
This level of mixing suggests an agressive injection method to stir up the flow sufficiently well, but this comes with an associated loss in stagnation pressure that terminates
supersonic flow by the early formation of a normal shock in the duct. This trade off has
been explored by scramjet fuel injection designers. Scramjets were introduced in Section 2.3

9.1

Scramjet injector technology

The problem of injection into a supersonic cross flow has important applications in thrust
vectoring of spacecraft and fuel injection in scramjet engines so the approach taken in this
work has been to build on the techniques in use in these environments.
Development of an efficient combustion chamber in scramjets hinges on a suitable means
of delivering the fuel. When injecting fuel, clearly the first aim is to mix the fuel and the
air, and this can be achieved by either a diffusion process (a step-slot in the combustor
wall) or by using a penetration mechanism, either by a physical strut or by injecting with
suitable momentum at an angle to the chamber wall. If the gentler diffusion process is used,
the combustion chamber is generally long, and therefore combustion chamber skin-fraction
losses may become significant. The use of a penetration mechanism usually result in a
shorter combustion chamber; however, shocks are produced by the penetration mechanism,
and the total pressure losses generated by these shocks are important. Hence both systems
have losses associated with them.

121

9.1.1

Mixing in a Supersonic Flow

Before examining the pros and cons of the various injecting techniques available for use in
the shock wave reactor, it is opportune clarify, briefly, the dynamics of supersonic mixing.
In a scramjet, because of the very high flight speed, the residence time for atmospheric
air injested is on the order of a millisecond, just as it is in the feedstock injection section.
Unfortunately, compressible flow mixing is slower than subsonic or incompressible flow.
Experiments and numerical studies [25] show that increased compressibility reduces mixing
layer growth rates and reduce mixing rates. The effect of compressibility is to reorganizes
the turbulence field and modify the development of turbulent structures by limiting the
propagtion of disturbances that promote mixing. Several phenomena result in the reduction
of mixing with increasing Mach, including the velocity difference between injected and main
flow.
The primary physical mechanism by which mixing occurs in jets injected into supersonic
flow are compressible shear-mixing layers; eddies in the shear layers entrain injectant into
the main flow and increase concentration gradients. Since the molecular diffusive flux is
proportional to this gradient, diffusion occurs quickly over short length scales and so molecular mixing is promoted inside shear layers. Additional vorticity is typically associated with
improved mixing.
In general supersonic mixing requires more space than subsonic mixing does and the shock
structures that form when the injectant jet behaves as an obstacle in the flow, cause significant pressure losses. At the same time, these shock structures are three-dimensional and
oblique and flow passing through them is imparted additional vorticity (which promotes
mixing). Additional vorticity comes from so called baroclinic torques generated from spatial mis-alignment between density gradients of mixing layer and pressure gradients across
the shock plane.

122

9.1.2

Possible Injector Configurations.

What follows is a summary of all the various injection schemes and techniques used in
scramjet combustors that have been examined in this work. A brief description is provided
alongside each scheme to highlight important flow characteristics. It should be restated that
the three key injector selection criteria were rates of mixing, penetration into the main flow
and jet spreading rates.
Typical types of injection for scramjets are :
- TRANSVERSE INJECTION: Good mixing and penetration (scaled with J, the jetto-cross flow momentum ratio [2]).

s
J=

(U 2 )jet
(U 2 )f reestream

(9.1.1)

Characterized by three dimensional shock structure separation zones upstream and


downstream. Compressibility effects that reduce the shear layer entrainment and
growth in transverse injection.
- OBLIQUE INJECTION: Slower mixing and less penetration than transverse injection
but much lower pressure losses due to weaker bow shocks.
- MODIFIED INJECTANT FLOWS: Introducing axial vorticity into the feedstock gas
via tabs or convoluted surfaces can improve mixing (though implementation requires
complex machining and higher injection pressures). Vorticity can be introduced into
the air upstream of the injector via wedge shaped bodies. At shock interfaces where
strong pressure and density gradients exist, baroclinic torques can introduce additional vorticity.
- INJECTOR PORT SHAPE: Circular, elliptic and diamond shape are effective injection port shapes. The circular injectors are the simplest to fabricate and easily modeled to predict mass flow rates.Performance wise, elliptcal and diamond shapes bring

123

improvements like increased lateral spread and for the for diamond shapes along, more
penetration and faster mixing. Unfortunately elliptical and diamond shapes present
more difficulty when machining. Also, mass flow predictions are not easily carried out
analytically
- RAMPS (SWEPT AND UNSWEPT): A small, often tethrahedral obstruction placed
upstream of an injection port to shelter the injectant jet. With small injection angles
they can produce near streamwise injection. Ramps cause streamwise vorticity and
enhanced mixing. Swept ramps create more vorticity than unswept ramps. Require
active cooling in high temperature flows
- IN-STREAM INJECTORS (STRUTS): mainly desirable due to streamwise injection
capability which reduces stagnation pressure and thrust loss. Require active cooling
in high temperature flow
- PULSE INJECTORS: Moving rotary valves with similar motions to those in a rotary
engine induce velocity fluctuations. Mixing is improved due to additional shear layer
area, though the machanism is complex.
- AERO-RAMPS: Aeroramps use the wake behind an upstream fluid jet to shelter a
downstream jet. Arrays of injector nozzles at various inclinations act to mimic the
physical swept ramp design.
Several of these items are pictured together in Figure 9.1

Figure 9.1: A Selection of Typical Injecting Techniques for Scramjets Ref.[42]

124

From the list above, it is clear that certain designs are fraught with practical complexity.
Any physical obstacle in the flow such as a strut or ramp requires cooling in the extremes
of temperature expected in this flow region. Likewise, it was not yet a requirement that
modified port shape or injectant streams were absolutely necessary.
Ease of manufacture was kept in mind at all times during selection of an appropriate injector configuration. Transverse or oblique injection provide a very simple solution (from a
manufacturing perspective) and their operation is studied and used extensively in scramjet
combustors. With this broad array of research experience backing up their performance,
they were chosen as a candidate design to stand against the thorough set of empirical and
numerical fluid flow analyses that follow.
In (Figure 9.2) a sketch of the complex structures within the flow field in the vicinity of a
transverse injection port is shown.

Figure 9.2: Sketch of a Flowfield in the Vicinity of A Transversely Injected and Underexpanded Jet. Ref.[40]

In the figure, a jet of injectant emerges into cross flow travelling from left to right, accelerates

125

out of the nozzle and rapidly bends toward the downstream direction due to mixing and
momentum exchange. The principal feature in the jet plume is the barrel shock structure
which terminates in a Mach disk and is followed by a less clearly defined expansion zone.
A strong bow shock forms at the leading edge of the jet which curves away and weakens
into a Mach wave further downstream. A separation zone ahead and behind the jet affects
the shock structures. Not shown in the illustration are the large scale vortical structures
although some of the shear layers can be seen and are labelled mixing boundary. Several
types of vortices develop in the vicinity of the jet. First, a shear layer forms around the
periphery of the jet and these eddies contribute to mixing (as seen in the image). Second,
a pair of horse shoe vortices (not shown) form near the wall and sweep around the injector.
[16]
To examine the pentration of such a jet, a model was sought that is able to predict the
penetration hights of a transversely injected gas into supersonic cross flow. Models typically
make use of flow parameters as inputs such as main flow velocity (Uf ), injection flow velocity
at the plane of the injector port (Uj ), injection pressure (pj ) and mass flow (m
j ). The model
allowed a better understanding of the requirements for methane injection pressure.
Studies in literature on the transverse jet are fairly consistent when describing the effects
of momentum flux ratio J in Equation 9.1.1 on jet penetration height. The model is generally accepted as a simple power-law and when normalized by the diameter of the port,
curves of penetration height collapse to a single line that depends strongly on J. Centerline trajectories of transverse jets with different momentum flux ratios have the form of
Equation 9.1.2:
 x n
y
=A
rd
rd

where

r = J 2m =

(U 2 )j
(U 2 )CF

m
(9.1.2)

In Equation 9.1.2, y is the distance from the jet exit in the original jet direction, and x

is distance in the crossflow direction.In the ratio r, the exponent m has set to to collapse
different momentum flux ratios flows to a single profile. This m value is also accepted in
supersonic flow. A skewed coordinate system, Equation 9.1.3, collapses all small angles
to a single profile, where the jet is inclined by an angle . This skewed coordinate system
is constructed by skewing the Y axis so that it is aligned with the initial jet velocity vector

126

and aligning the X axis with the crossflow velocity vector.

Y = y/cos()X = x ytan()

(9.1.3)

Slight variations on Equation 9.1.2 exist for different authors according to their modeling
approach. Determination of the empirical constants m and n are of interest here in order
to use the equations to predict penetration of methane into steam cross flow. The value of
J is easily calculable for the injector configuration under examination during design.
Unfortunately, increased J is, in general, associated with reduced mixing. Jets with low J
reach their fully developed state faster than jets with higher J. Injecting with an underexpanded jet (so called sonic injection) is advantageous to both mixing and penetration and
so it has been employed here. Sonic injection also allows straightforward analytical prediction of the expected flow parameters of injected feedstock by using choked flow equations
for a perfect gas (as in the pseudo 1-D model from Section 8.3.1). Since a total mass flow
rate of 0.1 kg/s is desired, this aids greatly in design iterations (described below) in which
various port sizes are evaluated.
9.1.3

Existing Models of Transverse Injection

Several authors have studied flow conditions similar to those present in the shock wave
reactor especially near the 2.1 Mach compressible flow regime. Their empirical studies or
and modeling enable the prediction of penetration height for the case of transversely injected
gasses. Before introducing the different models a a brief parametric analysis is carried out to
understand the sensivitivity and of penetration height on cross flow and injection conditions.
It is found that penetration is strongly dependent on momentum ratio (J), weakly dependent on free-stream Mach number (Mf ), and practically independent of jet Mach number
p

(Mj ), pressure ratio at the exit ( p2j , where p2 is the static pressure after the bow shock)

and density ratio ( 2j ) at the exit. A parametric statement including the aforementioned
parameters is:
h
h
=
d
d



pj j
J, Mf , Mj , ,
p2 2

(9.1.4)

127

Where

h
d

is the penetration distance scaled by the jet-exit diameter. It is important to

establish the interdependence of certain variables in Equation 9.1.4. For compressible flow,
the momentum flux ratio J can be written as:

J=

j pj Mj2

(9.1.5)

f pf Mf2

The normal-shock relation between p2 and pf

p2 = pf

2f Mf2 (f 1)

(9.1.6)

f + 1

Which for Mf 2 is well approximated by:

p2 pf

2f Mf2

(9.1.7)

f + 1

This enables us to relate J to p2 and write eq. 9.1.4 in the form




2f pj 2
pj j
h
h
=
J
M , Mf , Mj , ,
d
d
f + 1 p 2 j
p2 2
It is evident that, for a given

pj
p2 ,

(9.1.8)

the momentum ratio is a very strong function of Mj . For

fixed J, one has no choice but to decrease

pj
p2

while increasing Mj , and vice-versa. This

makes it difficult to distinguish between the individual effects of

pj
p2

and Mj . Equation 9.1.8

shows that the cross-flow Mach number has very weak influence on J, therefore Mf can be
varied by a large extent while keeping all other variables virtually fixed. The same applies
to density ratio, since J in the formulation of Equation 9.1.5 is independent of molecular
weight.
Seven applicable models (that are similar in terms of free stream Mach number, injection
pressure, injection velocity and momentum flux ratio) are now described briefly and all are
used to estimate penetration height. Note that their original notations for flow variables
are have not been modified. Their effectiveness shall be assessed thereafter.
1. Model by F.W. Spaid Ref. [44].
Spaid modeled the presence of the jet as an solid flow obstruction with the effective shape
of the injectant being represented as a quarter of a sphere of radius h followed by an axisymmetric half body (Figure 9.3 ).

128

Figure 9.3: Spaids model of the effective shape of the injectant causing body interaction in
the free stream. Ref.[44]

Spaid presents the following equation to predict the penetration height that the secondary
flow reaches.
h
=
c



1
M



Poj j 2
P CP




(j +1)/(j 1) 


 
2
P (j 1)/j 1/4
2
1
j 1 j + 1
Poj
(9.1.9)

Where h is Mach disk high (peak of concentration of the injected flow), c is a discharge
coefficient for the injector, Poj is the stagnation injecting pressure, P is static pressure of
the main flow and CP is the pressure coefficient corresponding to the stagnation pressure
behind a normal shock, found from theory of supersonic flow of an ideal gas:

CP

2
=
2
M



+ 1 2
M
2

+ 1
2 +1
2 M

1
1

2. Model by M.R. Gruber Ref. [19].


This model fits empirical data with Equation 9.1.2 and results in constants A = 1.23, m = 12 ,
n = 0.344
y
def f J


= 1.23

x
def f J

0.344
(9.1.10)

129

def f is the effective jet orifice diameter and y is the height of the mixing boundary. (height
of the contour line of 10% injected flow concentration).

3. Model by F.S. Billig and J.A.Schetz Ref. [11].


The authors considered prior analyses and disagree on the assumptions generally made in
the past to model jet penetration phenomena. They assume that the interaction af a secondary jet and subsequent dispersion into the primary fluid is a two-stage process. In the
penetration stage, the jet retains its identity while being accelerated and turned in the flow
direction of the primary fluid. The second stage is considered to be a coaxial turbulent mixing process and is not discussed in their paper. The assumption used in subsonic analyses
is of the jet acting as a solid body that merges from the port and is bent downstream by
drag and distorted in cross section by pressure differences on the front and back faces and
viscous shear, has been retained in thier analysis. By the suitable definition of an effective
back pressure (peb ), a correlation is obtained between the normal distance to the centre of
the Mach disk and the ratio of injection pressure to effective back pressure.
They model the jet penetration with the Equation 9.1.11, though it underestimates the
effective height by about 34%
1
y
= Mj4

Dj

pj
peb

1
2

(9.1.11)

Where y is the coordinate of the centre of the Mach disk, the apex stands for sonic conditions and pedix j stands for jet conditions at point of injection .
The effective back pressure is calculated as peb = 23 pta where pta is the total pressure that
we have in the main stream behind a the bow shock.

4. Model by F.S. Billig, R.C.Orth and M.Lasky Ref. [10].


Revisiting previous work in Ref. [11] a solid body drag model was introduced and was used
to calculate the trajectory of either a matched pressure or over-expanded jet and the portion
of jet trajectory downstream of the Mach disk in an under-expanded jet. The development
of this model rests on the similarity that exists between a jet discharging into a quiescent
medium and a jet discharging into crossflow. In fact the structure in the primary flow is

130

essentially caused by a blunt-body interaction. To model this situation, heat addition, heat
transfer and mixing neglected in this blunt body region. The assumption that no mixing occurs may appear to be overly restrictive, but the authors show that the exchange of normal
momentum occurs in the near region of the injector, and both primary and secondary flow
are nearly coaxial before substantial mixing occurs. In this paper a comparison with Spaids
previous work (Ref. [44]) has been made too, and ultimately a new model for injected flow
was proposed (Fig. 9.4).

Figure 9.4: Billings New Model Of The Effective Shape Of The Injectant Causing Body
Interaction In The Free Stream. Ref.[10]

Finally, the authors develop a new expression for the maximum ordinate of the Mach disk
that results as a correction to the previous model equation given by Billig and Schetz
(eq. 9.1.12).
 1


p
1.8[1 0.7exp(0.07 pebj )]
2
Dj
y
1/4 pj
= Mj
+
1
a 2 2
Dj
peb
Dj
{1 + 0.16[1 exp( M
Mj )] }

(9.1.12)

Where the subscript a stands for undisturbed free stream conditions.


Comparisons between this equation, Spaids model and experimental data have been shown

131

in figures below (fig. 9.5).

Figure 9.5: Comparison of calculated jet cross-sectional areas and measured jet concentration contours (on the left). Comparison of calculated values of the maximum ordinate of
the Mach disk (on the right). (Ref.[10])

In the right most figure EQUATION 44 is Equation 9.1.12. To define the trajectory of
the Mach disk centreline in both models 3 and 4, the trajectory shape is assumed to be
parabolic in the region after the injection using the following equation.
y
=
Dj



x
Dj



y1
Dj



x1
y1


(9.1.13)

Where x is the axial distance measured from centre of injection port, y is the normal distance
from the point of injection, y1 is the penetration high evaluated with previous equations
(eq. 9.1.11 and eq. 9.1.12) and x1 is Mach disk x coordinate.


Mf
x1
= 1.25[1 exp
]
y1
Mj

5. Model by L.S. Cohen and L.J. Coulter Ref. [14].


Based on experimental data the following penetration height behavior is suggested :


poj
H = 0.6455D
peb

1
2

(9.1.14)

132

Where poj is the total injecting pressure and H is the distance from the wall to the centre
of the Mach disc.

6. Model by R.C. Rogers Ref. [40].


Because the transverse jet interaction flow field contains shock waves, flow recirculation,
and separation zones, which are often three-dimensional, completely analytical solutions
require the use of full elliptic form of the partial differential equations of motion. Rogers
analysis aimed to find a supersonic combustion model. Using available data he assumed
that the flow in the vicinity of a two dimensional injector can be represented by the 2 D
model presented in Figure 9.6.

Figure 9.6: Flow Model Of Transverse Jet Interaction. (Ref.[40])

The flow is confined in a duct with walls that diverge at small angles T and B . The
secondary flow is injected at an angle f from the vertical. The leading shock wave is
assumed to be a linear combination of the separation and bow shocks. This means that,
at large distances from the injector, the turning imparted to the flow is equal to that of a
wedge with the same shock angle as the upstream separation region. Unlike other models a
distinction between the location of the local peak concentration zk (represented by the top

133

of the Mach disk) and the of the mixing zone z has been made too.

1
2
f Mf2 pf
zk
= 1.5
J
2
d
j Mj peb

(9.1.15)

To estimate z Rogers provides three different correlations from experiments with different
flow conditions. In the end he concludes that there is no reason to justify the use of one
over the other, the edge of the mixing region is taken as an average value given by equations
9.1.16 and 9.1.17: The first is obtained for Mf =2.72.

1  
Mf 2 1 x 0.0866
z
= 1.68
J2
d
Mj
d

(9.1.16)

Where x is the axial coordinate along the duct after the injection. While the second correlation is for Mf =4.03.
 x 0.143
1
z
= 1.93Mf 2 J 0.3
d
d

(9.1.17)

7. Method by D. Papamoshou Ref. [37].


Papamoschou outlines how penetration height is affected by main flow and injection parameters and also examines the merits of pressure-matched jets. Prior studies [22] had found
that injecting with a pressure equal to that found near the jet port instead of injecting with
the jet over or under-expanded, would give an enhancement in penetration. However, due
to compressibility in the cross flow, the static pressure around the jet varies greatly along
its circumference as well as along its trajectory, so it is difficult to define pressure matching.
Heister and Karagozian [22] showed that equivalent pressure (the static pressure averaged
around jet perimeter) is approximately 0.5 p2 where p2 is the pressure after the bow shock
After different experiments, penetration measurements showed that, at fixed J, Mf and Mj ,
the ratio

j
2

does not influence penetration. Furthermore, penetration at fixed J and Mf is

insensitive to Mj . This also means that penetration is insensitive to the pressure ratio, so
under-expanded jets achieve about the same penetration as those that are pressure-matched.
Increasing free-stream Mach number Mf results in a modest improvement penetration, while
the jet Mach number and density ratio have no noticeable effect. In the end it was found

134

that highly under-expanded jets achieve about the same penetration as pressure-matched
jets.
Papamoschou approaches the calculation of penetration height by the aforementioned power
law model. He uses an equation found experimentally by Prette and Baines [38] that tracks
the development of the upper edge of a subsonic transverse jet (Eq. 9.1.18) and supplies
an equation for supersonic flow (Equation 9.1.19).
 0.28
y
x
= 2.63
rd
rd
h
= 4.34J 0.36
d

(9.1.18)
(9.1.19)

Where h is defined here as the maximum height of the jet trajectory.Moreover, the jet
trajectory is shown to level off at

x
d

= 6. Comparing Papamoschous equivalent subsonic

model with experiment results we can see that it usually overestimates the real penetration
height. Figure 9.7 illustrates the errors in the prediction of +43% when J = 2 (shown in
green) and +9.7% when J = 3.5 (shown in blue).

Figure 9.7: Model Errors in Penetration versus Momentum Flux Ratio J.

In order to decide which injector configuration to adopt (transverse, oblique, under-expanded,

135

Table 9.1: Injection Conditions in the Shock Wave Reactor


Flow

P0

(kPa)

(kPa)

Jet

1034

598

850

Main Stream

270

250

2.12

2100

J
( KgK
)

( Kg
s )

1.13

519

0.1

1.19

470

0.35

(K)

ramps, fins, etc.) the flow in the injection section of the shock wave reactor and the conditions under which methane is to be injected are evaluated. Thermodynamic and flow
variables are tabulated in Table 9.1.3
Using these conditions, the maximum momentum flux ratio is J = 2.16. At this point, an estimate for penetration height is availble by direct calculation. As mentioned Papamoschous
model overpredicts penetration with known errors and it shall be shown that Spaids model
is the most conservative so these are the upper and lower bounds. Making use of the
pseudo-1-D model from Section 8.3.1, maximum penetration heights are plotted along the
internal wall of the supersonic nozzle in Figure 9.8 . Note that the plot does not indicate
the trajectory of a jet but rather, it illustrates, for a given point in the nozzle where pressure, temperature, and velocity are known, that the penetration height for those free stream
conditions would be as shown. Plotting penetration in this way gives an indication of how
much penetration would be achieved by the end of the nozzle if injection ports were to be
located at various downstream distances. Penetration at the exit of the nozzle corresponds
to the penetration that would be achieved by an injector placed in the constant area duct
anywhere downstream of the nozzle exit.
Penetration height at conditions found at the nozzle throat would be to the nozzle centerline
according to all but Rogers z model. This location cannot be used to inject methane
however, since the static temperature is still too high at the nozzle throat and pyrolysis
would begin immediately. At the nozzle exit, Spaids model predicts penetration halfway
to the centerline whilst Papamoschous model full centerline penetration as desired. Up to
43% error can be expected in the latter, and thus a conservative approach is taken using

136

Figure 9.8: Jet penetration Development According To The 7 Models Described.

Spaids model to calculate penetration heights.


It should also be noted that in Figure 9.8 of the supersonic nozzle, the exit diameter is 3.0
inches, whilst the nozzle inlet and all other sections in the reactor have 3.5 inch internal
diameters. The nozzle exit area thus was reduced for the express purpose of achieving
enhanced penetration. Downstream of injection ports, the 3.0 inch diameter expands in a
shallow cone to match the downstream mixing sections 3.5 inch diameter internal diameter.
9.2

Size and Number of Injector Ports

In scramjet combustors, more than one injector port is used to inject and mix fuel with air.
The use of a single injector brings with it a very strong bow shock and large pressure losses.
In some cases, supersonic flow upstream can be affected to the point where the combustion

137

chamber can unstart. Hence, the use of several injectors is beneficial, though more than 8
injectors is not feasible in this case for reasons to be discussed.
The choice of sizing and number of injection ports must begin by recognizing that the
required mass flow rate of CH4 is 0.1 kg/s. For N injectors, the sum of their mass flow
rates is then constrained to m
= 0.1kg/s. Consider N identical injectors with the same
supply pressure and diameter. For a given supply pressure, mass flow rate can be increased
by increasing injector diameter as given by Equation 8.1.1.This can be expressed more
precisely as m j = m/N

 d2 where m j is the mass flow rate through one of the N injectors.


Or,since m
= const,

1
d  0.01
N

(9.2.1)

Now turning to penetration height and examining it from the perspective of the aforementioned power-law relation in Equation 9.1.2. The equation can be re-expressed as Equation 9.2.2:

y = A(Jd)1n xn

(9.2.2)

Substituting for d gives:

y=A

J 1n
N

1n
2

xn

(9.2.3)

Transverse injection through N circular ports of diameter d therefore involves a trade-off


between mass flow rate in and penetration height. Too many injectors reduces penetration
height. Too few injectors implies a large diameter and the correspondingly larger jets
produce more losses.
Another consideration is that of jet spread and covering a uniform area of the duct cross section with mixed gasses. Areas with too low or high a concentration of CH4 are to be avoided
since yields may be affected by non-homogeneous concentration by virtue of the corresponding non-homogenous partial pressures and temperatures that would in turn make pyrolysis
selectivity and conversions impossible to quantitatively assess. A better homogeneity of
mixing can be achieved with more injectors because their orifice areas drop as porportional

138

to

1
N

(according to Equation 9.2.1) and hence the sum total of the area covered by the

injectors jets is larger. In addition, with more injectors, their arrangement can be chosen
such that their cloud or plume area more efficiently covers the area of the round duct.
It was found that using a configuration with 4 identical injector ports with diamters of 0.223
inches each and arranged equally spaced around the circumference offered a marginally acceptible cloud distribution and still showed low penetration. Using Spaids model as in
Figure 9.9 penetration is seen to reach slightly more than halfway to the centerline.

Figure 9.9: Jet Penetration For 4 Injector Configuration.

Three injector ports provided insufficient homogeneity in the mixture. Using one of injectors
as a larger primary port with secondary injectors all having identical and smaller diameter
ports was tested, though the trade off between coverage area and penetration did not see
significant improvement. In particular, the penetration of secondary ports drops off quickly
as the primary injector is made large enough to allow centerline flow penetration and hence,

139

other techniques of increasing penetration height were sought.

9.3

Injector Penetration Height Enhancement

Several methods are available to enhance the penetration height of a side wall injector
port. Penetration is limited in jets because their transverse momentum is lost to the cross
flow as mixing occurs. The basic concept for all penetration enhancement methods is to
place an obstruction of some kind upstream of an injection port and allow injection to
take place into the slower wake of the obstruction. Qualitatively speaking, this protects
the jet from disturbance by the cross flow until it exits the wake. By delaying momentum
exchange between the jet and cross flow, mixing begins higher up in the flow resulting in
jet penetration height increase.
A common method noted in literature was the use of fins or fins for fuel injection in
supersonic flows. Unlike wedge shaped ramps, a fin is tetrahedral and tapers as it protrudes
into the flow (See Figure 9.10). Fins have been shown to enhance mixing, improve penetration height though they reduce lateral spread. They also minimize the stagnation pressure
loss associated with fuel injection processes. In particular,reduction in shock-related pressure loss would be around 25% in the flow regime found in the reactor. Pressure loss
reduction is achieved by initiating planar oblique shocks from their leading edge, lowering
the Mach number and reducing the total pressure drop across the usual bow shocks forming ahead of the jet plume. Penetration beyond the fin height appears to follow a similar
trajectory as in the baseline case (transverse injection without fin) [3]. Also, after a downstream distance of 25 jet diameters the penetration is 1.5 times the height of the fin [33]
(exemplified in Figure 9.10).

The fin employed in these study was a triangular blade with a rectangular base characterized by the following ratios : h/d = 3, w/d = 1 and l/d = 10 and sketched in Figure 9.11.
XP is the injection proximity to the fin.

Another study by Gruber et al. 2005 [33] examined fin length:width:height ratios to op-

140

Figure 9.10: Comparison between instantaneous images of the case of no fin (top figure)
and fin-guided fuel injection at various axial downstream distances. (Ref.[33])

Figure 9.11: Fin Geometry Shown With Injection Port And Defined Axis System.

timize penetration, jet spreading rate, mixing and total pressure losses. Three different
fins geometries with optimal heights, widths and fin distances from injection were obtained
from a previous computational study [18]. The two parameters that were emphasized in
the numerical study were XP and W . Two fins in this experiment were designed using the
optimal case ( XP /d 2 and W/d 1). One is named medium fin (h/d=4 , w/d=1.12 ,
l/d=6.9) and one is named tall (h/d=6 , w/d=1.12 , l/d=10.4). The third fin uses the second best case ( XP /d 3 and W/d 1.5) and is named wide (h/d=4, w/d=1.6 , l/d=6.9).
Inclination angle () of each fin is derived from l, h, and w. Note that fins experiments
were carried out in a 2-dimensional planar flow flame cavity rather than an axisymmetric

141

flow duct as in the shock wave reactor. However, fins used in a shock wave reactor are
anticipated to be on a scale which is at least 15 times smaller than the inner diameter of the
reactor duct diameter, and so flow surrounding the fin is closely approximated by planar
flow.
Normalized penetration (yj /hf in ) and lateral spread results confirm the best configurations
are the medium and wide cases (20-25% better than the baseline). The wide fin is the most
effective at increasing penetration, the tall fin loses its effectiveness as injection pressure is
decreased, while the medium loses effectiveness at higher J values. The tall fin is causes
strong shock waves to develop and losses are high. All fins provided more penetration and
less lateral spread than the case with no fin. The wide fin was the best performer overall.
The optimal solution for the shock wave reactor was a fin with ratios : h/d=3 , w/d=1.12
and l/d=10. A narrower fin is more easily attached to the cylindrical walls of the reactor
and a longer length fin with shorter height has a shallower angle and does not create as
large a pressure loss. The height ratio was deemed sufficient achieve centerline penetration.
9.3.1

Fin analysis

Considering that fins can improve penetration by 1.5 times the height of the fin when the
width of the fin is close to the diameter of the jet, it can be seen that since such a fin is anticipated to 0.75 inches high, resulting jet penetration is 2.12 times the height reached without
the use of a fin. An improved arrangement of injectors is to use a single fin on one injector
to achieve centerline penetraiton and another 4 injectors (hence 5 total) as secondary injectors to fill in the remaining circumferential space.Coverage area is sketched in Figure 9.12.

With the chosen geometry a thermal analysis of the fin was carried out numerically with
ANSYS using five candiate materials: titanium, zirconia, alumina and copper. A solid
model was created in Solid Works, meshed with tethrahedral elements and run in transient
thermal simulation mode for 15 seconds.
For boundary conditions, the base of the fin would be attached to the reactor wall and thus
it was pinned at the reactor wall temperature of 500 K. Convection boundary conditions

142

Figure 9.12: Transverse Section Through Shock Wave Reactor Duct Showing Methane
Distribution At 12 Injector Port Diameters Downstream Of Injection. (approximately to
scale).

were applied to 3/4 of the upper lateral surfaces and the remaining lower portions were set
to act as an insulated wall to prevent unrealistic temperature gradients forming where the
convective and pinned surfaces meet. Bulk convective temperture was 2100 K to represent
peak temperature combustion augmented steam flow. The results of the analysis are shown
in Figure 9.13 and show that such a fin melts if constructed from any of the candidate
materials aside from zirconium.
At the laboratory scale, a zirconium fin is not entirely dismissable as a solution to the penetration issue. It can withstand the flow temperatures and enhance penetration. Practical
issues arise when one considers the zirconium fins attachement to the wall. Any fin of
this scale also blocks 30% of the duct flow area. When considering scale up to commercial
size reactors, the prospect of using a fin is low. They must survive 3100 K flow temperatures (certainly requiring advanced cooling techniques) and thus they lose their geometric
similarity and so they do not scale well between different size commercial reactors.

143

Figure 9.13: Thermal analysis for a fin made in titanium. After 15 sec it has exceeded its
melting temperature (1941 K).

9.3.2

Dual Transverse Injection : Aeroramp

Alternative penetration enhancements were explored. Dual transferse injection schemes or


aeroramps are attractive since, being fluid-only ramps, require no cooling and scale well.
Experimental and computational studies [17] , [6], and [28] have demonstrated that an aerodynamic ramp injection scheme has better mixing characteristics than a single transverse
one.
The flow field of dual transverse injection into a supersonic cross flow is very complex due
to various shock structures and vortical flows around injection flows. Figure 9.14 shows
schematic of such a system.

Three dimensional bow shocks form ahead of the front jet and ahead of the rear jet and these
are weaker than the bow shock of a single transverse injection scheme. A separation shock

144

Figure 9.14: Schematic View Of Mean Flow Field Of Dual Transverse Injection.
Ref.(Ref.[28])

is also generated by interaction between the front bow shock and boundary. Quantitatively,
the front jet blocks the supersonic air flow impinging on the rear jet leading to a reduction of
the momentum of the main flow . Thus the effective jet-to-cross flow momentum flux ratio
of the rear injection flow increased, which assists the rear injection flow causing stronger
expansion. This blocking of main flow has significant influences on the mixing process
by generation of additional vortical features: vortex pairs form along the jet flows, horse
shoe vortices, separation bubbles and recirculating wake flows. Streamwise vortices roll up
injection flows and thus accelerate mixing in supersonic flow fields.
A distinctive feature of the dual injection is that the Mach disc of the rear jet is located
at higher position from the wall and is larger than the one of the front jet. Figure 9.15
illustrates this increased penetration taken from a numerical study by S.H. Lee [28] examining the replacement of a single transverse injection scheme with an aeroramp for the case
of H2 injection into air cross flow. The figure compares the distance between injectors L/D
to the height of the front and rear Mach discs. Note that mass flow injected in both cases
is the same because the diameters of the aeroramp are such that the sum of the mass flow

145

through both of them is equal to the mass flow through a single injector.

Figure 9.15: Comparison of Penetration Height for Single And Dual Transverse Injection
Systems (Left) And Showing Mach Disc Position (Right). Ref.(Ref.[28])

In general, an efficient aeroramp injection system requires careful selection of a broad class
of parameters, primarily: position of each injector, distribution of mass flow rate and momentum flux, individual injection angle and combinations of injection angles.
In Figure 9.15 the height of the Mach disc of the rear injection flow is strongly related
with the distance between injectors, while the height of the Mach disc of the front injection
flow is mantained regardless of this distance. The Mach disc height of the front injector
in the dual injection system is smaller than that of the single injection system because the
diameter of injection hole of the dual injection is smaller than that of the single injection
system.
Due to the blockage effects of the front jet, the rear injection flow is strongly influenced
attaining higher total mixing rates compared to a single injector, have higher penetration but
also incur more losses of stagnation pressure. Figure 9.15 shows and penetrations increased
as the distance between injectors increased until a critical distance but then decreased after

146

that critical distance. Thus there existed an optimal distance between injectors. According
to S.H. Lees findings for a momentum flux ratio of J = 2 the ideal distance for penetration
between front and rear injector is 4 dj , while the one that for less stagnation pressure loss
is 1 dj (Figure 9.16 ). For a spacing of 4 dj a penetration enhancement of about 1.66 times
the height reached by a single injector is expected.

Figure 9.16: Comparison Of Penetration Distances (Left) And Of Stagnation Pressure Normalized By Stagnation Pressure Of The Cross Flow. Ref.(Ref.[28])

Other experimental and numerical studies for aeroramp applications to scramjets are available, though their concern is often with flame holding capabilities in combustion cavities.
These studies then observe the effect of the aeroramp downstream of the backward facing
step that provides this flame holding function. Aside from S.H. Lees study, few other studies to date have examined the tradeoffs between penetration, losses and mixing in a simple
cross flow (no steps) like that found in the reactor. S.H. Lees quantitative data for dual
transverse injection schemes was utilized as a starting point for aeroramp design.
9.3.3

Estimating Lateral Spread and Coverage Area

Aside from penetration height and pressure losses discussed in the preceeding sections, the
design of the injector system must carefully consider the mixing uniformity of the jet plumes

147

far downstream. Unlike penetration height which requires only 20-30 injector diameters for
jet plumes to essentially level off at their maximum height, the lateral spread of injector
plumes can take longer; they do not level off in the same way and continue to grow more
linearly.
When steam and methane are mixed in the feedstock injection section and carried downstream to enter the 32 inch mixing section, the gasses are allotted additional time (and
hence distance) to mix. During this period lateral spreading continues and the rate is to
be estimated by the procedures outlined below. As mentioned, uniformly distibuted binary
mixture is desired over the cross section of the duct.
Lateral spread has been imaged with schlieren and mie scattering [33] though few quantitative exerimental measurements exist for the conditions in the shock wave reactor. Numerical
results for aeroramps such as those of S.H. Lee [28] are shown in Figure 9.17.

Figure 9.17: Lateral Spread Obtained from Numerical Study (Ref [28])

Open symbols represent the effective radius of the plume as defined by:
uR
|r c|dydz
u Rjet
r(x) =
jet dydz
r and c is the postion vector inside the concentraiton field and c is the position vector of
the center of mass of the injectant. Closed symbols represent single standard deviation H2

148

of the concentration field and centered at the peak concentration. It is defined as:

(x) =

uR

jet (|r c| rjet )2 dydz


uR
jet dydz

The plot shows that the mixing characteristics of dual transverse injection systems are very
different from those of a single injection system. Not only does an optimal distance between
injectors exist (for J = 2 this distance is 4dj ) but the effective radius of the jet can be
roughly correlated to the penetration height by r 13 h.
For non-aeroramp lateral spread, images of single transverse injection jet plumes (especially
those acquired by Montes et al. [33]) were examined and a basic curve fit of lateral width (w)
vs. penetration height (h) was developed by direct measurement in the images. The curve
fit suggests the relation w = 1.3h. This relationship was validated from multiple sources
([50], [20], [21], and [10]). A rigorous treatment is not required because a computation fluid
dynamics study was later carried out (Section 10 to more thoroughly verify different design
choices.
To obtain the effective area (Ac ) covered by the jet plume from the relations above for the
single and dual injectors, a circlular shape is assumed for the latter. For the former, most
jet plumes appeared to be between elliptical and rectangluar in shape. An average was
taken between the two and normalized by the jet orifice area (Aj ).


Ac
Aj

1
=
2

h w

2
2

2

!
2

+ (h w)

(9.3.1)

149

Chapter 10
INJECTOR CONFIGURATION

Methods presented in Chapter 9 for calculating the mixing effectiveness values like penetration height (h) and effective area covered (Ac ) were applied in a preliminary design
study. Injector diameter (dj ), number (N ), and location were parameters of interest in the
study. A MATLAB code segment was included with the aforementioned pseudo-1-D code
(Appendix D) that calculates h and Ac over the practical range of parameters dj and N .
The locations of the injectors were initially chosen such that they were arranged around
the circumference of the duct and equally spaced. No out-of-plane injection arrangements
were tested at this stage. The plane referred to here is a cross-section perpendicular to
the direction of the steam flow. The single injector port injectors were circular and the
aeroramp had the same mass flow rate as the other single port injectors.
Based on this simple analysis results indicated that equal mass flow rates through all injectors could not produce a viable solution. When N = 3, penetration was only 0.6 inches,
or 40% of the centerline height (1.5 inches) and this worsened for increasing N . Hence,
more mass flow rate was supplied to the aeroramp as this increases dj and h in turn (From
Equation 9.2.2).
By increasing the mass flow through the aeroramp up to 40% of the total mass flow rate
of CH4 gives a satisfactory penetration height of just over 1.5 inches for the aeroramp and
a 41% coverage area. This area corresponds to 25 jet diameters downstream. The jet
diameters were 0.18 inches for each of the two identical ports in the aeroramp and 0.21
inches for the single port injectors. Thus after only 6 inches downstream of the injectors,
41.8% of the duct was predicted to contain mixing flow.

150

10.0.4

Injector Placement and Injection Angle

Initial CFD studies in ANSYS Fluent confirmed that penetration and mixing requirements
were met. ANSYS Fluent domains, discretization, boundary conditions and solver setup are
described in detail in Chapter 11 and only brief details will be included here. In addition
to penetration and mixing results, the CFD studies also highlighted the significant pressure
losses generated by using four in-plane injectors and by injecting transversely (90 ) to the
cross flow. Pressure losses were found to be so adverse that a normal shock was noted to form
upstream inside the nozzle due to the blockage posed by the jets (exemplified Figure 10.1).
Figure 10.1 shows a single transverse injection port in the upper frame and the equivalent
aeroramp in the lower frame. Their bow shocks clearly produce a degree of pressure loss
that drops the flow velocity to subsonic conditions downstream. The nozzle used in the
analysis is in fact the original bell nozzle design. As previously noted in Section 8.2, its
wall contour changes suddenly and the strong compression waves generated are clearly seen
in the image as concentrated regions of yellow contours extending from the nozzle wall and
intersecting the bow shock ahead of the jet. A parallel improvement effort for both nozzle
and injector arrangement was necessary for the final design specifications to be sufficiently
met. The nozzle was lengthened and the shape altered toward the more ideal profile of a
supersonic wind tunnel.
As for the injectors, through many phases of CFD testing, lower angles like 60 and 30 were
tested as well as the careful selection of the arrangement of the various injectors. Inclined
(or oblique) injection reduces pressure losses but at the expense of lowering penetration
height. To compensate for this reduced penetration height, additional mass flow supply to
the aeroramp is required. Since the aeroramp acts as a primary injector by supplying the
bulk of the mass flow, the other smaller injectors are referred to as secondary injectors.
Secondary injectors angles were modified and set to 30 from the horizontal. In addition,
the aeroramp injectors were inclined at different angles. Since the upstream (front) port was
to face the full Mach 2.1 crossflow, the upstream shock wave is stronger than that formed
ahead of the downstream (rear) port. The former port was inclined at 30 and the latter at
60 from the horizontal.

151

Figure 10.1: Mach contour for single transverse injection (top) and dual injection (bottom)
.

152

The resulting penetration heights can be compared between various inclination angles in a
plot of CH4 mass fraction as seen in Figure 10.2. Aeroramp penetration is clearly superior.

Figure 10.2: Comparison of mass fraction of methane for different jet angles.

The difference between using a transverse or 60 inclined secondary injector is illustrated


in Figure 10.3. In the upper frame, three transverse secondary injectors are situated inplane with the aeroramp. Only one secondary injector is visible because the image is a slice
through the centerline axis of symmetry of the duct and the other two injectors do not lie
on this plane of symmetry but rather in the cross sectional plane perpendicular to the steam
flow direction. A strong normal shock is generated in the upper frame and injection is into
subsonic flow (hence the additional penetration). In the lower frame, oblique lamda-shocks
at high shock angles emerge from the walls and coalesce into a structure reminiscent of a
Mach disc near the centerline. Pressure loss has been reduced by inclining the secondary

153

injectors, though the flow is still subsonic downstream of the injection plane. Further
inclination to 30 was ultimately found to give the combination of low pressure loss and
sufficient penetration height, but even with this inclination angle, and an improved nozzle
shape, injecting in one plane always results in subsonic or transonic flow downstream which
goes against the supersonic mixing paradigm in shock wave reactors.
The number of out of plane injector arrangements is almost limitless. One particular configuration that found promise in its simplicity involved placing the aeroramp first, and in
close proximity to the nozzle exit plane. Downstream of this, the secondary injectors were
placed in their own separate injection plane, although their circumferential positions were
not altered. Since penetration of secondary injectors is improved when injecting into lower
Mach number flows, it is beneficial to place them in a plane that is far enough downstream
that it takes advantage of the blockage effects from the upstream aeroramp. From the CFD
results, the bow shock from the aeroramp was noted to have an angle of 40-45 , suggesting
that the secondary injectors be placed at 4.6 inches downstream of the aeroramp in the 3
inch duct. Rounding up to 5 inches provided a little margin. This staggered injection
scheme showed vast improvements in stagnation pressure losses and these are quantified in
the following chapter (11) .
10.0.5

Further Optimizations

Five different aspects were identified that could lead to improvement in the injector design
with some fine adjustments:

The Distance Between The Nozzle Exit Plane And The Front Aeroramp Orifice. Shorter
distances minimize heat losses, but placing the injectors too close to the exit plane
allows the aeroramp bow shock to interact with the compression waves created in the
nozzle and the outcome is an intolerably strong shock system and high losses. Test
cases were carried out for 1 to 5 inches. A distance of 2 inches is used in the final
design as shock losses at this distance are minimized and increasing the distance shows
little observable benefit.

154

Figure 10.3: Mach contour comparison between transverse and oblique secondary jet.

155

The Distance Between Front And Rear Aeroramp Studies by S.H. Lee [28] examined this optimal distance but did not assess the optimal distance for dual transverse
aeroramps but not the case of inclined injectors. CFD trade studies were carried out
for 3dj , 5dj and 7dj and comparison plots are shown in Figures 10.4 and 10.5).

Figure 10.4: Mach contour comparison between different distance from front and rear jet
in the aeroramp configuration.

Figure 10.4 illustrates the nature of the shock structure for various distances and the
strong normal shock formed in the upper frame rules out the 3dj distance as a viable
option.
Figure 10.5 illustrates the level and distribution of stagnation pressure losses. From

156

Figure 10.5: Stagnation pressure comparison between different distance from front and rear
jet in the aeroramp configuration.

the lower plane, it is evident that a large area of low stagnation pressure is formed
in the 7dj case whilst this is slightly smaller in the 5dj case; losses are slightly more
pronounced in the latter. Overall there is not a significant difference between the 5dj
and 7dj case and hence the 5dj case is selected.

Aeroramp and Secondary Injector Inclination Angles Final inclination angles for the
aeroramp ports have already been stated. These are 30 for the front and 60 for the
rear ports. Arriving at these angles involved running CFD simulations for different
combinations of angles. The computational expense was high for each simulation.

157

Only 10 increments in inclination angle were examined.


The front injector was set to 30 because most oblique injectors described in literature
were 30 . The author is not aware of any quantitative studies that explain the use
of the 30 angle, though it has been shown to be effective here also. Reducing this
angle is not viable due to the difficulties and hence cost involved with fabricating such
ports (highly inclined machining requries undue precision) and increasing the angle
only serves to exacerbate the leading bow shock and increase losses.
CFD results of changing the rear injector angle showed that increases in angle above
the optimal 60 were lossy whilst decreases of to 50 and even 40 actually reduced
losses quite effectively.However, injection angles under 60 suffered from very poor
penetration and hence, not viable.
Based on the methods of estimating pentration and lateral spread developed in Chapter 9, the aforementioned injection angles and injector arrangements were assessed
with the secondary injectors arranged as in Figure 10.6. The sketch is proportinally
to scale and the locations and sizes of the jet plumes are pictured as they would be
expected at 25dj downstream of the injection point (approximately 6 inches). The
sketch shows good agreement with final CFD simulations in Chapter 11.

The Distance Between Primary And Secondary Injectors By reducing this distance,
duct length can be reduced and heat losses are further minimized. Figure 10.7 shows
a comparison between a distance of 4 and 5 inches and the sensitivity of the shock
structure on distance is clearly evident. In the lower frame, the 4 inch distance causes
the formation of a normal shock structure in the viscinity of the centerline and this
shock induces more losses. The 5 inch distance originally calculated in Section 10.0.5
was not changed and selected for the final design.

Profile of the Duct Downstream of the Injectors The nozzle exit and duct diameters
in the zone of injection were selected to be 3 inches for sufficient penetration to be
achieved. However, sections downstream of this are 3.5 inches an expander is re-

158

Figure 10.6: Scheme of the transverse injection 6 downstream the aeroramp injection point.

quired. Expansions to larger diamters in supersonic duct flows firstly consist of a


diverging section in which both axial and radial components of velocity exist and then
a straightening section where flow is turned axially. The former is well approximated
by an isentropic process, but, whilst the latter can also be isentropic, it is not so
in general. A wall profile that abruptly turns supersonic flow generates undesirable
shock waves. The entire expansion process can be made to be largely isentropic if
particular wall profile is used and yet, from a machinging perspective, such a profile
requires costly CNC lathing time. On the other hand, a step in the wall is the simplest
expander but is also chracterized by several undesirable flow features like separation
and recirculation zones including shock waves. A conical expander presents the best
compromise between machining complexity and shock waves since it can be noted that
a 3 to 3.5 inch expansion requires a very small cone half angle half .
Expansions taking place over an axial length of 20 inches (half = 0.71 ) and 4 inches
(half = 3.6 ) were analyzed using CFD. Note that the total length of injection section

159

Figure 10.7: Mach contour comparison between different distances from primary to secondary jets: 5 (top) vs 4 (bottom.

(minus the nozzle) is 20 inches and so injection takes place inside the 20 inch long
expander. The 4 inch expander was placed downstream of the injectors. Results are
tabulated (Table 10.1) to compared exit conditions like stagnation pressure, stagnation
temperature, Mach number and the degree of homogeneity. Homegeneity was judged
by the degree to which the mass fraction of CH4 approached the fully mixed condition
(23% CH4 )
The 20 inch long expander showed an additional 139K temperature drop over the 4
inch expander but no difference in stagnation pressure. Better mixing was noted in the
4 inch section because injection was not taking place over the length of the expander

160

Expander Length (in)

Exit P0 (kPa)

Exit Mach

Exit T0 (K)

Exit Homegeneity

150

1.55

1689

Homogeneous

20

150

1.55

1550

Non-homogeneous

Table 10.1: Comparison a 4 inch and 20 inch Long Conical Expander

as in the 20 inch section, and thus penetration heights using the 4 inch expander were
better. Therfore the 4 inch expander was utilized in the final design in Chapter 11.

161

Chapter 11
FINAL CFD SIMULATION RESULTS

This chapter provides a more detailed description of the CFD analysis used to acertain the
efficiency of the entire feedstock injection section. Described here are CFD meshing, fluid
models, discretizations, boundary conditions, convergence and finally results. The final
design was validated according to specificatioins outlined in Section 7 using ANSYS Fluent
software. The simulation was run on an Intel Core 2 Quad CPU Q8200 with 2.33GHz
clockspeed and 4 GB of RAM, limiting the total number of elements to just over 700,000.
11.1

Computational Domain

The physical scope of the computationl domain was to start at the entrance to the feedstock
injection section, slightly upstream of the supersonic nozzle inlet, and include the duct up
to the downstream end of the mixing section (as in Figure 11.1)

Figure 11.1: Final configuration of nozzle, injector and mixing section.

11.2

Meshing and Grid Resolution

As before, the geometry was created using Solid Works and imported into the ANSYS
meshing package. The domain was discretized using tethrahedral elements and a prismatic

162

boundary layer mesh. The physical size of the domain to be simulated was large in comparison to the detailed flow structures around injector ports and the sharp gradients across
shock waves. A coarse mesh was used throughout and refined around these flow features
such that the smaller grid size more accurately captured their structure and effect on the
downstream flow field. The entire domain contained an axis of symmetry, allowing a symmetry boundary to be used to save computational and memory costs.
A uniform body mesh was first applied to the geometry with element sizes of 1.27
104 inches. This mesh was then refined using so called spheres of influence. The body
mesh gradually becomes finer in the vicinity of each sphere until it reaches the element size
specified for the sphere. Multiple overlapping spheres of influence were used. Near injectors,
four different spheres of influence with gradually descending element sizes which were used
to create a very find mesh for capturing the jet shape and bow shock and gradually increase
the mesh size further away from a jet where flow property changes are less abrupt. This
can be seen in Figure 11.2. Using only the finest mesh size in one large sphere of influence
overstepped the memory capability of the simulation computer. In particular, sizes from
finest to coarsest are:

Size 1 - 1.27 10

inches

Size 2 - 2.28 10

inches

Size 3 - 2.28 10

inches

Size 4 - 7.62 10

inches

Around the injector ports Figure 11.3 shows the mesh refinement results. Jet inlet discretization must be fine to acquire greater accuracy than other regions as this determines
computed injected mass flow as well as bow shock strengths important for pressure loss
estimation.
Boundary layers contain high velocity gradients that are best captured by a by prismatic
elements. In the CFX-mesher built into Fluent, an inflation function which creates computationally efficient high aspect-ratio elements, stacking them in layers, as seen in Figure

163

Figure 11.2: Mesh Refinement Zones using Spheres of Influence.

Figure 11.3: Injector Inlet Mesh

164

11.4. The total thickness of all layers and the number in the stack can be specified. Equivalent accuracy can be obtained using a far larger number of tethrahedral elements but
their generation by the automatic mesher is prone to high levels of element distortion (a
significant source of numerical error).

Figure 11.4: Boundary Layer Prismatic Mesh

Boundary layers in the subsonic regions of the nozzle inlet are orders of magnitude thinner
than those found at the end of the nozzle and in the injection or mixing sections. Two
sets of boundary layer inflation meshes were selected to model this and by doing so, reduce
the required number of elements. Each mesh has 7 layers, 4, 36 105 inches thick for the
subsonic nozzle zone, and 7.62 105 inches for the remainder of the domain.

The final mesh contained 673554 elements, nearing the limit of computer memory capacity.
Fluent is able to adapt the automatically generated meshes in regions of high fluid prop-

165

erty gradients (such as those near shock waves) by refining individual elements. The adapt
feature was used to verify that the resolution of the mesh was sufficient. If regions were
identified for adaption in Fluent, these zones were manually refined by spheres of influence.
11.3

Solver Sets and Boundary Conditions

Solvers were setup according to the prior modeling procedure in Section 8.3.2 (density-based
equations sets, standard k- turbulence model, etc.). Unless explicitly mentioned hereafter,
all settings are identical to those used in the nozzle analysis of Section 8.3.2.
In this final simulation two species were involved, H2 Oand CH4 . An eddy dissipation model
was used to model the binary mixture. It is based on the assumption that chemical reactions are fast relative to the transport processes of the flow. When the reactants mix at a
molecular level they instantaneously form products. The model assumes that the reaction
rate may be directly related to the time required to mix the reactants at the molecular
level. In turbulent flows this time is determined mainly by eddy properties. Therefore, the
reaction rate is proportional to a mixing time defined by the turbulent kinetic energy, k,
and its dissipation,  according to

rate


k

This concept of reaction control is applicable to a wide range of industrial combustion


problems. Unlike the nozzle analysis in Section 8.3.2, the calculation enabled the use of
multi-species transport that treat mixing and transport of chemical species by solving conservation equations describing convection, diffusion and reaction for each component specie.
Thermal diffusion is included in this equation set. Flow species were treated as a mixture
with a density that follows the ideal gas law and specific heat (CP ) evaluated through thermodynamic mixing laws. This law computes the mixtures specific heat capacity as a mass
fraction average of the pure species heat capacities(eq: 11.3.1).

CP =

X
i

Yi C P i

(11.3.1)

166

Thermal diffusivity and mass diffusivity were set constant while viscosity was allowed to
vary with temperature following Sutherlands law.
The required set of boundary conditions now included not only the steam inlet and mixed
CH4 /H2 Ooutlet boundary conditions but also a boundary condition for each injector. The
steam inlet was of type pressure inlet with stagnation (1034214 Pa) and static (598240
Pa) conditions supplied. For the injectors, the velocity vector was chosen to be normal to
the inlet plane (rather than the plane of the orifice into the main duct). After verifying
that changing turbulence k- model coefficents held no noticable changes their values were
left as defaults. Injector total pressure and temperature were set to 1034 kPa (150 psi) and
850 K. In order not to influence the solution too greatly, the only condition specified at the
flow exit was pressure. To model the preheated reactor walls, a wall temperature was set
at 600 K and and average thickness to 0.75.

An implicit solver was initially selected as it is generally associated with greater numerical
stability. However, to the contrary, the explicit scheme proved more stable using low Courant
numbers between 0.05 and 0.1. The scheme was run in steady state mode but this quickly
became unstable if all injectors were active from the start of the simulation. To avoid
stability issues, an elaborate sequence was required in which injectors were turned on in
certain orders. The simulation was allowed to run with no injectors active until a supersonic
flow field was established throughout the domain and residuals indicated convergence. Then
the aeroramp was turned on and again the solution was allowed to converge. Finally the
secondary injectors were turned on, each in turn, and the solution allowed to converge in
between.
Various metrics are availble in CFD pacakges to gauge the level of convergence in a flow
solution. As with any numerical scheme, a solution cannot converge to a perfect solution
due to the inherent truncation errors in the discretized form of the governing flow equations.
The solution can only reach a state of minimum error. Error referred to here is the level to
which the governing equations are satisfied by the approximate flow solution. Residuals, net
mass balance and visual flow field inspection are the primary tools to evaluate convergence.
Fluents default thresholds for convergence are set to 103 for all flow variables (continuity,

167

momentum etc.) with the exception of energy which has a threshold of 106 . Residuals
are generally a type of mathematical norm of the individual errors at each point in the
mesh, and as such, they are effective but not authoritative sources for judging convergence.
Thus, the state of residuals, the imbalance between incoming and outgoing mass flows and
a critical judgement of the structures seen within the flow field are all used to evaluate
convergence in the final simulation. In particular, the simulation is considered converged
when residuals cease with a downward trend and when mass flow imblance is less than 2%
of the total mass flow through the system (that is m
imbalance < 0.0009kg/s)
11.4

Results and Conclusions

The paradigm of supersonic mixing in the shock wave reactor relies on a mixture of H2 Oand
CH4 at supersonic Mach and hence a flow temperature lower than the pyrolysis temperature
of 1600 K. In Figure 11.5, the exit Mach number from the final simulation confirms this at
M = 1.35.

Looking at Mach contour we can see that the nozzle accelerates the main flow up to Mach
1.9. Injectors produce under-expanded jets that that expand and terminate in Mach disks
with upstream bow shocks in a complex flow field. Downstream, the conical expander reaccelerates the mixing flow to Mach 1.7-1.8 and then a sequence of shocks decelerate the
flow down to M = 1.35.
Closer examination of the region downstream of the injectors (Figure 11.6) reveals a sequence
of lambda shocks developing. This structure develops at the end of the conical expander
where two very weak oblique shocks emerge from the walls when the flow is turned axially.
They intersect at the centerline, reflect and then intersect the walls further downstream. In
the region close to the wall, the interaction between the shock and boundary layer induces
an adverse pressure gradient and thus the well known separation region characterizing a
lambda shock.
The lambda shock sequence initiates gradual recovery in flow static temperature (Figure
11.7). Static temperatures near the walls are higher than pyrolysis temperature in the
separation regions of the lambda shocks; some pyrolysis is thus expected. Methane con-

168

Figure 11.5: Mach Contour Lines.

centrations are low in this region however, so not much of the total methane mass flow is
expected to be pyrolyzed. Even so, a temperature drop corresponding to the endothermic
reaction is anticipated near the walls, lowering the heat losses predicted by the simulation.
Downstream of this sequence of shocks, the static temperature at the end of the mixing
section is noted to be around 1550-1650 K which is at the 1600 K threshold of pyrolysis
(Not shown in Figure 11.7).
The total temperature will be approximately 1650-1700 K and the thermal mixing in the
flow appears to be complete by the exit of the mixing section, as illustrated in Figure 11.8).
Despite temperature losses, 1700 K mixture of CH4 and H2 Ocontains sufficient enthalpy for
pyrolysis
To have an estimate for the extent of heat losses across the each section and identify where
the most significant losses occur, two different simulations were run: A case with adiabatic
walls and a case with the wall pre-heated to 600K. Integrating the total temperature through
four different transverse planes along the channel and comparing these by subtraction gives

169

Figure 11.6: Sequence of Lambda Shocks And Section Geometry.

Figure 11.7: Static Temperature Contours at the Beginning of the Mixing Section.

170

Figure 11.8: Total Temperature Contours In Injector And Mixing Section.

the total temperature loss (T ) from the entrance to the nozzle up until that section .
Individual losses across just a section are given by Tloss of Section. This data is shown
in Table. 11.1. The diverging section of the nozzle can expect the most heat loss measured
by a temperature loss of 35 K which is 65% of the total 55 K loss across the whole length
of the feedstock injection section and mixing sections combined. The estimate is based on
the assumption of 600 K walls throughout as a worst case scenario as the liners and nozzle
are expected to heat up and operate at higher temperatures.
Refer to a plot of the stagnation pressure field in Figure 11.9. The stagnation pressure is
seen to drop from an initial 2.5 bar down to 1.2-1.3 bar at the exit of the mixing section.
The zone identified as the major contributor of this loss is that immediately downstream of
the aeroramp where mixing of the two aeroramp jets is occuring.
The jet penetration height and methane-steam mixing goals setout in Chapter 7 as shown

171

Section (in)

Adiabatic (kPa)

600 K Walls (K)

Tloss of Section

Nozzle Throat

2100

2093

Nozzle Exit

2100

2058

43

35

Upstream of Secondary Injectors

1867

1823

45

Mixing Section Exit

1705

1651

55

10

Table 11.1: Total Temperature Losses Along Nozzle-Injector And Mixing Sections

Figure 11.9: Stagnation pressure losses along nozzle-injector and mixing sections.

172

in Figure 11.11. Aeroramp and secondary injectors penetrate the flow more than predicted
by Spaids method [44] as in Figure 11.12.The method appears conservative and backs up
the results in Figure 9.8 in Section 9.1.3 that Spaids model underpredicts pentration. The
exit flow of the section appears well mixed where 80% of the flow area has a fully mixed
CH4 mass fraction of 23%.

Figure 11.10: Methane Mass Fraction Along The Duct And At The Exit Plane.

The development of the jet plume can be observed at various planes over a length of 8 inches
from the poinf of first injection. (Figure 11.11). Typical cloud structures rise progressively
up to the penultimate (second from the right) plane at 6 downstream of the first injection
point. Little increase in cloud height is noted beyond the is distance.

173

Figure 11.11: Jet Plume Development Over A Length Of 8 Downstream The First Injection
Point.

Figure 11.12: Jet Plume Development 8 Downstream The First Injection.

174

Chapter 12
CONCLUDING REMARKS

Firstly, the analytical power of the quasi 1-D model cannot be overstated. Its congruence
with a full 3-D axisymmetric CFD analysis is remarkable and there is a difference of only
around 7% for major flow variables which is likely due to boundary layer corrections. The
quasi 1-D model predicts Mach 2.12 at the nozzle exit whilst CFD predicts 1.9. Comparing
the calculation time involved for each, the former proves its utility by its varied use in this
design process for rapid iteration design, in the heat transfer analysis (Section 8.3.4) as well
as the penetration height estimates (Section 9.1.3).
Secondly, regarding the various injector schemes useful in axisymmetric geometries for ultrahigh temperatures above 3000 K (such as those expected in a commercial scale shock wave
reactor), it is clear that physical ramps lose much of their feasibility in this high temperature
regime. Active cooling requirements for a physical ramp make a configuration such as
an aeroramp far more suitable. Aeroramps are attractive for this purpose due to their
simple fabrication requirements, their improved penetration of 1.5x an equivalent transverse
injector and their ability to reduce pressure losses significantly when inclined.
Thirdly, using CFD results, it has been confirmed that the use of planar penetration height
and mixing data is a good approximation when evaluating axisymmetric geometries.
At the time of writing, fabrication of the supersonic nozzle and methane injectors are
complete and await insertion into the reactor upon completing combustion testing. Their
running efficiency is to be assessed within the final stages of the second shock wave reactor
project.

175

BIBLIOGRAPHY
[1] A.T. Mattick A. Hertzberg and D.A. Russell. Apparatus for initiating pyrolysis using
a shock wave. Academic Press, New York, 1994.
[2] G.N. Abramovich. The Theory Of Turbulent Jets. Cambridge, MIT Press, 1963.
[3] C. Aguilera, B. Pang, A. Ghosh, A. Winkelmann, and KH Yu. Scramjet Mixing Control
Using Fin-Guided Fuel Injection. AIAA Journal, 2009.
[4] L.F. Albright, McConnell C.F., and K. Welther. Thermal hydrocarbon chemistry. In
A.G. Oblad, H.G. Davis, and R.T. Eddigner, editors, Advances in Chemistry Series.
American Chemical Society: Washington, 1979. pp. 175191.
[5] L.F. Albright and J.C. Marek. Mechanistic model for formation of coke in pyrolysis
units producing ethylene. Ind. Eng. Chem. Res. American Chemical Society, Vol. 27(pp
751-759), 1988.
[6] C. D. Anderson and J.A. Schetz. Performance of an aerodynamic ramp fuel injector in
a scramjet combustor. Journal of Propulsion and Power, 21(2):371 to 374, 2005.
[7] A. Hertzberg A.T. Mattick and D.A. Russell. Proceedings of the 18th international
symposium on shock waves, sendai. In Shock controlled reactors. Sendai Japan, July
1992.
[8] A. Hertzberg A.T. Mattick, D.A. Russell and C. Knowlen. Shock waves, proceedings of
the 19th international symposium on shock waves, marseille. In R. Brun and L. Z. Dumitrescu, editors, Shock-controlled chemical processing. Springer-Verlag, Berlin, 1983.
pp. 209-214.
[9] D.B. Spalding B.E. Launder. Lectures in mathematical models of turbulence. In -.
Academic Press, London, England, 1972.
[10] F.S. Billing, R.C. Orth, and M. Lasky. Unified analysis of gaseous jet penetration.
AIAA Journal, 9(6):10481058, 1971.
[11] F.S. Billing, J.A. Schetz, , and W.F. Ng. Penetration of gaseous jets injected into a
supersonic stream. Journal of Spacecraft and Rockets, 3(11):16581665, 1966.
[12] B.L.Wang, H. Olivier, and H. Gronig. Ignition of shockheated h2 airsteam mixtures.
Combustion and Flame, Vol. 133(pp 96106), 2003.

176

[13] D.A. Russell C. Knowlen, A.T. Mattick and R.K. Masse. Dept. of energy paper refs
12-20. ETC, Vol. 4(Num. 1,3-15), 2000.
[14] L.S. Cohen and L.J. Coulter. Jet penetration control. United States Patent, Issued
Aug 1973. Patent number: 3752172, Filing date: Jun 14, 1971.
[15] D.A.Spence. The growth of compressible turbulent boundary layers on isothermal and
adiabatic walls. A.R.C., 21(44), 1959.
[16] J.E. Fay and T. Rossmann. Mixing measurements of transverse and oblique sonic jets in
supersonic cross flow. In Aerospace Sciences Meeting and Exhibit. AIAA, Reno,Nevada,
2006.
[17] R.P. Fuller, P.K. Wu, A.S. Nejad, and J.A. Schetz. Comparison of physical and aerodynamic ramps as fuel injectors in supersonic flow. Journal of Propulsion and Power,
14(2):134135, 1998.
[18] OV Gouskov, VI Kopchenov, KE Lomkov, VA Vinogradov, and PJ Waltrup. Numerical
researches of gaseous fuel pre-injection in hypersonic 3-D inlet. AIAA Paper 2000-3599,
July 2000.
[19] M.R. Gruber and A.S. Nejad. Mixing and penetration studies of sonic jet in mach 2
freestream. Journal of Propulsion and Power, March-April 1995.
[20] MR Gruber, AS Nejad, TH Chen, and JC Dutton. Mixing and penetration studies of
sonic jets in a Mach 2 freestream. Journal of Propulsion and Power, 11(2):315323,
1995.
[21] MR Gruber, AS Nejad, TH Chen, and JC Dutton. Compressibility effects in supersonic
transverse injection flowfields. Physics of fluids, 9:1448, 1997.
[22] S.D. Heister and A.R. Karagozian. Gaseous jet in supersonic crossflow. AIAA Journal,
28:819827, 1990.
[23] C. Hulet, C. Briens, F. Berruti, and E. Chan. A review of short residence time cracking
processes. International Journal Of Chemical Reactor Engineering, Vol. 3(Review R1),
2005.
[24] Dieter K. Huzel and David H. Huang. Modern engineering for design of liquid-propellant
rocket engines. American Institute of Aeronautics and Astronautics, Washington, 1992.
[25] G.S. Diskin J.P. Drummond. Fuel-air mixing and combustion in scramjets. In Joint
Propulsion Conference Exhibit. AIAA, Indianapolis, Indiana, 2002.

177

[26] E.S. Stern J.P. Stern. Petrochemicals Today. Edward Arnold, London, 1971.
[27] L. Kniel, O. Winter, and K. Stork. Ethylene: Keystone to the Petrochemical Industry.
Marcel Dekker, New York, 1980.
[28] Sang-Hyeon Lee. Characteristics of dual transverse injection in scramjet combustor,
part 1: Mixing. Journal Of Propulsion And Power, 22(5), 2006.
[29] R.K. Masse, D.A. Russell, A.T. Mattick, and C. Knowlen. Study of shock recompression
parameter dependence. In 22nd International Symposium on Shock Waves. Imperial
College, London, UK, 1999. Jan 1823.
[30] A.T. Mattick. The supersonicmixing, shockwave reactor: An innovative approach
for efficient chemical production. In . U.S. Department of Energy, 2002.
[31] Hatch Matur. Chemistry of Petrochemical Processes. Gulf Publishing Company, Texas,
1994.
[32] A. Mol. Transfer line exchangers. In L.F. Albright, B.L Crynes, , and W.H. Corcoran,
editors, Pyrolysis: Theory and Industrial Practice. Academic Press, New York, 1983.
Chapter 13: Pyrolysis furnace design: conventional and novel, p327.
[33] D.R. Montes, P.I. King, M.R. Gruber, C.D. Carter, and K.Y. Hsu. Mixing Effects of
Pylon-Aided Fuel Injection Located Upstream of a Flameholding Cavity in Supersonic
Flow, 2005.
[34] Chris Morley. Gaseq. www.gaseq.co.uk. Software: Chemical Equilibria in Perfect
Gasses, Version 0.79.
[35] L.D. Nill and A.T. Mattick. An experimental study of shock structure in a normal
shock train. In AIAA Paper 96-0799. 34th AIAA Aerospace Sciences Meeting, Reno,
1996. Jan 1518.
[36] R. Orriss and H. Yamaguchi. Idemitsus chiba ethylene plant proves modern technology.
Oil & Gas J., Vol. 9(Num 27), 1987.
[37] D. Papamoschou and DG Hubbard. Visual observations of supersonic transverse jets.
Experiments in Fluids, 14(6):468476, 1993.
[38] B.D. PRETTE and D. BAINES. Profiles of the round turbulent jet in a cross flow.
Hydronaut. Div. ASCE 92, p. 53, 1967.
[39] Tao Ren. Petrochemicals from Oil, Natural Gas,Coal and Biomass: Energy Use, Economics and Innovation. Ipskamp Drukkers B.V, Enschede, The Netherlands, 2009.

178

[40] RC Rogers. Model of Transverse Fuel Injection in Supersonic Combustors. AIAA


Journal, 18:294301, 1980.
[41] L. H. Ross. Pyrolysis: Theory and industrial practice. In L.F. Albright, B.L Crynes,
and W.H. Corcoran, editors, Chapter 13: Pyrolysis furnace design: conventional and
novel, p327. Academic Press, New York, 1984.
[42] Sofia Sartori. Tesi di laurea: Shock wave reactor. Masters thesis, Politechnico di
Torino, Corso Duca degli Abruzzi, 24, Italy, May 2010.
[43] N. Shuichi. Methane conversion by various metal, metal oxide and metal carbide
catalysts. Catalysis Surveys from Japan 4, Vol. 4(Num. 1,315), 2000.
[44] F.W. Spaid. A study of secondary injection of gases into a supersonic flow. PhD thesis,
California Institute of Technology Pasadena, 1964.
[45] Thomas Strobridge, John Moulder, and Alan Clark. Titanium combustion in turbine engines. National Bureau of Standards, Report Nos. FAA-RD-79-51, NBSIR 791616(July), 2000.
[46] Vincent Tanguay, Patrick Batchelor, Ramzi El-Saadi, and Andrew Higgins. Metal
combustion in high-speed flow. In 43rd AIAA Aerospace Sciences Meeting and Exhibit.
AIAA, Reno, Nevada, 2005.
[47] NASA X-43 Development Team. X-43. www.apod.nasa.gov.
[48] T.C. Tsai and L.F. Albright. Importance of surface reactions in pyrolysis units. In
W.H. Corcoran L.F. Albright, B.L. Crynes, editor, Pyrolysis: Theory and Industrial
Practice. Academic Press: New York, 1983. Chapter 10.
[49] T.C. Tsai and L.F. Albright. Thermal cracking of hydrocarbons. Encyclopedia of
Chemical Processes, Vol. 5(pp 29752986), 2006.
[50] S.T. Yu and H. He. Three-dimensional simulation of transverse injection in a supersonic
flow by the CESE method. In 41 st AIAA Aerospace Sciences Meeting & Exhibit, Reno,
NV, 2003.
[51] M. J. Zucrow and Joe D. Hoffman. Gas Dynamics. Wiley, 1976.

179

Appendix A
LABVIEW BLOCK DIAGRAMS

LabVIEW block diagram images have been spliced to fit.

Figure A.1: Block Diagram 1

Figure A.2: Block Diagram 2

180

Figure A.3: Block Diagram 3

Figure A.4: Block Diagram 4

181

Figure A.5: Block Diagram 5

182

Table A.1: Flow Valve Opening Sequence

Time (s)
-5.0
-5.0
-5.0
-0.5
0.0
1.0
10.0
10.0
10.0
10.0
10.0
10.0
13.0
13.0
14.0
16.0
16.0
16.0
16.0
16.0
16.1
19.0
20.0
29.0
29.0

Phase
Prestart
Prestart
Prestart
Prestart
Steam
Steam
Combustion
Combustion
Combustion
Combustion
Combustion
Combustion
Feedstock Injection
Feedstock Injection
Feedstock Injection
Shutdown Sequence
Shutdown Sequence
Shutdown Sequence
Shutdown Sequence
Shutdown Sequence
Shutdown Sequence
Shutdown Sequence
Shutdown Sequence
Shutdown Sequence
Shutdown Sequence

Signal Name
Data Acquisition & Failsafe & Safety Logic
Dump Tank "Partial Open" Ballast Valve
Heater Isolation Valve
Heater Isolation Valve
PHB Valve
Spray Bar Valve (Low Flow Rate)
Spray Bar Valve (Low Flow Rate)
Spray Bar Valve (High Flow Rate)
Dump Tank "Partial Open" Ballast Valve
Dump Tank "Full Open" Ballast Valve
H2 Isolation Valve
O2 Isolation Valve
Methane Isolation Valve
Heater Isolation Valve
Sample Bottles
Methane Isolation Valve & Sample Bottles
Methane Purge Ballast Valve
O2 Isolation Valve
H2 Isolation Valve
Spray Bar (Low Flow Rate)
Spray Bar (High Flow Rate)
PBH Valve
Spray Bar (Low Flow Rate)
Methane Purge Ballast Valve
Data Acquisition

Position
Start
Preset OPEN
Preset OPEN
Preset CLOSE
OPEN
OPEN
CLOSE
OPEN
CLOSE
OPEN
OPEN
OPEN
OPEN
OPEN
CAPTURE
CLOSE
OPEN
CLOSE
CLOSE
OPEN
CLOSE
CLOSE
CLOSE
CLOSE
Stop

Pin
1-44 SAVE
1
3
3
4
5
5
6
1
2
7
8
9
3
10
9
11
8
7
5
6
4
5
11
1-44 SAVE

11 10 9 8 7 6 5 4 3 2 1 Decimal
0 0 0 0 0 0 0 0 0 0 0
0
0 0 0 0 0 0 0 0 0 0 1
1
0 0 0 0 0 0 0 0 1 0 1
5
0 0 0 0 0 0 0 0 0 0 1
1
0 0 0 0 0 0 0 1 0 0 1
9
0 0 0 0 0 0 1 1 0 0 1
25
0 0 0 0 0 0 0 1 0 0 1
9
0 0 0 0 0 1 0 1 0 0 1
41
0 0 0 0 0 1 0 1 0 0 0
40
0 0 0 0 0 1 0 1 0 1 0
42
0 0 0 0 1 1 0 1 0 1 0
106
0 0 0 1 1 1 0 1 0 1 0
234
0 0 1 1 1 1 0 1 0 1 0
490
0 0 1 1 1 1 0 1 1 1 0
494
0 1 1 1 1 1 0 1 1 1 0
1006
0 0 0 1 1 1 0 1 1 1 0
238
1 0 0 1 1 1 0 1 1 1 0
1262
1 0 0 0 1 1 0 1 1 1 0
1134
1 0 0 0 0 1 0 1 1 1 0
1070
1 0 0 0 0 1 1 1 1 1 0
1086
1 0 0 0 0 0 1 1 1 1 0
1054
1 0 0 0 0 0 1 0 1 1 0
1046
1 0 0 0 0 0 0 0 1 0 1
1029
0 0 0 0 0 0 0 0 0 0 1
1
0 0 0 0 0 0 0 0 0 0 0
0

183

Appendix B
MATLAB CODE

The file post processor gui.m is the main processing file. It is associated with a .fig GUI
object.post processor gui.m makes calls to four functions.
1. get Line Info.m reads line names and meta information from an excel spreadsheet
prepared for each run
2. calibrate.m which applies calibrations to each sensor
3. distPlot.m which plots pressure and temperature vs. distance
4. transientPlot.m which plots pressure and temperature vs. time
Code for each is presented below:
B.1

post processor gui.m

function varargout = post processor gui(varargin)

2
3

%% Begin initialization code - DO NOT EDIT

gui Singleton = 1;

gui State = struct('gui Name',

mfilename, ...

'gui Singleton',

'gui OpeningFcn', @post processor gui OpeningFcn, ...

'gui OutputFcn',

@post processor gui OutputFcn, ...

'gui LayoutFcn',

[] , ...

'gui Callback',

[]);

10
11

if nargin && ischar(varargin{1})

12

gui State.gui Callback = str2func(varargin{1});

13

end

14

if nargout

15
16

gui Singleton, ...

[varargout{1:nargout}] = gui mainfcn(gui State, varargin{:});


else

17

gui mainfcn(gui State, varargin{:});

18

end

19

% End initialization code - DO NOT EDIT

184

20
21

% --- Executes just before post processor gui is made visible.

22

function post processor gui OpeningFcn(hObject, eventdata, handles, varargin)

23

global PLOT ON PLAY SAVE DATA data file loc string plot time run time CALIBR DATA TABLE N time points

24

global line names line units line plot range lower line plot range upper line channel line type

25
26

% Choose default command line output for post processor gui

27

handles.output = hObject;

28
29

% Update handles structure

30

guidata(hObject, handles);

31
32

% This sets up the initial plot - only do when we are invisible

33

% so window can get raised using post processor gui.

34

if strcmp(get(hObject,'Visible'),'off')

35
36

plot(0);
end

37
38

% UIWAIT makes post processor gui wait for user response (see UIRESUME)

39

% uiwait(handles.figure1);

40

PLOT ON = 2;

41

PLAY = 0;

42

plot time = 0.1;

43

SAVE DATA = 1;

44
45

%Get the data directory

46

load listbox('C:\JF\Work\Uni\UW\2009\SWR\Measurement\Data Acquisition and Sequencing\Data Files',handles)

47

runs = get(handles.popupmenu3,'String')

48

run = runs{length(get(handles.popupmenu3,'String'))}

49

data file loc string = strcat('..\Data Files\',run,'\')

%location of the data files

50
51
52

%populate GUI

53

set(handles.listbox1,'Value',9)

54

set(handles.slider1,'Value',0.1)

55

set(handles.edit1,'String',num2str(plot time));

56

set(handles.popupmenu plot type,'Value',1)

57
58

%refresh data for the first time

59

pushbutton update Callback(hObject, eventdata, handles)

60

popupmenu plot type Callback(hObject, eventdata, handles)

61

set(handles.text5,'String',data file loc string);

62

set(handles.popupmenu3,'Value',length(get(handles.popupmenu3,'String')))

63

warning off MATLAB:hg:uicontrol:ParameterValuesMustBeValid

64

warning off MATLAB:polyfit:RepeatedPointsOrRescale

65
66
67

% --- Outputs from this function are returned to the command line.

68

function varargout = post processor gui OutputFcn(hObject, eventdata, handles)

69

% varargout

cell array for returning output args (see VARARGOUT);

70

% hObject

handle to figure

71

% eventdata

reserved - to be defined in a future version of MATLAB

72

% handles

structure with handles and user data (see GUIDATA)

73
74

% Get default command line output from handles structure

75

varargout{1} = handles.output;

76

185

77

% --- Executes on button press in pushbutton update.

78

function pushbutton update Callback(hObject, eventdata, handles)

79

global CALIBR DATA TABLE PLOT ON PLAY SAVE DATA data file loc string run dt plot time time CALIBR DATA TABLE N time points

80

global line names line units line plot range lower line plot range upper line channel line type

81

axes(handles.axes1);

82

cla; clc;

83

warning off MATLAB:polyfit:RepeatedPointsOrRescale

84
85

%% READ THE RAW DATA FILES

86

RAW DATA Ch0=[];RAW DATA Ch1=[];RAW DATA Ch2=[];RAW DATA Ch3=[];

87

RAW DATA Ch0 = dlmread(strcat(data file loc string,'ch0raw.txt'), '\t');

88

RAW DATA Ch1 = dlmread(strcat(data file loc string,'ch1raw.txt'), '\t');

89

RAW DATA Ch2 = dlmread(strcat(data file loc string,'ch2raw.txt'), '\t');

90

RAW DATA Ch3 = dlmread(strcat(data file loc string,'ch3raw.txt'), '\t');

91
92

%% READ THE LABVIEW CALIBRATED DATA FILES

93

CALIBR DATA Ch0=[];CALIBR DATA Ch1=[];CALIBR DATA Ch2=[];CALIBR DATA Ch3=[];

94

% try

95

CALIBR DATA Ch0 = dlmread(strcat(data file loc string,'ch0.txt'), '\t');

96

CALIBR DATA Ch1 = dlmread(strcat(data file loc string,'ch1.txt'), '\t');

97

CALIBR DATA Ch2 = dlmread(strcat(data file loc string,'ch2.txt'), '\t');

98

CALIBR DATA Ch3 = dlmread(strcat(data file loc string,'ch3.txt'), '\t');

99

% catch excpt

100

101

% end

disp(getReport(excpt));

102
103

%% CALCULATE THE NUMBER OF DATA POINTS

104

N lines = 16;

105

N total = min([length(RAW DATA Ch0),length(RAW DATA Ch1),...

106
107

length(RAW DATA Ch2),length(RAW DATA Ch3)]);


N time points = N total/(N lines+1);

108
109

%% RESHAPE THE DATA INTO A LARGE TABLE OF t vs. PRESS or TEMP

110

k = 0;

111

for i = 1:N time points

112

for j = 1:(N lines+1)

113

k = k+1;

114

RAW DATA TABLE 0(i,j) = RAW DATA Ch0(k,2);

115

RAW DATA TABLE 1(i,j) = RAW DATA Ch1(k,2);

116

RAW DATA TABLE 2(i,j) = RAW DATA Ch2(k,2);

117

RAW DATA TABLE 3(i,j) = RAW DATA Ch3(k,2);

118
119

DATA TABLE 0(i,j) = CALIBR DATA Ch0(k,2);

120

DATA TABLE 1(i,j) = CALIBR DATA Ch1(k,2);

121

DATA TABLE 2(i,j) = CALIBR DATA Ch2(k,2);

122

123
124

DATA TABLE 3(i,j) = CALIBR DATA Ch3(k,2);


end

end

125
126

% All data is contained in DATA TABLE

127

RAW DATA TABLE = [RAW DATA TABLE 0(:,2:end),RAW DATA TABLE 1(:,2:end),...

128

RAW DATA TABLE 2(:,2:end),RAW DATA TABLE 3(:,2:end)];

129

% DATA TABLE = [DATA TABLE 0(:,2:end),DATA TABLE 1(:,2:end),...

130

131

132

time = RAW DATA TABLE 0(:,1);

133

DATA TABLE 2(:,2:end),DATA TABLE 3(:,2:end)];

%extract the time data

186

134

%% SENSOR CALIBRATION

135

% Pressures:

Temperatures:

136

% 8-10 Combustor

40 R1 PBH Top

137

% 11-15 Mixer

41 R2 PBH Bottom

138

% 16-18 Transition

42 R3 PBH Mid starboard

139

% 19-21 Reactor

43 R4 X

140

% 22 Diagnostic

44 KR1 PBH Head

141

% 23 Dummy

46/47 KR3 H2/O2 Manifold

142

% 32/33 PBH Bottom/Top

48 K1 Comb 1

143

49 K2 Comb External Wall

144

50 K3 Mixer End Wall (M5)

145

51 K4 Reactor Beginning (R1)

146

52 K5 Diagnostic

147

56 C1 Combustor End (C3)

148
149

% Refer to calibrate.m -> type: 1-Pr 2-C 3-K 4-R 5-Opt

150

type = zeros(N lines*4);

151

type(1:6) = [5,6,7,8,9,10]; %H2 Manifold Pressure

152

type(7) = 13;

153

type(9:25) = ones(1,17);

%Pressure Transducers

154

type(33:34) = ones(1,2);

%Pressure Transducers

155

type(35) = 11*ones(1,1);

%Optical temp sensor

156

type(36) = 12*ones(1,1);

%Spray bar water flow rate

157

type(40:43) = 4*ones(1,4);

158

type(44:56) = 3*ones(1,13);

159

type(57:64) = 2*ones(1,8);

160

CALIBR DATA TABLE=[];

161

for i = 1:N lines*4

162
163

CALIBR DATA TABLE(:,i) = calibrate(RAW DATA TABLE(:,i),time,type(i));


end

164
165

% Calculate water mass flow rate by differentiating a polynomial fit of

166

% water mass vs. t

167

% P = polyfit(time,CALIBR DATA TABLE(:,36),9); Psh = circshift(P',1); Pderiv = Psh(2:end);

168

% CALIBR DATA TABLE(:,(N lines *4+1)) = polyval(Pderiv,time);

169

CALIBR DATA TABLE(:,(N lines *4+1)) = smooth(gradient(CALIBR DATA TABLE(:,36),time),10,'rlowess',3);

170
171

% Calculate the H2-O2 Injector Calibration Curve Averages

172

[t min diff,ti10] = min(abs(time-10)); %calculate the index at which t=10s occurs

173

[t min diff,ti60] = min(abs(time-(time(length(time))-60))); %calculate the index that is 60s from the end

174

Comb Inj Initial 10s Avg = mean(CALIBR DATA TABLE(1:ti10,5))/6.89475729-14.7

175

Comb Inj Final 60s Avg = mean(CALIBR DATA TABLE((end-ti60):end,5))/6.89475729-14.7 %average of last 60s in psig

%average of first 10s in psig

176
177

%Read the excel spreadsheet (Line Names.xls)

178

[line names,line units,line plot range lower,line plot range upper,line channel,line type] = get Line Info();

179

dt = mean(diff(time));

180
181

%% PLOTS

182

PLOT TYPE = get(handles.popupmenu plot type, 'Value');

183

PLOT NUMBER = get(handles.listbox1, 'Value');

184

set(handles.edit dt,'String',strcat(num2str(dt*1000,'%3.0f'),' ms'));

185

Tinit = -5; Tfinal = 35; Nt = (Tfinal-Tinit)/dt;

186

plot time = Tinit+get(handles.slider1,'Value')*(Tfinal-Tinit);

187

set(handles.edit1,'String',plot time);

188

listbox1 Callback(hObject, eventdata, handles)

189
190

SAVE DATA = get(handles.checkbox1,'Value');

187

191

if SAVE DATA == 1

192

save file name = strcat(data file loc string,'DATA TABLE-',run)

193

save title{1} = 'Time'; save title{66} = 'Mass Flow Rate';

194

for i = 2:65

195

save title{i} = strcat('Line-',num2str(i-1));

196

end

197

save DATA = [save title;num2cell([time,CALIBR DATA TABLE])];

198

cellwrite(strcat(save file name,'.csv'),save DATA)

199

disp('Export complete')

200

end

201

% --------------------------------------------------------------------

202

% Additional Plots

203

figure(1)

204

plot(time,-CALIBR DATA TABLE(:,60)+CALIBR DATA TABLE(:,61),'.-k')

205

title('Temperature Drop Due to Heat Loss

206

xlabel('time (s)'); ylabel('\DeltaT (\circC)')

207

% --------------------------------------------------------------------

vs. Time'); grid on;

208
209

function FileMenu Callback(hObject, eventdata, handles)

210
211

% --------------------------------------------------------------------

212

function OpenMenuItem Callback(hObject, eventdata, handles)

213
214

file = uigetfile('*.fig');

215

if isequal(file, 0)

216
217

open(file);
end

218
219

% --------------------------------------------------------------------

220

function PrintMenuItem Callback(hObject, eventdata, handles)

221

printdlg(handles.figure1)

222
223

% --------------------------------------------------------------------

224

function CloseMenuItem Callback(hObject, eventdata, handles)

225

selection = questdlg(['Close ' get(handles.figure1,'Name') '?'],...

226

['Close ' get(handles.figure1,'Name') '...'],...

227
228

'Yes','No','Yes');
if strcmp(selection,'No')

229
230

return;
end

231
232

delete(handles.figure1)

233
234
235

% --- Executes on selection change in popupmenu plot type.

236

function popupmenu plot type Callback(hObject, eventdata, handles)

237

global PLOT ON PLAY data file loc string plot time time CALIBR DATA TABLE N time points

238

global line names line units line plot range lower line plot range upper line channel line type

239
240

% Hints: contents = get(hObject,'String') returns popupmenu plot type contents as cell array

241

242

PLOT TYPE = get(handles.popupmenu plot type, 'Value');

243

cla(handles.axes1)

244

switch PLOT TYPE

245

contents{get(hObject,'Value')} returns selected item from popupmenu plot type

case 1

246

%Transient Plot

247

set(handles.listbox1,'String',line names);

188

248

line names;

249

line = get(handles.listbox1,'Value');

250

transientPlot(handles,line,time,'pressure',CALIBR DATA TABLE)

251

case 2

252

%Distance Plot

253

ss = line names(1:2);

254

ss(1) = 'Pressure'; ss(2) = 'Temperature'

255

256

set(handles.listbox1,'String',ss);
set(handles.listbox1,'Value',1);

257

plot time = str2double(get(handles.slider1,'Value'));

258

slider1 Callback(hObject, eventdata, handles)

259

end

260
261

% --- Executes during object creation, after setting all properties.

262

function popupmenu plot type CreateFcn(hObject, eventdata, handles)

263

% Hint: popupmenu controls usually have a white background on Windows.

264

265

if ispc && isequal(get(hObject,'BackgroundColor'), get(0,'defaultUicontrolBackgroundColor'))

266
267

See ISPC and COMPUTER.

set(hObject,'BackgroundColor','white');
end

268
269

set(hObject, 'String', {'Transient Plot', 'Pressure vs. Distance Plot'});

270
271
272

% --- Executes on slider movement.

273

function slider1 Callback(hObject, eventdata, handles)

274

global PLAY dt plot time time CALIBR DATA TABLE

275

% Hints: get(hObject,'Value') returns position of slider

276

277

dt = str2double(get(handles.edit dt,'String'));

278

Tinit = -5; Tfinal = 35; Nt = (Tfinal-Tinit)/dt;

279

cla;

get(hObject,'Min') and get(hObject,'Max') to determine range of slider

280

plot time = Tinit+get(handles.slider1,'Value')*(Tfinal-Tinit);

281

set(handles.edit1,'String',plot time);

282

PLOT TYPE = get(handles.popupmenu plot type, 'Value');

283
284

switch PLOT TYPE

285

case 2

286

%Pressure vs Dist Section Plots

287

PLOT NUMBER = get(handles.listbox1,'Value');

288

play speed = 1;

289

switch PLOT NUMBER

290

case 1

291

distPlot(handles,time,CALIBR DATA TABLE,'Pressure',PLAY,play speed);

292

case 2

293

distPlot(handles,time,CALIBR DATA TABLE,'Temperature',PLAY,play speed);

294
295

end
end

296
297
298

% --- Executes during object creation, after setting all properties.

299

function slider1 CreateFcn(hObject, eventdata, handles)

300

% Hint: slider controls usually have a light gray background.

301

if isequal(get(hObject,'BackgroundColor'), get(0,'defaultUicontrolBackgroundColor'))

302
303
304

set(hObject,'BackgroundColor',[.9 .9 .9]);
end

189

305
306

function edit1 Callback(hObject, eventdata, handles)

307

% Hints: get(hObject,'String') returns contents of edit1 as text

308

str2double(get(hObject,'String')) returns contents of edit1 as a double

309
310

% --- Executes during object creation, after setting all properties.

311

function edit1 CreateFcn(hObject, eventdata, handles)

312
313

% Hint: edit controls usually have a white background on Windows.

314

315

if ispc && isequal(get(hObject,'BackgroundColor'), get(0,'defaultUicontrolBackgroundColor'))

316
317

See ISPC and COMPUTER.

set(hObject,'BackgroundColor','white');
end

318
319

% --- Executes on selection change in listbox1.

320

function listbox1 Callback(hObject, eventdata, handles)

321

global PLOT ON PLAY data file loc string dt plot time time CALIBR DATA TABLE N time points

322

global line names line units line plot range lower line plot range upper line channel line type

323

% Hints: contents = get(hObject,'String') returns listbox1 contents as cell array

324

325

PLOT TYPE = get(handles.popupmenu plot type, 'Value');

326

cla(handles.axes1)

327

switch PLOT TYPE

328

contents{get(hObject,'Value')} returns selected item from listbox1

case 1

329

line = get(handles.listbox1,'Value');

330

transientPlot(handles,line,time,'Pressure',CALIBR DATA TABLE);

331

case 2

332

%Pressure vs Dist Section Plots

333

PLOT NUMBER = get(handles.listbox1,'Value');

334

play speed = 1;

335

switch PLOT NUMBER

336

case 1

337

distPlot(handles,time,CALIBR DATA TABLE,'Pressure',PLAY,play speed);

338

case 2

339

distPlot(handles,time,CALIBR DATA TABLE,'Temperature',PLAY,play speed);

340

end

341
342

end

343
344

% --- Executes during object creation, after setting all properties.

345

function listbox1 CreateFcn(hObject, eventdata, handles)

346
347

% Hint: listbox controls usually have a white background on Windows.

348

349

if ispc && isequal(get(hObject,'BackgroundColor'), get(0,'defaultUicontrolBackgroundColor'))

350
351

See ISPC and COMPUTER.

set(hObject,'BackgroundColor','white');
end

352
353

function edit dt Callback(hObject, eventdata, handles)

354

% Hints: get(hObject,'String') returns contents of edit dt as text

355

str2double(get(hObject,'String')) returns contents of edit dt as a double

356
357
358

% --- Executes during object creation, after setting all properties.

359

function edit dt CreateFcn(hObject, eventdata, handles)

360

% Hint: edit controls usually have a white background on Windows.

361

See ISPC and COMPUTER.

190

362

if ispc && isequal(get(hObject,'BackgroundColor'), get(0,'defaultUicontrolBackgroundColor'))

363
364

set(hObject,'BackgroundColor','white');
end

365
366
367

% --- Executes during object creation, after setting all properties.

368

function pushbutton update CreateFcn(hObject, eventdata, handles)

369
370

% --- Executes on selection change in popupmenu3.

371

function popupmenu3 Callback(hObject, eventdata, handles)

372

global data file loc string run

373

% Hints: contents = get(hObject,'String') returns popupmenu3 contents as cell array

374

375

runs = get(handles.popupmenu3,'String')

376

run = runs{get(handles.popupmenu3,'Value')}

contents{get(hObject,'Value')} returns selected item from popupmenu3

377

data file loc string = strcat('..\Data Files\',run,'\');

378

pushbutton update Callback(hObject, eventdata, handles)

379

popupmenu plot type Callback(hObject, eventdata, handles)

380

set(handles.text5,'String',data file loc string)

%location of the data files

381
382

% --- Executes during object creation, after setting all properties.

383

function popupmenu3 CreateFcn(hObject, eventdata, handles)

384

if ispc && isequal(get(hObject,'BackgroundColor'), get(0,'defaultUicontrolBackgroundColor'))

385
386

set(hObject,'BackgroundColor','white');
end

387
388

% ------------------------------------------------------------

389

% Read the current directory and sort the names

390

% ------------------------------------------------------------

391

function load listbox(dir path,handles)

392

global data file loc string

393

temppwd = pwd;

394

cd (dir path);

395

dir struct = dir(dir path);

396

[sorted names,sorted index] = sortrows({dir struct.name}');

397

file names = sorted names;

398

is dir = [dir struct.isdir];

399

set(handles.popupmenu3, 'String', file names(3:end));

400

set(handles.text5,'String',data file loc string)

401

cd (temppwd);

402
403
404

% --- Executes on button press in checkbox1.

405

function checkbox1 Callback(hObject, eventdata, handles)

406

% hObject

handle to checkbox1 (see GCBO)

407

% eventdata

reserved - to be defined in a future version of MATLAB

408

% handles

structure with handles and user data (see GUIDATA)

409
410

% Hint: get(hObject,'Value') returns toggle state of checkbox1

B.2

get Line Info.m

191

function [line names,units,line plot range lower,line plot range upper,channel,line type] = get Line Info()

global data file loc string

3
4

[ln,header,raw] = xlsread(strcat(data file loc string,'Line Names.xlsx'),1,'A1:N66');

5
6

line names = raw(2:end,6);

units = raw(2:end,12);

line plot range lower = raw(2:end,13);

line plot range upper = raw(2:end,14);

10

channel = raw(2:end,3);

11

line type = raw(2:end,10);

12
13

for i = length(line names)

14

if isnan(line names{i})

15

line names{i} = '.';

16

end

17

if isnan(units{i})

18

units{i} = '.';

19

end

20

if isnan(line plot range lower{i})

21

line plot range lower{i} = '.';

22

end

23

if isnan(line plot range upper{i})

24

line plot range upper{i} = '.';

25

end

26

if isnan(channel{i})

27

channel{i} = '.';

28

end

29

if isnan(line type{i})

30

line type{i} = '.';

31
32

end
end

B.3

calibrate.m

function val = calibrate(input,time,type)

switch type

3
4
5

case 1
val = input.*100;
case 2

%Type C

input = input*10;

B(1) = 74.20821;

B(2)= -4.345242;

10

B(3)= 0.5411627;

11

B(4)= -4.8762064E-02;

12

B(5)= 3.0525550E-03;

13

B(6)= -1.2473883E-04;

14

B(7)= 3.1673633E-06;

15

B(8)= -4.5258943E-08;

16

B(9)= 2.7811453E-10;

17

val=0;

192

18

for i = 1:9

19
20
21

val = val+B(i)*input.(i);
end
case 3

22

%Type K

23

input = input*10;

24

B(1)=

25

B(2)= -0.2482746;

26

B(3)=

27

B(4)= -6.2740748E-03;

28

B(5)=

29

B(6)= -1.0039074E-05;

30

B(7)=

31

B(8)= -1.6374382E-09;

32

B(9)=

24.81809;

6.0291328E-02;

3.3348822E-04;

1.7449725E-07;

6.4360188E-12;

33
34

val=0;

35

for i = 1:9

36
37
38

val = val+B(i)*input.(i);
end
case 4

39

%Type R (Mattick calibration)

40

input = input*2.5;

41

B(1)=

173.3246

42

B(2)=

-36.98014 ;

43

B(3)=

44

B(4)=

-1.948491;

45

B(5)=

0.2250111 ;

46

B(6)= -1.6329067E-02;

47

B(7)=

48

B(8)= -1.7749791E-05;

49

B(9)=

50

val=0;

51

for i = 1:9

52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72

10.60782 ;

7.2203868E-04;

1.8577632E-07;

val = val+B(i).*input.(i);
end
case 5
%H2 Manifold Pressure (OMEGADYNE CALIBRATION SHEETS)
val = 6.89475729*(14.6959488+interp1([0.016,2.525,5.018],[0,500,1000],input,'linear','extrap'));
case 6
%O2 Manifold Pressure (OMEGADYNE CALIBRATION SHEETS)
val = 6.89475729*(14.6959488+interp1([0.016,2.525,5.017],[0,500,1000],input,'linear','extrap'));
case 7
%H2 Tank Pressure (OMEGADYNE CALIBRATION SHEETS)
val = 6.89475729*(14.6959488+interp1([0.039,2.54,5.05],[0,1500,3000],input,'linear','extrap'));
case 8
%O2 Tank Pressure (OMEGADYNE CALIBRATION SHEETS)
val = 6.89475729*(14.6959488+interp1([0.015,2.525,5.023],[0,1500,3000],input,'linear','extrap'));
case 9
%CH4 Ballast Pressure (OMEGADYNE CALIBRATION SHEETS)
val = 6.89475729*(14.6959488+interp1([0.011,2.521,5.023],[0,500,1000],input,'linear','extrap'));
case 10
%CH4 Manifold Pressure (OMEGADYNE CALIBRATION SHEETS)
val = 6.89475729*(14.6959488+interp1([0.003,2.508,4.989],[0,250,500],input,'linear','extrap'));
case 11

73

%Optical Temperature Sensor (in Kelvin)

74

val = 53.969*input.4 + 356.16*input.3 + 930.48*input.2 + 1385.5*input + 1588.6 - 273.15;

193

75

case 12

76

%Water spraybar Mass / volt

77

%Data smoothing in 3 steps:

78

% 1. Use the smooth function to apply a 15 span

79

% Savitzky-Golay filter on the data, removing higher frequency noise.

80

% 2. Alias and anti-alias data via interpoltion - Select every 5th

81

% data point (by using the J vector) and interpolate with a cubic

82

% spline to remove mid frequency noise, preserve shape and avoid

83

% Gibb's phenomena

84

% 3. Aias and anti-alias again using the J2 vector and selecting

85

% every 15th data point to remove small remaining bumps so that the

86

% dM/dt plot is smoother.

87

N = length(input);

88
89

%Polynomial fit

90

P = polyfit(time,input,4);

91

input1 = polyval(P,time);

92
93

%Sliced Domain

94

[t1,i1] = min(abs(time-1));

95

[t2,i2] = min(abs(time-15));

96

[t3,i3] = min(abs(time-19.8));

%end of the high flow rate phase

97

[t4,i4] = min(abs(time-30));

%end of the low flow rate phase

%start of the low flow rate phase


%start of the high flow rate phase

98
99

istart=[1,i1+1,i2+1,i3+1,i4+1];

100

end=[i1,i2,i3,i4,N];

101

input2 = [];

102

alias small = 5; alias big = 10; smooth span = 240;

103

for i = 1:length(istart)

104

T = time(istart(i):alias small:end(i));

105

J = istart(i):alias small:end(i);

106

input smooth = zeros(N,1);

107

input alias = zeros(N,1);

108

input smooth(istart(i):end(i)) = interp1(T,smooth(T,input(J),smooth span,'rlowess'),time(istart(i):end(i)),'pchip','extrap')

109
110

T2 = time(istart(i):alias big:end(i));

111

J2 = istart(i):alias big:end(i);

112

input alias(istart(i):end(i)) = interp1(T2,input smooth(J2),time(istart(i):end(i)),'pchip');

113

input2 = [input2;input alias(istart(i):end(i))];

114

end

115
116

%Custom smoothing and aliasing technique

117

J = 1:5:N; J2 = 1:10:N;

118

input3 = interp1(J,smooth(input(J),14,'rlowess'),1:N,'pchip');

119

input3 = interp1(J2,input3(J2),1:N,'pchip');

120
121

%Linear regression smoothing

122

input4 = smooth(input,60,'rlowess');

123

input4 = smooth(input4,60,'rlowess');

124
125

val = 21.991*input2 + 1.0682;

126

case 13

127

%Injector Calibration

128

val = input;

129

otherwise

130
131

val = input;
end

194

132

end

B.4

distPlot.m

function distPlot(handles,time,CALIBR DATA TABLE,PorT,PLAY,play speed)

global PLOT ON plot time

global line plot range lower line plot range upper

if PLOT ON

5
6
7

switch PorT
case 'Pressure'
%Pressure vs Dist Section Plots

plot lines = [33,34,9:11,24,12:23,25];

section = [1:8,3:5,7:18,6,19,26:32,1,2]; %downstream distance of the sensor (index i corresponds to line i, value at that in

10

%Section Pressure Transducers in the order that they are plotted

dist = section(plot lines); %select distances

11

dist index range = 1:length(dist);

12

animation speed = play speed;

13
14

j = 0;

15

for i = plot lines %for each line corresponding to sensor location, interpolate at time plot time

16

j = j+1;

17

pressure data(j) = interp1(time,CALIBR DATA TABLE(:,i),plot time,'linear','extrap');

18

end

19

hold(handles.axes1,'on')

20

plot(handles.axes1,dist(1:2),pressure data(1:2),'sr')

21

plot(handles.axes1,dist(3:5),pressure data(3:5),'or','MarkerFaceColor',[1 0 0])

22

plot(handles.axes1,dist(6)

23

plot(handles.axes1,dist(7:11)

24

plot(handles.axes1,dist(12:14),pressure data(12:14),'ok','MarkerFaceColor',[0 0 0])

25

plot(handles.axes1,dist(15:17),pressure data(15:17),'og','MarkerFaceColor',[0 1 0])

%Reactor

26

plot(handles.axes1,dist(18)

,pressure data(18),'oc','MarkerFaceColor',[0 1 1])

%Diagnostic

27

plot(handles.axes1,dist(19)

,pressure data(19),'ko')

28

legend('Pebble Bed Heater',...

29

'Combustor',...

30

'Dummy',...

31

'Mixer',...

32

'Transition',...

33

'Reactor',...

34

'Diagnostic',...

%PBH
%Combustor

,pressure data(6),'ob','MarkerFaceColor',[0 0 1])

%Dummy

,pressure data(7:11),'om','MarkerFaceColor',[1 0 1])

%Mixer
%Transition

%Dump Tank o

35

'Dump Tank')

36

plot(handles.axes1,dist(19)

37

grid on

38

xlim([dist(1)-1,dist(end)+1]);

39

ylim([line plot range lower{plot lines(1)},line plot range upper{plot lines(1)}])

40

xlabel('Station'); ylabel('Pressure (kPa)');

41

title(strcat('Pressure vs. Downstream Distance at t=',num2str(plot time,'%11.4f'),'s'))

42

hold(handles.axes1,'off')

,pressure data(19),'kx')

43
44

case 'Temperature'

45

%Temperature vs Dist Section Plots

46

% 1

48 pbh bottom (K)

47

% 2

45 pbh top (K)

48

% 3

57 C3 (C)

%Dump Tank x

195

49

% 4

56 C3 Liner (K)

50

% 5

54 D1 (K)

51

% 6

50 D1 Liner (K)

52

% 7

51 M5 (K)

53

% 8

55 M3 Liner (K)

54

% 9

52 R1 (K)

55

% 10 58 R2 outer flow (C)

56

% 11 59 R2 core flow (C)

57

% 12 53 Diag (K)

58

plot lines = [48,45,57,56,54,50,51,55,52,58,59,53];

59

section = [1:44,2,46:47,1,49,5,7,9,12,5,7,3,3,10,10]; %downstream distance of the sensor (index i corresponds to line i, val

60

dist = section(plot lines); %select distances

61

dist index range = 1:length(dist);

62

animation speed = play speed;

%Section Pressure Transducers in the order that they are plotted

63
64

j = 0;

65

for i = plot lines %for each line corresponding to sensor location, interpolate at time plot time

66

j = j+1;

67

temperature data(j) = interp1(time,CALIBR DATA TABLE(:,i),plot time,'linear','extrap');

68

end

69

hold(handles.axes1,'on')

70

plot(handles.axes1,dist(1:2),temperature data(1:2),'sr')

71

plot(handles.axes1,dist(3),temperature data(3),'or','MarkerFaceColor',[1 0 0])

%Combustor

72

plot(handles.axes1,dist(4),temperature data(4),'pr','MarkerFaceColor',[1 1 1])

%Combustor Liner

73

plot(handles.axes1,dist(5)

,temperature data(5),'pb','MarkerFaceColor',[0 0 1])

74

plot(handles.axes1,dist(6)

,temperature data(6),'pb','MarkerFaceColor',[1 1 1])

75

plot(handles.axes1,dist(7)

,temperature data(7),'pm','MarkerFaceColor',[1 0 1])

76

plot(handles.axes1,dist(8)

,temperature data(8),'pm','MarkerFaceColor',[1 1 1])

77

plot(handles.axes1,dist(9),temperature data(9),'pg','MarkerFaceColor',[1 1 1])

78

plot(handles.axes1,dist(10:11),temperature data(10:11),'og','MarkerFaceColor',[0 1 0])

%Reactor C

79

plot(handles.axes1,dist(12)

%Diagnostic

80

grid on

81

xlim([dist(1)-1,dist(end)+1]);

82

ylim([line plot range lower{plot lines(1)+1},line plot range upper{plot lines(1)+1}])

83

xlabel('Dist'); ylabel('Temperature (\circC)');

84

title(strcat('Temperature vs. Downstream Distance at t=',num2str(plot time,'%11.4f'),'s'))

85

%Mixer Liner

,temperature data(12),'pc','MarkerFaceColor',[0 1 1])

end

transientPlot.m

function transientPlot(handles,line,time,PorT,CALIBR DATA TABLE)

global PLOT ON N time points

global line names line units line plot range lower line plot range upper line channel line type

4
5
6
7

if PLOT ON
switch PorT
case 'Pressure'

time index range = 1:N time points;

conversion = 1; offset = 0;

10

%Dummy Liner

%Reactor K

end

B.5

%Dummy

%Mixer

hold(handles.axes1,'off')

86
87

%PBH

if strcmp(line units(line),'Pressure (psi)')

196

11

conversion = 1/6.89475729;

12

offset = 14.7;

13

units = 'Pressure (psig)';

14

elseif strcmp(line units(line),'Pressure (psig)')

15

conversion = 1/6.89475729;

16

offset = 14.7;

17

units = 'Pressure (psig)';

18

elseif strcmp(line units(line),'Pressure (psia)')

19

conversion = 1/6.89475729;

20

offset = 1;

21

units = 'Pressure (psia)';

22

else

23

units = line units(line);

24

end

25

line,size(time index range),size(CALIBR DATA TABLE)

26

plot(handles.axes1,time(time index range),CALIBR DATA TABLE(time index range,line).*conversion-offset,'.-k')

27
28

% Attempt to set the domain and range if supplied in

29

% Line Names.xls.

30

try

31

xlim([time(1)-1,time(end)+1]); ylim([line plot range lower{line},line plot range upper{line}])

32

catch me

33

end

34

grid on

35

xlabel('time (s)'); ylabel(units);

36

title(strcat(line names(line),'

37

case 'temperature'

38
39
40

end
end

[',line type(line),'] - Channel:',line channel(line),' Line:',num2str(line-1)))

197

Appendix C
SPRAY BAR MASS FLOW RATE - NOISE REDUCTION
TECHNIQUES

During a run, on the order of 2-5 kg of water is sprayed into the test section over 35 s,
representing a very small change in mass when compared with the range of the electronic
scale. Hence, signal noise is found to be high, generally about

0.5kg peak-to-peak. This

is exemplified in Figure C.1 which illustrates measurements acquired during a run with two
phases, both low and high water flow rate. Note that, in this example, water injection
timing is not the same as that used in nominal reactor conditions; the low flow rate phase
runs from 1.0 to 15.0 seconds and from 20.0 to 30.0 seconds whilst the high flow rate phase
runs from 15.0 to 20.0 seconds.

Figure C.1: Noise in the water mass vs. time plot

Ideally, the curve of water ballast tank mass vs. time is expected to contain discontinuities

198

in gradient when water mass flow is instantaneously changes from one phase to another. In
reality a small transient exists due to the time required to fill manifolds and it smoothes the
transition between each phase. The duration of the transient is expected to be small due
the incompressible nature of water flow and high flow rates involved. Consequently, two
distinct gradients are expected, and this has been approximated by the straight solid lines
sketched in the figure. The use of straight lines implies a constant mass flow rate assumption
during each of the low and high flow rate phases. This corresponds to a constant driving
pressure from the inert gas pressure inside the ballast tank and relatively invariant back
pressure inside the reactor during a run. This a priori knowledge of timing and constant
flow rates is to be made use of in the analysis that follows. Water mass vs. time trends are
approximated with a graphical method in Figure C.2. From the plot, low and high flow rate
phases are 0.12 kg/s and 0.16 kg/s, respectively. Mass flow rate approximations sketched
in Figure C.3.

Figure C.2: Noise in the water mass vs. time plot with trendline

199

Figure C.3: Graphically determined mass flow rate

The quantity of interest from the water mass measurements is the mass flow rate (m
=

dm
dt )

because it is insightful to know if sufficient quench water is entering the system to maintain
effective cooling. Although the level of cooling may be ascertained, to an extent, by observing flow temperatures downstream of the spray bar, the approximate mass flow rate is also
useful for hydrogen balances in the pyrolysis chemistry. Theoretical calculation of quench
water mass flow rate is complicated by the inability to directly examine flow conditions
inside the sealed reactor system for use modeling the flow. Numerical simulation of the
spray bar system was not practical from a time and computational resource standpoint as
the domain complexity of a two-phase-with-injection computational fluid dynamics model
would exceed available computer memory. Instead, an experimental approach for determining mass flow has been employed by measuring water tank mass vs. time, and subsequent
differentiation of this curve.
It is well known that differentiation of a signal is highly sensitive to noise, and in general,

200

data smoothing techniques are employed to artificially eliminate noise thus allowing more
sensible derivatives to be calculated. In Figure C.1 calculated mass flow rate before noise
reduction clearly shows how uninformative this data can be without noise reduction.

Figure C.4: Mass flow rate vs. time before noise reduction

Several noise reduction techniques were examined to smooth water mass vs. time data
effectively. Use of a variety of Fourier decompositions, frequency domain weightings and
high frequency cutoffs proved ineffective as spurious Gibbs phenomena were unavoidable.
Polynomial fits of lower order failed to preserve the curve shape as seen in Figures C.5.
Higher order polynomial fits, seen in Figures C.6, generated highly spurious oscillations near
the start and end of a run.
Key data in the middle of the run was of the correct order of magnitude, yet it was highly
smoothed and failed to capture the distinct yet subtle changes in slope that are expected
to occur at the transitions from low to high water mass flow rate, and vice versa.
Local averaging, implemented using MATLAB smooth function, proved more successful

201

Figure C.5: Mass flow rate vs. time: Lower Order Fits

Figure C.6: Mass flow rate vs. time: High Order Fits

202

at noise reduction, though only when combined with a priori knowledge of the timing for
low and high flow rate phase. Simple moving average techniques are known to be effective
for functions with zero second derivatives. However, non-zero second derivative functions
experience smoothing and reduction of local maxima and minima which are to be preserved
in this case since they quantify the magnitude of the mass flow rate. More sophisticated
local averaging methods were examined. The optimal smoothing technique was found to
be local regression smoothing using robust weighted linear least squares on a first degree
polynomial model. In MATLAB, the function call is of the form smooth(x,y,rlowess,1),
henceforth referred to as rlowess smoothing. Regression weights are given by a tricube
function given by Equation C.0.1:
wi = (1 |

x xi 3 3
| )
d(x)

(C.0.1)

x is the predictor value associated with the response value to be smoothed, xi are the nearest
neighbors of x as defined by the span, and d(x) is the distance along the abscissa from x
to the most distant predictor value within the span. Used alone, however, the results of
rlowess smoothing in Figure C.7 are not quite satisfactory as they do not capture the low
and high flow rate phases distinctly.
Performing separate smoothing in carefully chosen time segments shows a marked improvement. The full procedure is as follows. The time domain is sliced into five segments, one
for each phase, as shown in Table C.1.
Time Segment

Start Time (s)

End Time (s)

Water Off

-5.0

2.0

Low Flow Rate

2.0

15.2

High Flow Rate

15.2

19.8

Low Flow Rate

19.8

30

Water Off

30

39.9

Table C.1: Time domain slicing

In this example, rlowess smoothing is then performed in each segment using a span of 120

203

Figure C.7: Mass flow rate vs. time: rlowess Fits

points. This span represents approximately one third of the overall time domain. Figure C.8
illustrates the results.
Slicing the time domain has clearly allowed distinct changes in gradient to remain, but more
importantly, the slopes of the curve have been revealed and are readily used to calculate
the mass flow rates, as seen in Figure C.9.
The procedure for calculating the mass flow rate is to take the finite difference approximation
to the first derivative by using MATLABs 2nd order accurate gradient() function. This
procedure is complicated by the act of slicing of the time domain since large discontinuities
are present in the mass vs. time curve. These discontinuities are eliminated by an additional
rlowess smoothing step using a minimal span of only 10 points.

204

Figure C.8: Water Tank Mass vs. time: Sliced rlowess Fits

205

Figure C.9: Water Tank Mass vs. time: Sliced rlowess Fits

206

Appendix D
MATLAB CODE FOR PSEUDO 1-D MODEL, BASIC FEEDSTOCK
MIXING AND NOZZLE HEAT TRANSFER

The file main.m is the main processing file and makes calls to nine functions.
1. calcCombustorPenetration.m estimates the penetration of the jet from the H2 / O2 injectors
in the combustion section, ignoring impingements.
2. calcNozzleFlowProperties.m takes a given nozzle geometry and uses the 1-dimensional
isentropic flow equations to calculate flow properties
3. getArea.m retrieves the nozzle profile x,r coordinates from a database of different
nozzle shapes.
4. getCH4Viscosity.m calculates the viscosity of methane from tabulated values. Lookup
is based on temperature and pressure.
5. getGasProperties.m determines thermodynamic properties of steam.
6. getMach.m implicitly solves for Mach number from the classic area ratio equation for
isentropic flow.
7. getNozzleMaterialConstants.m takes a material type (e.g. titanium) as an input and
returns heat transfer material constants.
8. nozzleHeatTransfer.m performs the numerical finite difference simulation given an
initial condition and a nozzle temperature vs. time profile.
9. plotCrossSectionalMixing.m rough calculations for lateral spread of jets.
Code for each is presented below:

207

D.1

main.m

%% MIXER SIMULATION & HEAT TRANSFER SIMULATION

% AUTHOR: Robert Cerff

% DATE: 02/09/2009

% This code simulates the mixing section of the shock wave reactor. It

% produces estimates for the temperature, pressure, density, momentum

% flux profiles as well as other parameters. Estimates are based on

% assumptions such as, minimual heat loss, no flow losses, flow separation

% or significant boundary layer effects, hence allowing the flow to be

10

% assumed adiabatic and isentropic.

11

12

% VERSIONS:

13

% v1.0 - H2+O2 Combustion, Mass flow rates, Nozzle Property Profiles,

14

% Penetration Height, Injector Sizing

15

16

% v1.1-v1.3 - Heat Transfer Finite Difference Code

17

% v1.4 - Conical Nozzle changed to Bell Nozzle

18

% v1.5a - Refactoring into functions. Heat Transfer Function

19

% v1.5b - Refactoring: Steam Pressure/Temp Profile Function

20

% v1.6 - Nozzle Temperature profile in 2 stages.

21

% v1.7 - UOP Updates:

22

% v1.8 - Combustor Penetration Heights

23

% v1.9 - Required duct expansion after injection and mixing

24
25

clc; clear all; close all; format short g

26

global PLOT ON HEAT TRANSFER ON NOZZLE TYPE

27

global in2m m2in psi2Pa Pa2psi Patm Pbar

28

global G R bath G P0 bath G T0 bath G Cp bath G Cp H2O G T0 bath heat loss

29

global G Cp mix

30
31

PLOT ON = 0;

32

HEAT TRANSFER ON = 0;

33

NOZZLE TYPE = 'cfd';

34

in2m = 2.54e-2;

35

m2in = 1/2.54e-2;

36

Patm = 101325;

37

Pbar = 1e5;

38

psi2Pa = 1/14.7*Patm;

39

Pa2psi = 14.7*Patm;

40
41

%% INCOMING GAS PROPERTIES

42

% CALCULATIONS FOR H20 AND H2 ENTERING THE STEAM MIXER

43
44

%COMBUSTION REACTANTS (per second basis)-

45

M H2 = 2;

%molecular weight (g/mol)

46

M O2 = 32;

%molecular weight (g/mol)

47

M H2O = 18; %molecular weight (g/mol)

48

M H = 1;

%molecular weight (g/mol)

49

M O = 16;

%molecular weight (g/mol)

50

M OH = 17;

%molecular weight (g/mol)

51

M CH4 = 16; %molecular weight (g/mol)

52

REACT RN H2O H2 = 7.16/2.5; %ratio of H2O moles to H2 moles

208

53

REACT RN H2 O2 = 2.5/1;

%ratio of H2 moles to O2 moles

54

PROD RN CH4 H2O = 1/3;

%ratio of CH4 moles to H2O moles

55

mdot H2O = 0.2825;

56

REACT N H2 = (mdot H2O/M H2O)/REACT RN H2O H2 *1000;

%massflow (kg/s) rate of steam from PBH

57

REACT N O2 = REACT N H2/REACT RN H2 O2;

58

mdot H2 = REACT N H2 * M H2/1000

59

mdot O2 = REACT N O2 * M O2/1000

60

mdot comb = mdot H2+mdot O2

61

mdot bath = mdot comb+mdot H2O

%massflow rate of H2 and O2 in combustor


%massflow rate of bath into mixer

62
63

%COMBUSTION PRODUCTS FROM GASEQ @ PROD TEMP = 2300K (per second basis)-

64

PROD N H2O = 0.01536+0.00436%5.1826;

65

PROD N H2 = 2.749e-4+0.00108%1.56992;

66

PROD N total = PROD N H2O + PROD N H2;

67

mdot CH4 = mdot bath* PROD RN CH4 H2O * M CH4/M H2O %massflow of CH4

68

mdot total = mdot bath+mdot CH4

%massflow of the total reacted system

69
70

%MIXING OF COMBUSTION PRODUCTS @ 3167K AND STEAM @ 1400K

71

Caa=32; Cbb=0.1923e-2; Ccc=1.055e-5; Cdd=-3.595e-9;

72

T comb = 3192;

%combustion adiabatic flame temp from GASEQ = 3167K

73

T H2O = 1400;

%temperature of steam from PBH

74

Cp comb = 3562.71;

%specific heat (J/kg.K) from GASEQ

75

G Cp H2O =(Caa+Cbb* T H2O+Ccc* T H2O2+Cdd* T H2O3)*1000/M H2O %specific heat (J/kg.K) Incropera & Dewitt based on H2O @ 1400K

76

Tbath = 2300; %2037.8

%Cp constants for steam

77

G Cp bath = (Caa+Cbb*Tbath+Ccc*Tbath2+Cdd*Tbath3)*1000/M H2O %specific heat (J/kg.K) Incropera & Dewitt based on H2O @ 2200K

78

CPP = (mdot comb* Cp comb* T comb+mdot H2O* Cp comb * T H2O)/(mdot bath*Tbath)

79

G Cp bath = (mdot comb*3329*2300+mdot H2O*2939*2300)/(mdot bath*Tbath)

80

%STEAM GAS PROPERTIES ENTERING MIXER

81

Mhat bath = (M H2O * PROD N H2O + M H2 * PROD N H2)/PROD N total;

82

Mhat bath = M H2O;

83

Ru = 8314;

84

G R bath = Ru/Mhat bath;

85

G T0 bath heat loss = 120;

86

G T0 bath = 2300-G T0 bath heat loss;%(mdot comb*Cp comb* T comb+mdot H2O* G Cp H2O * T H2O)/(mdot bath* G Cp bath)

87

G P0 bath = 2.39*Pbar;

88

rho0 in = G P0 bath/(G R bath * G T0 bath);

89
90

%% STEAM MIXER FLOW PROPERTIES

91

% ISENTROPIC APPROXIMATIONS ARE USED TO GENERATE FLOW PROPERTY PROFILES

92

% INSIDE THE SUPERSONIC EXPANDER.

93

% STAGE 1 (S1): 1400K steam only from Pebble Bed Heater

94

% STAGE 2 (S2): 2100K steam + H2 from combustor and Pebble Bed Heater.

95
96

%NOZZLE PROPERTIES

97

D spool inlet = 3.5*in2m;

98

D spool outlet = 3.0*in2m;

99

z throat = 1.289*in2m;

%inches

100

z nozzle length = 6.0*in2m;

101

NQ = 100; %number of axial points in the grid

102

dz = z nozzle length/(NQ-1);

103

z = (0:dz:z nozzle length);

104
105

%INCOMING GAS MIXTURE PROPERTIES

106

[g S1,gm S1,gp S1,mu S1,R S1,P0 S1,T0 S1,Cp S1] = getGasProperties('precombusted S1');

107

[g S2,gm S2,gp S2,mu bath S2,R bath S2,P0 bath S2,T0 bath S2,Cp bath S2] = getGasProperties('bath S2');

108
109

%CALCULATE THROAT DIAMETER BASED ON CHOKED FLOW IN STAGE 2

209

110

P star bath S2 = P0 bath S2/(gp S2/2)(g S2/gm S2);

111

T star bath S2 = T0 bath S2/(gp S2/2);

112

A star = mdot bath/P star bath S2 *sqrt(R bath S2 * T star bath S2/g S2);

113

D star = sqrt(4* A star/pi);

114

disp(['Throat diameter: ' num2str(D star/2.54e-2) 'in'])

115
116

%CALCULATE STAGE 1 MASS FLOW RATE

117

P star S1 = P0 S1/(gp S1/2)(g S1/gm S1);

118

T star S1 = T0 S1/(gp S1/2);

119

mdot S1 = A star * P star S1/sqrt(R S1 * T star S1/g S1)

120

mdot S2 = mdot total

121
122

%CALCULATE NOZZLE GEOMETRY

123

[A nozzle,D nozzle,z,nozzle dir cosine,throat index z,z throat] = getArea(z,z throat,D star,z nozzle length,D spool inlet,D spool outlet

124

options=optimset('display','off');

125

Minit S2 = getMach(0.1, options, g S2, A nozzle(1)/A star); %Incoming Mach Number

%get the nozzle area

at position z from geometric parameters


%sets options for fsolve

126
127

%CALCULATE NOZZLE FLOW PROFILES

128

[M S1,a S1,P S1,T S1,rho S1,u S1,q S1,Re S1,mu CH4 S1] = calcNozzleFlowProperties(z,T0 S1,P0 S1,A nozzle,D nozzle,A star,g S1,z throat,R

129

[M S2,a S2,P S2,T S2,rho S2,u S2,q S2,Re S2,mu CH4 S2] = calcNozzleFlowProperties(z,T0 bath S2,P0 bath S2,A nozzle,D nozzle,A star,g S2,

130
131

%CH4 MIXING

132

T CH4 = 850;

133

Cp CH4 = 1e3*(-1.46e-07*(T CH4)2 + 3.34e-03* T CH4 + 1.29); %Excel Spreadsheet CH4 gamma at 2100K.xlsx

134
135

%METHANE INJECTION TOTAL TEMPERATURES WITH HEAT LOSS

136

Qdot mixing INCFD = -26885

137

mdot H2O mixing IN = mdot bath

138

mdot CH4 mixing IN = mdot CH4

139

Cp H2O mixing IN = G Cp bath

140

Cp CH4 mixing IN = 66.5/16e-3 %NIST Chemistry Web Book - Methane at 850K is 66.5J/mol.K

141

T0 H2O mixing IN = G T0 bath

142

T0 CH4 mixing IN = 850

143

mdot mixing OUT = mdot total

144

Cp mixing OUT = 3130

145

T0 mixing = (mdot H2O mixing IN * Cp H2O mixing IN * T0 H2O mixing IN...

146

+mdot CH4 mixing IN * Cp CH4 mixing IN * T0 CH4 mixing IN+Qdot mixing INCFD)...

147

/(mdot mixing OUT * Cp mixing OUT)

148
149
150

% DUCT EXPANSION AFTER INJECTION OF METHANE

151

[g exp,gm exp,gp exp,mu exp,R exp,P0 exp,T0 exp,Cp exp] = getGasProperties('CH4-H2O Mix');

152

M exp = 1.0;

153

P exp = P0 exp/(1+gm exp/2* M exp2)(g exp/gm exp);

154

% P exp = 0.5*Pbar;

155

T0 exp

156

T exp = T0 exp/(1+gm exp/2* M exp2)

157

RT exp = R exp * T exp;

158

u expansion = M exp *sqrt(g exp * RT exp);

159

rho expansion = P exp/(RT exp);

160

A exp = mdot total/(rho expansion*u expansion);

161

D exp inches = sqrt(4* A exp/pi)*m2in

162
163

%% COMBUSTOR PENETRATION HEIGHT

164

[CombustorPenetrationHeight(:,1),CombustorPenetrationHeight(:,2)] = ...

165

calcCombustorPenetration();

210

166
167

%% INJECTOR SIZING

168

% DETERMINATION OF REQUIRED INJECTION PRESSURE, INJECTOR THROAT DIAMETERS

169

% ETC.

170

PRIMARY INJECTOR = 1;

171

WITH FIN = 1;

172

scaled fin height = 1.66;

173

gJ = 1.2;

174

gJp = gJ+1;

175

gJm = gJ-1;

176

T0 jet = T CH4;

177

P0 jet = 150*psi2Pa;

178

num injectors = 4;

%fin increases penetration of the primary jet by this factor

%feedstock temperature

179

num sec injectors = 0;

180

num primary injectors = WITH FIN;

181

if PRIMARY INJECTOR

%value is set below

182

num sec injectors = num injectors-1;

183

injector spacing = 2*pi* D spool inlet/2/(num sec injectors+1);

184

if WITH FIN

185

primary jet scale = 2.15*1/3*num sec injectors;

186

else

187

primary jet scale = 3/4*num sec injectors;

188

end

189

mdot primary jet = primary jet scale*(mdot total-mdot bath)/(num sec injectors+1);

190

mdot jet = (mdot total-mdot bath-mdot primary jet)/num sec injectors;

191
192

R jet = Ru/M CH4;

193

P star jet = P0 jet/(gJp/2)(gJ/gJm);

194

T star jet = T0 jet/(gJp/2);

195

rho star jet = P star jet/(R jet * T star jet);

196

a star jet = sqrt(gJ* R jet * T star jet);

197

A star jet = mdot jet/P star jet *sqrt(R jet * T star jet/gJ);

198

D throat jet = sqrt(4* A star jet/pi);

199

A star primary jet = mdot primary jet/P star jet *sqrt(R jet * T star jet/gJ);

200

D throat primary jet = sqrt(4* A star primary jet/pi);

201

fprintf('\n---INJECTOR PROPERTIES---');

202

fprintf(['\nAeroramp injector diameters: ' num2str(D throat primary jet/2.54e-2/sqrt(2)) 'in\n'])

203

fprintf(['Secondary injector diameter: ' num2str(D throat jet/2.54e-2) 'in\n'])

204
205

fprintf(['\nInjector spacing: ' num2str(injector spacing/2.54e-2) 'in\n'])


else

206

injector spacing = 2*pi* D spool inlet/2/num injectors;

207

mdot jet = (mdot total-mdot bath)/num injectors;

208

mdot primary jet = mdot jet;

209

R jet = Ru/M CH4;

210

P star jet = P0 jet/(gJp/2)(gJ/gJm);

211

T star jet = T0 jet/(gJp/2);

212

rho star jet = P star jet/(R jet * T star jet);

213

a star jet = sqrt(gJ* R jet * T star jet);

214

A star jet = mdot jet/P star jet *sqrt(R jet * T star jet/gJ);

215

D throat jet = sqrt(4* A star jet/pi);

216

D throat primary jet = D throat jet;

217

fprintf(['Injector diameter: ' num2str(D throat jet/2.54e-2) 'in'])

218
219

fprintf(['\nInjector spacing: ' num2str(injector spacing/2.54e-2) 'in\n'])


end

220
221

fprintf(['\nPrimary injector mass flow: ' num2str(num primary injectors) ' x '

222

fprintf(['Secondary injector mass flow: ' num2str(num sec injectors) ' x ' num2str(mdot jet) 'kg/s\n'])

num2str(mdot primary jet) 'kg/s\n'])

211

223
224

% MOMENTUM FLUX RATIO

225

rhoJ = rho star jet;

226

rhoInf = rho S2;

227

uJ = a star jet;

228

uInf = u S2;

229

J ratio = sqrt(rhoJ./rhoInf).*uJ./uInf;

230
231

%% MIXER PENETRATION HEIGHT

232

% ANALYTICAL EXPRESSION FROM

233

% Rosen, Spaid, Zukoski: Secondary injection of gases into supersonic flow

234

% SHOWING PENETRATION HEIGHT VS MACH NUMBER AND STAGNATION PRESSURE RATIO

235

Minf = M S2;

236

Pinf = P S2;

237

gInf = g S2;

238

Cp star = real(2./(g S2 *Minf.2).*((gp S2/2*Minf.2).(g S2/gm S2)...

239
240

.*(gp S2./(2* g S2 *Minf.2-g S2+1)).(1/gm S2)-1));


if PRIMARY INJECTOR

241

dcp = D throat primary jet; %prim. injector diameter * disch coefficient

242
243

dc = D throat jet; %injector diameter * discharge coefficient


else

244

dcp = D throat jet; %prim. injector diameter * disch coefficient

245

dc = D throat jet; %injector diameter * discharge coefficient

246

end

247

h dc = ((1./Minf).*(P0 jet./Pinf*gJ/gInf*2./Cp star).0.5...

248

.*(2./gJm*(2./gJp).(gJp/gJm).*(1-(Pinf./P0 jet).(gJm/gJ))).0.25);

249
250

zerow = zeros(1,length(h dc));

251

zerow(end-1:end) = [0,1];

252

if WITH FIN

253
254

hp = real(h dc)*dcp*scaled fin height;


else

255

hp = real(h dc)*dcp;

256

end

257

h = real(h dc)*dc;

258

h = zerow.*h;

259

hp = zerow.*hp;

260

rc = hp/3;

261

Acp = pi*rc.2;

262

w = 1.3*h;

263

Ac = num sec injectors*h.*w;

264

Aduct = pi*(D spool outlet).2/4;

265

coverage = (Acp(end)+Ac(end))/Aduct

266
267
268

fprintf(['\nPrimary penetration height: ' num2str(hp(end)/2.54e-2) 'in\n'])

269

fprintf(['Secondary penetration height: ' num2str(h(end)/2.54e-2) 'in\n'])

270

fprintf(['\nMomentum Flux Ratio J = ' num2str(J ratio(end))])

271
272
273

if PLOT ON

274

figure()

275

Dmax = max(D nozzle);

276

hold on

277

plot(z(end-1:end)*m2in,(-D nozzle(end-1:end)./2+Dmax/2+hp(end-1:end))*m2in,'ro-','LineWidth',2);

278

plot(z(end-1:end)*m2in,(D nozzle(end-1:end)./2+Dmax/2-h(end-1:end))*m2in,'bo-');

279

plot(z*m2in,(D nozzle./2+Dmax/2)*m2in,'k--')

212

280

legend('Primary Injector Penetration',...

281

'Secondary Injector Penetration',...

282

'Nozzle Wall','Location','SW');

283

plot(z*m2in,(-D nozzle./2+Dmax/2)*m2in,'k--')

284

zlabel('z (in)'); ylabel('h (in)'); xlim([0,z nozzle length*m2in])

285

ylim([0,max(D nozzle)*m2in]); title('Penetration Height (in)')

286

axis equal

287

plotCrossSectionalMixing(PRIMARY INJECTOR,num injectors,hp(end),...

288
289

h(end),D throat primary jet,D throat jet);


end

290
291

%% FINITE DIFFERENCE HEAT TRANSFER ANALYSIS

292

% MODELS TRANSIENT 2D CONDUCTION IN THE NOZZLE WALL VIA A FINITE DIFFERENCE

293

% DISCRETIZATION OF THE HEAT EQUATION IN POLAR COORDINATES.

294

% BOUNDARY CONDITIONS:

295

% INNER WALL TEMPERATURE IS DEPENDENT ON CONVECTIVE HEAT TRANSFER

296

% RATE THAT IS CALCULATED USING BARTZ'S EQUATION.

297

% OUTER WALL TEMPERTURE IS DEPENDENT ON RADIATIVE HEAT TRANSFER RATE.

298

% NOZZLE ENDWALLS ARE FIXED AT T=400K

299

if HEAT TRANSFER ON

300

NP = 30; %number of radial points in the grid

301

nozzle material = 'Stainless316';

302

TT init = 0;

303

final time S1 = 10;

304

dt1 = 0.0005;

305

[ZZ1,RR1,TT1,T peak1,Nt1,qdot1,hg1,dRR1,dZZ1] = nozzleHeatTransfer(TT init,NP,NQ,dt1,final time S1,D nozzle,nozzle dir cosine,z nozzle le

306

stability1 = dt1/(max(dRR1,dZZ1))2

307

TT init = TT1(:,:,end);

308
309

final time S2 = 5;

310

dt2 = 0.0005;

311

[ZZ2,RR2,TT2,T peak2,Nt2,qdot2,hg2,dRR2,dZZ2] = nozzleHeatTransfer(TT init,NP,NQ,dt2,final time S2,D nozzle,nozzle dir cosine,z nozzle le

312

stability2 = dt2/(max(dRR2,dZZ2))2

313

ZZ = ZZ1(2:end-1,2:end-1);

314

RR = RR1(2:end-1,2:end-1);

315

TT = cat(3,TT1(2:end-1,2:end-1,:),TT2(2:end-1,2:end-1,:));

316

T peak Cer = [T peak1,T peak2]; T peak = T peak Cer;

317

time1 = 1:Nt1; time1 = time1*dt1;

318

time2 = 1:Nt2; time2 = time2*dt2;

319

time Cer = [time1,final time S1+time2]; time = time Cer;

320

final time = final time S1+final time S2;

321

Temp Loss = [qdot1 qdot2]/(Cp bath S2 *mdot bath);

322

T loss = Temp Loss;

323

Peak Temp Loss = max(Temp Loss)

324
325

%% PLOT TEMP RESULTS

326

% load loss stainless0-25.dat time Cer Temp Loss -ascii;

327

% time025L=time Cer; Temp Loss025=Temp Loss;

328
329
330
331

figure(77)

332

plot(time,T loss,'o'); xlim([0,final time])

333

hold on

334

% plot(time025L,Temp Loss025,'*'); xlim([0,final time])

335

xlabel('Time (s)'); ylabel('Temperature Loss (K)');

336

title('Wall Temperature Loss vs. Time');

213

337
338

figure(88)

339

pcolor(ZZ.*m2in,RR*m2in,TT(:,:,end))

340

colorbar('EastOutside');

341

hold on

342

pcolor(ZZ.*m2in,-RR*m2in,TT(:,:,end))

343

axis equal;

344

colormap hot; shading interp;

345

xlim([ZZ(1,1)*m2in,ZZ(1,end)*m2in]); ylim([-D spool inlet/2*m2in*1.3,D spool inlet/2*m2in*1.3]);

346

ylim([-2,2])

347

xlabel('z (in)'); ylabel('R (in)');

348

title(['Temperature Distribution in Nozzle Walls t=' num2str(final time) 's, Max Temp = ' num2str(max(T peak Cer)) 'K'])

349

hold off;

350

view(0,90)

351
352

figure(99)

353

hold off

354

save b.dat time Cer Temp Loss -ascii;

355

save b.dat time Cer T peak Cer -ascii;

356
357

load titanium.dat time T peak -ascii; time Ti=titanium(1,:); T peak Ti=titanium(2,:);

358

% load loss stainless0-25.dat time025L Temp Loss025 -ascii;

359

% load peak temp stainless0-25.dat time025P T peak 025 -ascii;

360

361

% load loss stainless0-5.dat time050L Temp Loss050 -ascii;

362

% load peak temp stainless0-5.dat time050P T peak 050 -ascii;

363

364

% load loss stainless1-0.dat time100L Temp Loss100 -ascii;

365

% load peak temp stainless1-0.dat time100P T peak 100 -ascii;

366
367
368

% load loss stainless0-5.dat time Cer Temp Loss -ascii;

369

% load peak temp stainless0-5.dat time Cer T peak Cer -ascii;

370

% load loss stainless1-0.dat time Cer Temp Loss -ascii;

371

% load peak temp stainless1-0.dat time Cer T peak Cer -ascii;

372
373

plot(time,T peak,'-b'); xlim([0,final time])

374

hold on;

375

load a.dat time Cer T peak Cer -ascii;

376

time025P=time Cer; T peak 025=T peak Cer;

377
378
379

plot(time025P,T peak 025,'mo'); xlim([0,final time])

380

legend(nozzle material,'Titanium')

381

xlabel('Time (s)'); ylabel('Temperature (K)');

382

title('Peak Wall Temperature vs. Time');

383

hold off

384

end

385
386

fprintf(['\n\nMax Mach Number = ' num2str(max(M S2))])

387

fprintf(['\nT exit = ' num2str(T S2(end))])

388
389

fprintf('\nDone');

214

D.2

calcCombustorPenetration.m

%% COMBUSTOR JET PENETRATION HEIGHT

% O2/H2 mixtures are injected as a jet into the 1400K coreflow steam that

% comes from the PHB. The penetration height achieved by the jet is

% calculated based on the usual powerlaw relations for transverse jets with

% circular orifices. Specifically, the powerlaw used here is found in

% "Circular and noncircular subsonic jets in cross flow" - Gutmark et al.

% Details of the following calculations are in Nov 6th Notes.

% NOTE: This calculation is independent of main.m

9
10

function [x,y1] = calcCombustorPenetration()

11

global in2m m2in psi2Pa Pa2psi Patm PLOT ON

12

format short g

13

Ru = 8314; %J/mol.K

14
15

%% Massflow ratio between Jet and Freestream

16

mdotH2 = 0.0116;

17

mdotO2 = 0.0741;

18

mdotJ = 0.0857; %kg/s

19

mdotInf = 0.25; %kg/s

20

massflowRatio = mdotJ/mdotInf;

21
22

%% Area Ratio between Jet and Freestream

23

DJ = 0.4*in2m; %from the injector drawing

24

DInf = 3.5*in2m;

25

AJ = pi*DJ2/4;

26

AInf = pi*DInf2/4;

27

areaRatio = AJ/AInf;

28
29

%% Density Ratio between Jet and Freestream

30

%Density of Freestream

31

PInf = 2.5e5;

32

RInf = Ru/18;

33

TInf = 1400;

34

rhoInf = PInf/(RInf*TInf);

%Nominally 2.5 Bar from PBH


%Gas constant for steam
%Nominally 1400K from PBH

35
36

%Density of Jet

37

MO2 = 32;

%kg/kmol

38

MH2 = 2;

%kg/kmol

39

RO2 = Ru/MO2;

40

RH2 = Ru/MH2;

41

MolH2 = mdotH2/MH2;

%kmols of H2 = kg of H2 / kg/kmol

42

MolO2 = mdotO2/MO2;

%kmols of O2 = kg of O2 / kg/kmol

43

MolFracH2 = MolH2/(MolH2+MolO2);

44

MolFracO2 = MolO2/(MolH2+MolO2);

45

CpH2 = 28.21; %J/mol.K from NIST webbook at 450K

46

CpO2 = 30.50; %J/mol.K from NIST webbook at 450K

47

CpmixJ = MolFracH2*CpH2+MolFracO2*CpO2;

48

gJ = CpmixJ/(CpmixJ-Ru/1000);

49

gJm = gJ-1;

50

gJp = gJ+1;

51

RJ = (RO2*MO2+RH2*MH2)/(MO2+MH2);

52

%Gas constant for O2 and H2 mixture

215

53

MJ = 0.9;

54

T0J = 450;

55

P0J = 400*psi2Pa;

56

PJ = P0J/(1+gJm/2*MJ2)(gJ/gJm);

57

TJ = T0J/(1+gJm/2*MJ2);

58

rhoJ = PJ/(RJ*TJ);

59

densityRatio = rhoJ/rhoInf;

%Nominally 450K from O2 and H2 heaters


%Nominally 400psi from O2/H2 regulators

60
61

%% Calculate Momentum Flux Ratio r and penetration profile y

62

r = massflowRatio*sqrt(1/densityRatio)/areaRatio;

63

d = 0.25*in2m;

64

Aconst = 2.05;

65

mconst = 0.28;

66

x = (0:0.05:20)*in2m;

67

y1 = ((r*d)*Aconst*(x./(r*d)).mconst)*m2in;

68

if PLOT ON == 1

69

figure()

70

ymid = x./x*3.5/2;

71

yspool = x./x*3.5;

72

hold on

73

plot(x*m2in,y1,'r-')

74

plot(x*m2in,ymid,'--k')

75

plot(x*m2in,yspool,'b')

76

xlim([-1,x(end)*m2in])

77

xlabel('x (in)'); ylabel('Penetration Height (in)');

78
79

title('Combustion Gas Penetration Height vs Downstream Distance')


end

D.3

calcNozzleFlowProperties.m

function [M,a,P gas,T gas,rho,u,q,Re,mu,mu CH4] = calcNozzleFlowProperties(z,T0 bath,P0 bath,A nozzle,D nozzle,A star,g,z throat,R,visc,

%% CALCULATE INCOMING GAS PROPERTIES

% Temperature, Pressure, Density, Velocity etc, of a gas based on

% isentropic choked flow.

5
6

%Inlet Conditions

gm = g-1;

gp = g+1;

9
10

options=optimset('display','off');

%sets options for fsolve

if PLOT PROFILE == 1

11

Minit = getMach(0.1, options, g, A nozzle(1)/A star); %Incoming Mach Number

12

T in = T0 bath/(1+gm/2*Minit2);

13

P in = P0 bath/(1+gm/2*Minit2)(g/gm); %pressure at the nozzle inlet

14

rho in = P in/(R* T in);

%not used

15

ainit = sqrt(g*R* T in);

%not used

16

%temperature at the nozzle inlet

end

17
18

%Calculate properties at position z in the nozzle

19

for i = 1:length(z);

20

if z(i)z throat

21
22

M0 = 0.7;
elseif z(i)>z throat;

216

23

M0 = 1.8;

24

end

25

M(i) = getMach(M0, options, g, A nozzle(i)/A star); %finds Mach # for fanno and isentropic flow

26

T0 T = 1+(g-1)/2*M(i)2; %Ratio of stagnation to static air pressure

27

T gas(i) = T0 bath/T0 T;

28

P gas(i) = P0 bath/(T0 T)(g/gm);

29

rho(i) = P gas(i)/(R* T gas(i));

30

a(i) = sqrt(g*R* T gas(i));

31

u(i) = M(i)*a(i);

32

q(i) = rho(i)*u(i).2;

33

visc(i) = 4.90E-04* T gas(i)0.656;

34

Re = (rho(i)*u(i)*D nozzle(i))/visc(i);

%calculate local static temp

%calculate local static density


%calculate local speed of sound

%calculate local flow velocity

35
36

end

37

mu = visc;

38

mu CH4 = getCH4Viscosity(T gas,P gas);

39
40
41

%% PLOTS

42

%NOZZLE PROFILES

43

if PLOT PROFILE

44

global m2in;

45

figure()

46

grid on

47

zin = z*m2in;

48

plot(zin,M,'k'); xlabel('z (in)'); ylabel('M'); %title('z vs Mach Number')

49

figure()

50

grid on

51

plot(zin,T gas,'k'); xlabel('z (in)'); ylabel('T (K)'); %title('z vs Temperature')

52

figure()

53

grid on

54

plot(zin,P gas),'k'; xlabel('z (in)'); ylabel('P (Pa)'); %title('z vs Pressure')

55

figure()

56

grid on

57

plot(zin,rho),'k'; xlabel('z (in)'); ylabel('\rho (kg/m3)'); %title('z vs Density')

58

figure()

59

grid on

60

plot(zin,u),'k'; xlabel('z (in)'); ylabel('u (m/s)'); %title('z vs Velocity')

61

figure()

62

grid on

63

plot(zin,q,'k'); xlabel('z (in)'); ylabel('\rho u2 (kg/m.s-2'); %title('z vs Momentum Flux')

64

figure()

65

grid on

66
67

plotyy(zin,q,zin,T gas); title('z vs Momentum Flux & Temperature')


end

D.4

getArea.m

function [A,D,z,dir cosine,throat index z,z throat] = getArea(dist,z throat,D star,z nozzle length,D spool inlet,D spool outlet)

global in2m m2in

global NOZZLE TYPE

217

switch NOZZLE TYPE

%CONICAL NOZZLE

case 'conical'

Asub=[0 1;z throat 1]; bsub=[D spool inlet;D star];

Asup=[z throat 1;z nozzle length 1]; bsup=[D star;D spool outlet];

10
11

mc sub=Asub\bsub; msub=mc sub(1); csub=mc sub(2);

12

mc sup=Asup\bsup; msup=mc sup(1); csup=mc sup(2);

13
14

for i = 1:length(dist)

15

z = dist(i);

16

D(i)=0;

17

if z 0 && z z throat

18

D(i) = msub*z+csub;
elseif z > z throat && z z nozzle length

19
20

D(i) = msup*z+csup;

21

else

22

A(i) = 1; D(i) = 1; disp(['Error: z=' num2str(z)])

23

end

24

A(i) = pi*D(i).2/4;

25

end

26

z = dist;

27

dz = z(2)-z(1);

28

slope = diff(D./2);

29

gradnt = slope/dz;

30

dir cosine = cos(atan([gradnt(1) gradnt gradnt(end) 0]));

31

[minz,throat index z] = min(D);

32
33

%SOFIA'S EXPERIMENTALLY DETERMINED PROFILE FROM CFD

34

case 'cfd'

35

z = dist;

36

R spool inlet = D spool inlet/2;

37

R spool outlet = D spool outlet/2;

38

R star = D star/2;

39
40

%CALCULATE SUBSONIC SECTION PARAMETERS

41

Rad1 = 1.5* R star;

42

Rc1 = 2.5* R star;

43

d = 1.5* R star-sqrt(Rad12-(R spool inlet-Rc1)2);

44

zc1 = 1.5* R star;

%subsonic radius of 1.5R* circle center coord R

%subsonic radius of 1.5R* circle center coord z

45
46

%INTERPOLATE NOZZLE PROFILE DATA POINTS FOR SUPERSONIC SECTION

47

z profile = [1.289,1.682,2.075,2.467,2.860,3.252,3.645,4.037,4.430,4.780,5.215,5.449,5.724,6.0]*in2m;

48

R profile = [1.066,1.087,1.122,1.169,1.217,1.260,1.307,1.350,1.390,1.421,1.457,1.472,1.488,1.5]*in2m;

49
50
51

%DETERMINE NOZZLE SHAPE R VS. Z

52

for i = 1:length(z)

53

%SUBSONIC

54

if z(i)0 && z(i)zc1

55

R(i) = Rc1-sqrt(Rad12-(z(i)-zc1)2);

56

%SUPERSONIC

57

elseif z(i)>zc1 && z(i)<z profile(end)

58
59

R(i) = interp1(z profile,R profile,z(i),'spline');


else

60
61

R(i)= R spool outlet; %disp(['Error: z=' num2str(z(i))]);


end

218

62

end

63
64

D = R*2;

65

A = pi*R.2;

66

figure()

67

hold off;

68

grid on

69

plot(z*m2in,R*m2in,'k','MarkerSize',2)

70

ylim([0,1.8])

71

xlabel('z (in)'); ylabel('Nozzle Radius (in)')

72
73

dz = z(2)-z(1);

74

slope = diff(D./2);

75

gradnt = slope/dz;

76

dir cosine = cos(atan([gradnt(1) gradnt gradnt(end) 0]));

77

[minz,throat index z] = min(D);

78

z throat = zc1;

79
80

%BELL NOZZLE

81

case 'bell'

82

z = dist;

83

R spool inlet = D spool inlet/2;

84

R spool outlet = D spool outlet/2;

85

R star = D star/2;

86
87

%CALCULATE PARABOLIC SECTION PARAMETERS

88

theta n = 19;

89

theta e = 12;

90

Re = R spool outlet;

91

R0382 = 0.382;

%R-coord of end of parabola (Sept 30 notes)

92
93

%CALCULATE SHIFT (d) in subsonic circular arc (Oct 4 notes) so that

94

%inlet radius is the same as the spool radius.

95

Rad1 = 1.5* R star;

96

Rad2 = R0382* R star;

97

Rc1 = 2.5* R star;

98

d = 1.5* R star-sqrt(Rad12-(R spool inlet-Rc1)2);

%subsonic radius of 1.5R* circle center coord R

99
100

%CALCULATE CENTERS OF CIRCULAR ARCS FOR SUBSONIC & THROAT SECTIONS

101

zc1 = 1.5* R star-d;

102

Rc2 = (1+R0382)* R star; %throat radius of 0.382R* circle center coord R

103

zc2 = 1.5* R star-d;

%subsonic radius of 1.5R* circle center coord z

%throat radius of 0.382R* circle center coord z

104
105

%CALCULATE PARABOLA CONSTANTS FOR BELL NOZZLE APPROXIMATION

106

zn = zc1+R0382* R star *sind(theta n);

107

Rn = R star *(1+R0382*(1-cosd(theta n)));

108

Lf = 1;

109

Lconical = (Re-Rn)/tand(15);

110

Ln = Lconical*Lf;

111

Lp = (Ln-R0382* R star *sind(theta n));

112

ze = zn+Lp;

%z-coord of start of parabola (Oct 2 notes)


%R-coord of start of parabola (Oct 2 notes)

%z-coord of end of parabola (Sept 30 notes)

113
114

ke = 1/(tand(theta e))2;

115

kn = 1/(tand(theta n))2;

116
117
118

z0 = [-20,3,0];
%

% Make a starting guess at the solution

optns=optimset('Display','notify','Algorithm','levenberg-marquardt','TolFun',1e-8,'MaxFunEvals',5e4);

219

119

optns=optimset('Display','off','TolFun',1e-8,'MaxFunEvals',5e5);

120

[u,fval] = fsolve(@NozzleParameterFSolve,z0,optns);

121

a=u(1); b=u(2); c=u(3);

122

theta n = atand(1/sqrt(b2-4*a*(c-zn)));

123

theta e = atand(1/sqrt(b2-4*a*(c-ze)));

124

theta ne = [theta n,theta e]

% Call optimizer

125
126
127

%DETERMINE NOZZLE SHAPE R VS. Z

128

for i = 1:length(z)

129

%SUBSONIC RADIUS OF 1.5R*

130

if z(i)0 && z(i)zc1

131

R(i) = Rc1-sqrt(Rad12-(z(i)-zc1)2);

132

%THROAT RADIUS OF 0.382R*

133

elseif z(i)>zc1 && z(i)zn

134

R(i) = Rc2-sqrt(Rad22-(z(i)-zc2)2);

135

%PARABOLIC BELL SECTION

136

elseif z(i)>zn && z(i)<ze

137

R(i) = (-b+sqrt(b2-4*a*(c-(z(i)))))/(2*a);

138

else

139

R(i)= R spool outlet; %disp(['Error: z=' num2str(z(i))]);

140

end

141

end

142
143

D = R*2;

144

A = pi*R.2;

145

figure()

146

hold off;

147

ZZ = zn:0.001:ze;

148

ZZin = ZZ*m2in;

149

RR = (-b-sqrt(b2-4*a*(c-ZZ)))/(2*a);

150

plot(ZZ*m2in,RR*m2in)

151
152

plot(z*m2in,R*m2in,'k','MarkerSize',2)

153

hold off; ylim([0,1.8])

154

grid on

155

xlabel('z (in)'); ylabel('Nozzle Radius (in)')

156
157

dz = z(2)-z(1);

158

slope = diff(D./2);

159

gradnt = slope/dz;

160

dir cosine = cos(atan([gradnt(1) gradnt gradnt(end) 0]));

161

[minz,throat index z] = min(D);

162
163

z throat = zc1;
end

164
165

function F = NozzleParameterFSolve(u)

166

F = [u(2)2 - 4*u(1)* (u(3) - zn)- kn

167

1/(2*u(1))* (-u(2) + sqrt(u(2)2-4*u(1)*(u(3)-ze)))-Re;

168

1/(2*u(1))* (-u(2) + sqrt(u(2)2-4*u(1)*(u(3)-zn)))-Rn];

169
170

end
end

D.5

getCH4Viscosity.m

220

function mu gas = getCH4Viscosity(T,P)

2
3

%From engineering toolbox

Pd = [1,2,5,10,20,50].*6894.75729;

Td = (1500:-50:300-32)./1.8;

[Pdata,Tdata] = meshgrid(Pd,Td);

mudata = [...

0.041

0.041

0.041

0.041

0.041

0.041;

0.04

0.04

0.04

0.04

0.04

0.04;

10

0.039

0.039

0.039

0.039

0.039

0.039;

11

0.038

0.038

0.038

0.038

0.038

0.038;

12

0.037

0.037

0.037

0.037

0.037

0.037;

13

0.035

0.035

0.035

0.035

0.035

0.035;

14

0.034

0.034

0.034

0.034

0.034

0.034;

15

0.034

0.034

0.034

0.034

0.034

0.034;

16

0.032

0.032

0.032

0.032

0.032

0.032;

17

0.031

0.031

0.031

0.031

0.031

0.031;

18

0.03

0.03

0.03

0.03

0.03

0.03;

19

0.029

0.029

0.029

0.029

0.029

0.029;

20

0.028

0.028

0.028

0.028

0.028

0.028;

21

0.026

0.026

0.026

0.026

0.026

0.026;

22

0.025

0.025

0.025

0.025

0.025

0.025;

23

0.024

0.024

0.024

0.024

0.024

0.024;

24

0.023

0.023

0.023

0.023

0.023

0.023;

25

0.022

0.022

0.022

0.022

0.022

0.022;

26

0.021

0.021

0.021

0.021

0.021

0.021;

27

0.02

0.02

0.02

0.02

0.02

0.02;

28

0.019

0.019

0.019

0.019

0.019

0.019;

29

0.018

0.018

0.018

0.018

0.017

0.017;

30

0.016

0.016

0.016

0.016

0.016

0.016;

31

0.015

0.015

0.015

0.015

0.015

0.015;

32

0.014

0.014

0.014

0.014

0.014

0.014;

33

];

34

for i = 1:length(T)

35
36

mu gas = interp2(Pdata,Tdata,mudata,P(i),T(i),'spline');
end

D.6

getGasProperties.m

function [g,gm,gp,visc,R,P0,T0,Cp] = getGasProperties(gas)

global G R bath G P0 bath G T0 bath G Cp bath G Cp H2O Patm G Cp mix

switch gas

case 'precombusted S1'

% STAGE 1 (S1): 1400K steam only from Pebble Bed Heater

visc = 9.68e-5;

R = 461.5;

P0 = 2.5*Patm;

T0 = 1400;

10

Cp = G Cp H2O;

11

g = Cp/(Cp-R);

12

gp = g+1;

13

gm = g-1;

221

14

case 'bath S2'

15

% STAGE 2 (S2): 2100K steam + H2 from combustor and Pebble Bed Heater.

16

visc = 9.68e-5;

17

R = 8314/18.02;

18

P0 = G P0 bath;

19

T0 = G T0 bath;

20

Cp = G Cp bath;

21

g = Cp/(Cp-R);

22

gp = g+1;

23

gm = g-1;

24

case 'CH4-H2O Mix'

25

% Mixture of Methane and Steam.

26

visc = 9.68e-5;

27

R = 8314/16.042;

28

mdot total = 0.4351;

29

mdot bath S2 = 0.3356;

30

mdot CH4 = 0.099451;

31

Cp CH4 = 4177.7; %J/kg.K at 900K

32

T01 = 2100; %Steam temp after nozzle

33

T0j = 900;

34

P0 = G P0 bath *0.9;

35

Cp = (mdot bath S2 * G Cp bath+mdot CH4* Cp CH4)/mdot total;

36

T0 = (G Cp bath *T01* mdot bath S2+Cp CH4 *T0j*mdot CH4)/(Cp*mdot total); %Energy Eq from Nov 23 Notes;

%Temp of injected methane

37
38

g = Cp/(Cp-R);

39

gp = g+1;

40

gm = g-1;

41

otherwise

42

g = 1.33;

43

gp = g+1;

44

gm = g-1;

45

visc = 9.68e-5;

46

R = G R bath;

47

P0 = G P0 bath;

48

T0 = G T0 bath;

49
50

Cp = G Cp bath;
end

D.7

function M air = getMach(M0 air, options, gam, A At)

invMachArea = @(M,options)(((1+(gam-1)/2)-((gam+1)/(2*(gam-1)))/(M*(1+((gam-1)/2)*M2)-((gam+1)/(2*(gam-1)))))-A At);

3
4

getMach.m

M air =fsolve(invMachArea,M0 air,options);


end

D.8

getNozzleMaterialConstants.m

%solve for mach number given area ratio

222

function [k,rho,Cp,alpha] = getNozzleMaterialConstants(material)

2
3

switch material

case 'Titanium'

%Titanium data

k = 21.9;

%Thermal Conductivity

rho = 4540;

%Density

Cp = 540;

%Specific Heat

alpha = k/(rho*Cp);

10

W/m.K

gm/cm3
J/kg.K

%Thermal Diffusivity m2/s

case 'Copper'

11

%Copper data from google

12

k = 400;

%Thermal Conductivity

13

rho = 8960;

%Density

14

Cp = 385;

%Specific Heat

15

alpha = k/(rho*Cp);

16

W/m.K

gm/cm3
J/kg.K

%Thermal Diffusivity m2/s

case 'Stainless316'

17

%Stainless Steel 316 data from

18

%http://www.azom.com/details.asp?Articleid=863

19

k = 21.5;

%Thermal Conductivity 16W/m.K @ 100C, 21.5W/m.K @ 500C

20

rho = 8000;

%Density

21

Cp = 500;

%Specific Heat

22

alpha = k/(rho*Cp);

23

gm/cm3
J/kg.K

%Thermal Diffusivity m2/s

case 'Zirconium'

24

%Zirconium from CiDRA

25

k = 2.2;

26

rho = 5720;

%Density 5.5-6.1 gm/cm3

27

Cp = 400;

%Specific 400-600 Heat J/kg.K

28

alpha = k/(rho*Cp);

29

%Thermal Conductivity 1.5-2.9 W/m.K

%Thermal Diffusivity m2/s

case 'Alumina'

30

%Aluminum Oxide (Al2O3 90-95%) Data from CiDRA Ceramics

31

k = 25.6;

%Thermal Conductivity 14-25.6 W/m.K

32

rho = 3800;

%Density 3.4-3.8 gm/cm3

33

Cp = 850;

%Specific Heat 850-1050 J/kg.K

34

alpha = k/(rho*Cp);

35

%Thermal Diffusivity m2/s

case 'Mullite'

36

%Mullite (Al2O3-SiO2)

37

k = 5.0;

%Thermal Conductivity 1.5-5 W/m.K http://www.earthwaterfire.com/pdf library/material mullite.pdf

38

rho = 2770;

%Density 2.3-2.7 gm/cm3

39

Cp = 850;

%Specific Heat 850-1050 J/kg.K

40

alpha = k/(rho*Cp);

41

%Thermal Diffusivity m2/s

end

D.9

nozzleHeatTransfer.m

%% FINITE DIFFERENCE HEAT TRANSFER ANALYSIS

% MODELS TRANSIENT 2D CONDUCTION IN THE NOZZLE WALL VIA A FINITE DIFFERENCE

% DISCRETIZATION OF THE HEAT EQUATION IN POLAR COORDINATES.

% BOUNDARY CONDITIONS:

% INNER WALL TEMPERATURE IS DEPENDENT ON CONVECTIVE HEAT TRANSFER

% RATE THAT IS CALCULATED USING BARTZ'S EQUATION.

% OUTER WALL TEMPERTURE IS DEPENDENT ON RADIATIVE HEAT TRANSFER RATE.

% NOZZLE ENDWALLS ARE FIXED AT T=400K

223

9
10

function [ZZ,RR,TT,T peak,Nt,qdot,hg mtx,max r ,max z] = nozzleHeatTransfer(TT initial,NP,NQ,dt,final time,D nozzle,nozzle dir cosine

11

global HEAT TRANSFER ON PLOT ON

12

global m2in in2m

13
14

if HEAT TRANSFER ON

15
16

%GRID

17

Ti thickness = 1*in2m; %nozzle wall thickness

18

R nozzle = D nozzle./2;

19

r mean = [R nozzle(1) R nozzle R nozzle(end)];

20

z0 = 0;

21

zend = z nozzle length;

22

dZZ = dz

23
24

%NP = number of radial points in the grid

25

%NQ = number of axial points in the grid

26

%i indexes radially, j indezes axially

27

for j = 1:NQ+2

28

r1(j) = r mean(j);

29

r2(j) = r mean(j)+Ti thickness./nozzle dir cosine(j);

30

dRR(j) = (r2(j)-r1(j))/(NP-1);

31

for i = 1:NP+2

32

ZZ(i,j) = z0+(j-2)*dZZ;

33

RR(i,j) = r1(j)+(i-2)*dRR(j);

34

end

35

end

36

disp('...')

37

max z = dZZ

38

max r = max(dRR)

39
40

if PLOT ON == 2

41

figure()

42
43

plot(RR(end,:)); ylim([0,max(RR(end,:))])
end

44
45

%INITIAL CONDITIONS

46

Nt = floor(final time/dt)+1;

47

TT default = 400;

48

final time = dt*Nt;

49

TT = ones(NP+2,NQ+2,Nt).*TT default;

50

if TT initial == 0

51

else

52

%intitial nozzle temperature at time t=0

TT(:,:,1) = TT initial;

53

end

54

TT(:,:,1);

55
56

%CONDUCTION CONSTANTS

57

[k wall,rho wall,Cp wall,alpha wall] = getNozzleMaterialConstants(nozzle material);

58
59

%CONVECTION CONSTANTS

60

D star = sqrt(4* A star/pi);

61

C star = P0 bath * A star/mdot bath;

62

k steam = 0.000128* T gas + 0.00365; %engineering toolbox @ 125C

63

mu steam = 4e-08* T gas + 9e-06; %trendline in Steamproperties.xlsx

64
65

omega gas = 0.69;

%curve fitting a power trendline in Excel

224

66

Cp steam = Cp bath;

67

rc = D star/2*1.5; %Modern Engineering for design of liquid-propellant rocket engines - Dieter, Huzel, Huang pg 76

68

hg const = (0.026/D star0.2).*(mu steam.0.2*Cp steam/(mu steam.*Cp steam/k steam)0.6).*...

69
70

(P0 bath/C star)0.8.*(D star/rc).0.1;


%RADIATION CONSTANTS

71

sigma SBconst = 5.6704e-8;

72

nozzle emissivity = 0.6;

73

nozzle absorptivity = 0.8;

74

Tinf env = 298;

75

rt(1)=0; rt(2)=0;

76

T peak = zeros(1,Nt);

77

qconv Area total = 0;

78

qconv total(1) = 0;

79

qconv rad(1) = 0;

80

nn = 0; %index for storing film coefficient values

81
82

TTprime = TT;

83

for n = 1:Nt

84

tic

85

time(n) = n*dt;

86

qrad total = 0;

87

qconv Area total = 0;

88

%UPDATE BOUNDARY HEAT FLUXES AND END WALL FIXED BOUNDARIES

89

for j = 1:NQ

90

%CONVECTION VIA BARTZ EQUATION

91

Tw1(j) = TT(2,j,n);

92

Tinfg(j) = T gas(j);

93
94

sig(j) = 1/((0.5*Tw1(j)/T0 bath*(1+gm/2*M(j)2)+0.5)...

95

(0.8-0.2*omega gas)*(1+gm/2*M(j)2)(0.2*omega gas));

96

hg(j) = hg const(j)*(A star/A nozzle(j))0.9*sig(j);

97
98

if j == throat index z && rem(n,20) == 0

99

[time(n) max(Tw1) T0 bath./(1+gm/2*M(j)2) hg(j)]

100

if isnan(Tw1(j))

101

disp('Error: Wall temp is NaN, reduce timestep')

102

break

103

end

104

end

105
106

qconv flux(j) = real(hg(j)*(Tinfg(j)-Tw1(j))); % heat flux W/m2

107

if (-Tw1(j)+T0 bath)<50 && n>Nt/10

108

qconv flux(j) = 0;

109

gh = 1

110

end

111

TT(1,j+1,n) = qconv flux(j)*dRR(j)/k wall + TT(2,j+1,n); %update boundary ghost points

112
113
114

%HEAT LOSS FROM STEAM INTO NOZZLE VIA CONVECTION

115

R1 = R nozzle(j);

116

z1 = ZZ(1,j);

117

if j < NQ

118

R2 = R nozzle(j+1);

119
120

z2 = ZZ(1,j+1);
else

121
122

R2 = 0; R1 = 0; z2 = 1;
end

225

123

m conv const = (R1-R2)/(z1-z2);

124

c conv const = R1-m conv const *z1;

125

qconv Area = 2*pi*(0.5* m conv const *(z22-z12)+c conv const *(z2-z1)); %Oct 23 notes

126

qconv Area total = qconv Area total+qconv Area;

127

qconv total(j+1) = qconv total(j) + qconv flux(j)*qconv Area; % Sum up all heat transfer J/s

128
129

%RADIATION LOSS TO SURROUNDINGS AT TEMP Tinf env

130

Tw2 = TT(end-1,j,n); %temperature of the outer nozzle wall

131

qrad flux(j) = sigma SBconst*(nozzle emissivity*Tw24-nozzle absorptivity*Tinf env);

132

TT(end,j+1,n) = qrad flux(j)*dRR(j)/k wall + TT(end-1,j+1,n);

133
134

%HEAT LOSS FROM NOZZLE INTO SURROUNDINGS VIA RADIATION

135

qrad Area = qconv Area;

136

qrad total(j+1) = qrad total(j) - qrad flux(j)*dRR(j)*2*pi*(R nozzle(j)+Ti thickness./nozzle dir cosine(j));

137
138

end

139

if rem(n,floor(1/dt)) == 0

140

nn = nn+1;

141

hg mtx(nn,:) = hg; %store film coefficient

142

end

143

TT(:,1,n) = TT(:,2,n);

%inlet (adiabatic Neumann BC)

144

TT(:,end,n) = TT(:,end-1,n);

%outlet (adiabatic Neumann BC)

145
146

for j = 2:NQ+1

147

for i = 2:NP+1

148

%Update each temperature node

149

Tij = TT(i,j,n);

150

Timj = TT(i-1,j,n);

151

Tipj = TT(i+1,j,n);

152

Tijm = TT(i,j-1,n);

153

Tijp = TT(i,j+1,n);

154
155

Tr = (Tipj-Timj)/2/dRR(j);

156

Trr = (Timj-2*Tij+Tipj)/(dRR(j))2;

157

rz = (ZZ(i,j+1)-ZZ(i,j-1))/2/dRR(j);

158

rzz = (ZZ(i,j-1)-2*ZZ(i,j)+ZZ(i,j+1))/(dRR(j))2;

159

Tzz = Trr*rz2+Tr*rzz;

160

TTprime(i,j,n) = alpha wall*(1/RR(i,j)*Tr+Trr+Tzz);

161

if n==1

162

sn = 1; sA = -1;

163

else

164

sn = n-1; sA = 1;

165

end

166

TT(i,j,n+1) = TT(i,j,n)+ dt/2*(3*TTprime(i,j,n)-sA*TTprime(i,j,sn));

167
168

end
end

169
170
171

rt(1) = rt(1)+toc;

172

tic

173

if (rem(n,50) == 0 | | n == 1) && PLOT ON == 1

174

format short g

175

fprintf(['Conv = ' num2str(sum(qconv total)) 'J/s

176

figure(99)

177

subplot(1,2,1)

178

pcolor(ZZ.*m2in,RR*m2in,TT(:,:,n))

179

if n == Nt

Rad = ' num2str(sum(qrad total)) 'J/s'])

226

180

colorbar('EastOutside');

181

end

182

hold on

183

pcolor(ZZ.*m2in,-RR*m2in,TT(:,:,n))

184

axis equal; colormap hot; shading flat;

185

xlim([ZZ(1,1)*m2in,ZZ(1,end)*m2in]); ylim([-D spool inlet/2*m2in*1.3,D spool inlet/2*m2in*1.3]);

186

zlabel('z (in)'); ylabel('y (in)');

187

title(['Temperature Distribution in Nozzle Walls t=' num2str(final time) 's, Max Temp = ' num2str(max(max(TT(2:end-1,2:end-1,n))

188

hold off;

189

subplot(1,2,2)

190
191

for nn = 1:n

192

Temp Throat(nn) = TT(2,throat index z,nn);

193

end

194

plot(time,Temp Throat); xlim([0,final time])

195

xlabel('time (s)'); ylabel('Throat Wall Temperature (K)');

196

hold on;

197
198

title(['Throat Wall Temperature vs. Time']);

199

pause(0.01)

200

hold off

201

qconv Area total;

202

end

203

rt(2) = rt(2)+toc;

204

T peak(n) = max(max(TT(2:end-1,2:end-1,n)));

205

qdot(n) = qconv total(j);

206

if rem(n,10) == 0

207

qdot(n);

208
209

end
end

210
211

TT(2:end-1,2:end-1,end);

212

else

213
214

ZZ=1;RR=1;TT=1;T peak=1;Nt=1;
end

D.10

plotCrossSectionalMixing.m

function err = plotCrossSectionalMixing(PRIMARY INJECTOR,num injectors,hp,h,D throat primary jet,D throat jet)

2
3

% for i = 1:num injectors

% end

[injX(i,:),injY(i,:)] = genCircle(c,r)

6
7

err = '';

227

Appendix E
NOZZLE AND INJECTOR FABRICATION

In Figure E.1 a CAD model of the fully assembled feedstock injection section is shown. A
cutaway exploded view is shown in Figure E.2 showing various parts as they were before
being welded together.

Figure E.1: CAD Model Of Injection Section.

Injector port detail can be seen in Figure E.3. Injection manifolds attach to the Swagelok
weld adapaters shown in the image. Gasses enter the adapters perpendicularly and are
turned obliquely as they enter the injection port flow channels that are inclined. The
aeroramp is at the bottom of the image whilst a single secondary injector can be seen at
the top of the image.

228

Figure E.2: Exploded Cutaway Of Injection Section CAD Model

Figure E.3: Cutaway of CAD Model Showing Injetor Ports

229

The nozzle and injection section were fabricated from stainless steel 304 and a CNC lathe
cut their internal profiles. Photographs are shown in Figure E.4 of the outside of the nozzle
(left), inside nozzle profile (mid), and injection section with aeroramp injector ports (right).

Figure E.4: Photographs of fabricated parts: outside of the nozzle (left), inside nozzle profile
(mid), and nozzle with injection section showing aeroramp injector ports (right)

You might also like