You are on page 1of 229

UNIVERSITY OF CALGARY

Numerical and Experimental Studies of Catalytic Combustion

in aPacked Bed Reactor

by

LUBNA BASHEER YOUNIS

A DISSERTATION

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE

DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF MECHANICAL AND MANUFACTURING ENGINEERING

CALGARY, ALBERTA

NOVEMBER, 2003

- © LUBNA BASHEER YOUNIS 2003


UNIVERSITY OF CALGARY

FACULTY OF GRADUATE STUDIES

The undersigned certify that they have read, and recommend to the Faculty of Graduate

Studies for acceptance, a thesis entitled "Numerical and Experimental Studies of

Catalytic Combustion in aPacked Bed Reactor" submitted by Lubna Younis in partial

fulfillment of the requirements of the degree of Doctor of Philosophy.

Supervisor, Dr. I. Wierzba


Department of Mechanical and
Manufacturing Engineering

Dr. G A. Karim
Department of Mechanical and
Manufacturing Engineering

Department of Chemical and


Petroleum Engineering

Dr. A. Jeje
Department of Chemical and
Peleum..Egineering

External Examiner, Dr. M. D. Checkè


Department of Mechanical Engineering
University of Alberta

November 28th 2003


Date

11
ABSTRACT

A one-dimensional, non-equillibrium mathematical model of the catalytic combustion in

apacked bed reactor was developed. The model includes heat conduction, convective and

radiative heat transfer, axial diffusion, and accounts for both homogeneous (gas-phase)

and heterogeneous (surface) reactions of the Arrhenius type. The unsteady state equations

of conservation of mass, chemical species, and energy with appropriate boundary and

initial conditions were solved using afinite-difference procedure. Numerical simulation

was conducted with mainly lean premixed preheated methane-air mixtures over platinum

and abinary mixture of C0304/Cr2O3 catalysts for arange of equivalence ratios, approach

velocities, and inlet temperatures for three different cases: (i) inert bed (gas-phase), (ii)

heterogeneous reaction only, and (iii) both heterogeneous and homogeneous reactions.

The effects of the bed porosity, solid thermal conductivity, and bed length on methane

conversion were also studied. It was found that the solid thermal conductivity has a

strong influence on methane conversion when both homogeneous and heterogeneous

reactions were included. The results also showed that the bed length affected the methane

oxidation significantly especially when the bed porosity was greater than 0.45.

The catalytic combustion of other gaseous-fuels (ethane, propane, carbon monoxide, and

hydrogen) over platinum and binary mixture of C0304/Cr2O3 catalysts was also examined

numerically. The results show that the conversion of these fuels exhibits similar behavior;

i.e. the conversion of the fuel increased with an increase in the inlet temperature and

equivalenáe ratio, and a decrease in the approach velocity. It was also found that the

complete conversion of hydrogen took place at much lower inlet temperatures than those

required for the complete conversion of other fuels. The obtained results are consistent

with the reactivity of theses fuels as well as with the observed experimental trends.

Experimental studies for model validation were also conducted for the catalytic oxidation

of lean methane-air mixtures within the packed bed reactor. The model developed
together with the corresponding experimental data was used to establish the kinetic data

for global surface reaction (the activation energy and the pre-exponential factor), over

platinum and C0304/Cr2O3 catalysts for the fuels employed within the packed bed

reactor.

111
ACKNOWLEDGMENTS

The author wishes to express her sincere gratitude to her supervisor, Dr. I. Wierzba for

her valuable advice, support and guidance during the course of thi research and thesis
writing.

The author also would like to extend her thanks to Dr. G. A. Karim, Dr. D. Checkel, Dr.

S. A. (Raj) Mehta, and Dr. A. Jeje for serving on her exam committee and providing

valuable suggestions in conjunction with this study. Dr. Karim's continuous support and

encouragement throughout the duration of this investigation is highly appreciated.

Thanks should also go to Dr. A. A. Mohamad for his invaluable and endless advises and

guidance regarding the model development part of this research.

Thanks should go to the technical staff of the Department of the Mechanical and

Manufacturing Engineering machine shops, especially to R. Apperal, G. East, B.

Ferguson, A. Moehrle, R. Mosher, and B. Sanders for their assistance in instrumentation


and apparatus preparation and modifications.'

Sincere thanks go to J. McNeely, for his prompt response to any assistance required by

the author for the apparatus and instrumentations repair and preparation with

considerable effort. The author also acknowledges N. Vogt for providing technical staff

promptly when they are needed.

The author would like to thank her office-colleague H. Li for his constructive comments

and discussion during the progress of this work. Thanks should also go to her lab-

colleagues Q. Wang and C. Veer.

The cooperation of the secretarial staff in the Department of the Mechanical and
Manufacturing Engineering is also acknowledged.

Special thanks are attributed to the author's family for their continuous encouragement,

devotion and moral support throughout the duration of this study.

Finally, the financial assistance of the University of Calgary and National Science and

Engineering Research Council of Canada (NSERC) is greatly acknowledged.

iv
DEDICATION

To my beloved parents

who taught me that knowledge is the

key to success

To my daughters, Shand and Danya,

my son, Mustafa, and

my wonderful husband Majeed


TABLE OF CONTENTS

PAGE
ABSTRACT

ACKNOWLEDGMENTS S iv

DEDICATION

TABLE OF CONTENTS Vi

LIST OF TABLES X

LIST OF FIGURES Xi

LIST OF SYMBOLS xxii

CHAPTER 1: INTRODUCTION 1
1.1 Background .. 1

1.2 Objectives of the Study .12

1.3 Thesis Organization 13

CHAPTER 2: LITERATURE REVIEW 14


2.1 Introduction 14

2.2 Experimental Work 14

2.2.1 Fundamental Studies . 15

2.2.2 Applications .17

2.2.3 Non-Noble Metal Catalysts .19

2.2.4 Catalytic Oxidation of Some Common Gaseous Fuels 20

2.3 Theoretical and Computational Studies 27

2.3.1 Bulk Diffusion 28

2.3.2 Intraparticle Diffusion 28

2.3.3 Coupled Heterogeneous and Homogeneous Reactions . .29

2.3.4 Transport Mechanisms 33

vi
2.4 Summary = 35

CHAPTER 3: ANALYSIS AND MATHEMATICAL MODELING 37


3.1 Introduction .37

3.2 Description of the Physical Model 38

3.3 Description of the Mathematical model 39

3.3.1 Governing Equations ... 39

3.3.2 Kinetic Modeling .44

3.3.3 Initial Conditions ..45

3.3.4 Boundary Conditions 45

3.4 Numerical Model 46

3.4.1 Method of Solution . .46

3.4.2 Computer Code Development .46

3.5 Preliminary Results of Numerical Simulation of Methane 49

3.5.1 Inert Bed 52

3.5.2 Heterogeneous Reactions 55

3.5.3 Coupled Homogeneous-Heterogeneous Reactions 58

3.5.4 Effect of Bed Porosity .60

3.5.5 Effect of Bed Length 64

3.5.6 Effect of Thermal Conductivity of the Bed of Solids •75

CHAPTER 4: EXPERIMENTAL APPARATUS AND PROCEDURE 79


4.1 Motivation 79

4.2 Experimental Apparatus .80

4.2.1 Catalytic Reactor .83

4.2.2 Compressed Air and Main Heater Assembly .86

4.2.3 The Choked Nozzle Meter System .. 86

4.2.4 Fuel Injection System ..88

4.2.5 Exhaust Gas Analysis System 88

vi'
4.2.6 Instrumentation and Measurements 92

4.3 Experimental Procedure .93 -

4.4 Pellet Materials for the Packed Bed Reactor 95

4.5 Apparatus Limitations and Operational Parameters .96

CHAPTER 5: EXPERHV[ENTAL RESULTS 98

5.1 Introduction 98

5.2 Inert Packed Bed 99

5.3 Catalytic Packed Bed .. 101

5.3.1 Effect of Catalyst Aging 101

5.3.2 Combustion of Methane over Aged Platinum Catalyst 107

54 Oxidation of Methane over Binary (C0304/Cr2O3) Catalyst .111

5.5 Summary ...116

CHAPTER 6: NUMERICAL SIMULATION OF METHANE

COMBUSTION IN A PACKED BED REACTOR 117


6.1 Introduction . .117
6.2 Determination of Kinetic Data for Surface Reactions 117
6.2.1 Platinum Catalyst 117

6.2.2 C0304/Cr2O3 Catalyst 122


6.3 Results of Model Prediction 125

6.3.1 Methane Oxidation within the Inert Bed 125

6.3.2 Methane Oxidation within the Heterogeneous Bed 127

6.4 Summary 128


I

CHAPTER 7: OXIDATION OF OTHER GASEOUS FUELS

IN A PACKED BED REACTOR 130

7.1 Introduction 130

viii
7.2 Ethane Combustion 134

7.3 Propane Combustion 139

7.4 Carbon Monoxide Combustion .144

7.5 Hydrogen Combustion .148

CHAPTER 8: COMPARISON OF MODEL PREDICTION WITH

THE EXPERIMENTAL DATA 155

8.1 Introduction 155

8.2 Determination of Kinetic Data for Surface Reactions of Common


Gaseous Fuels .155

8.3 Summary ..162

CHAPTER 9: CONCLUSIONS AND RECOMMENDATIONS 166

9.1 Summary and Conclusions 166,


9.2 Recommendations for Further Work .169

REFERENCES 170

Appendix A: Thermophysical Properties 188

A.1 Gas Specific Heats 188

A.2 Diffusion Coefficient 189

A.3 Gas Thermal Conductivity 191

Appendix B: Choked Nozzles Calibration .......................................192

Appendix C: Thermocouple Correction 195

Appendix D: Pressure Drop .............................. 196

Appendix E: Gas Analysis ......... ............................... .198

Appendix F: Program Listing.............................. ... 199

ix
LIST OF TABLES

TABLE PAGE

Table 3.1 Values of operating parameters employed in the numerical

calculations 49

Table 4.1 Comparison between the measured methane using the gas

chromatograph and the gas analyzer for 4= 0.35 at different

inlet temperatures 90

Table 7.1 Kinetic parameters for gas-phase reactions 132

Table 7.2 Kinetic parameters for surface reactions 133

Table 8.1 Kinetic data for platinum and C0304/Cr2O3 catalysts within the

packed bed reactor for all the fuels employed in this study 161

Table A. 1 Gas specific heat correlation constants 188

Table A.2 Gas specific heat correlation (Eq. A.2) constant 189

Table A.3 Lennard-Jones parameters for species employed in the


calculation 191
LIST OF FIGURES

FIGURE PAGE

Figure 1.1 Classification of catalysts (Hagen, 1999) . .3

Figure 1.2 Reactions of synthesis gas (Hagen, 1999) 5

Figure 1.3 Mechanisms of catalyst deactivation (M =metal) (Hagen, 1999) 8

Figure 3.1 Schematic diagram of the physical model 38

Figure 3.2 Control volumes and mesh generation .47

Figure 3.3 Flow chart of the numerical model 48

Figure 3.4 Comparison between predicted methane conversion using the

kinetic data of Westbrook (1981) and experimental data for

methane conversion from Depiak and Wierzba (1999); u = 1.0

m/s, 4= 0.35, k = 1.0 W/m.K, inert bed 52

Figure 3.5 Comparison between calculated and experimental results of

methane conversion; u = 1.0 m/s, 4= 0.35, ks 1.0 W/m.K,

inert bed 53

Figure 3.6 Calculated methane conversion as a function of inlet

temperature at different equivalence ratios; u= 1.0 m/s, k = 1.0

W/m.K, inert bed, (*) Depiak and Wierzba (1999) 54

Figure 3.7 Calculated methane conversion as a function of inlet

temperature at different approach velocities; 4= 0.35, k = 1.0


W/m.K, inert bed 54

Figure 3.8 Calculated methane conversion as a function of inlet

temperature for various equivalence ratios; u= 1.0 m/s, lc = 1.0

W/m.K, Pt catalyst . 56

Figure 3.9 Calculated methane conversion along the bed; = 0.71, u= 1.0

m/s, Tin = 1200 K, Pt catalyst 56

xi
Figure 3.10 Calculated methane conversion as a function of inlet

temperature for different approach velocities; 4) = 0.35, k = 1.0

W/m.K, Pt catalyst 57

Figure 3.11 Calculated methane conversion as a function of inlet

temperature at different equivalence ratios; u= 1.0 m/s, k = 1.0

W/m.K, Pt catalyst 59

Figure 3.12 Calculated methane conversion as a function of inlet

temperature for different approach velocities; 4) = 0.35, k = 1.0

W/m.K, Pt catalyst 59

Figure 3.13 Calculated methane conversion as a function of inlet

temperatur&for various bed porosities; 4) = 0.35, u= 1.0 mIs, k5

= 1.0 W/(m.K),L50 mm 61

Figure 3.14 Calculated methane conversion as afunction of bed porosity for

different inlet temperatures; 4) = 0.35, u = 1.0m/s, k = 1.0

W/m.K,L50mm 61

Figure 3.15 Effect of bed porosity on methane mass fraction along the

packed bed for 4) = 0.35, u= 1.0 m/s, k1.0 W/m.K, and T1

750 K; homogeneous-heterogeneous model 63

Figure 3.16 Temperature distribution for the gas (solid line) and the solid

(broken line) for various bed porosities with 4) = 0.35, u = 1.0

m/s, k3 = 1.0 W/m.K and Tin = 750 K; homogeneous-

heterogeneous model 63

Figure 3.17 Calculated methane conversion as a function of inlet

temperature for different bed lengths; c= 0.50, 4) 0.35, u 1.0

m/s, k3= 1.0 W/ m.K; Pt catalyst 65

Figure 3.18 Calculated methane conversion as afunction of bed length for

different inlet temperatures; u= 1.0 mIs, 4) = 0.35, c= 0.2, k

1.0 W/m.K; Pt catalyst 65

xli
Figure 3.19 Calculated methane conversion as a fqnction of inlet

temperature fOr different porosities; L = 75 mm, 4) = 0.35, u=

1.0 m/s, k = 1.0 W/ m.K; pt catalyst 66

Figure 3.20 Calculated methane conversion as a function of inlet

temperature different bed porosities; L = 100 mm, 4) = 0.35, u=

1.0 m/s, ks= 1.0 W/ m.K; Pt batalyst 66

Figure 3.21 Methane mass fraction along the packed bed at various inlet

temperatures for a= 0.2 (thick line) and a=0.9 (thin line;

0.35, u=1.0 m/s; heterogeneous reaction, Pt catalyst 68

Figure 3.22 Calculated metliane conversion as afunction of bed porosity for

different bed lengths; u = 1.0 m/s, 4) 0.35, Ti,, = 1100 K, k =

1.0 WIm.k; Pt catalyst 68

Figure 3.23 Bed Solid temperature at the outlet as afunction of porosity for

various inlet temperatures; u 1.0 m/s, 4) = 0.35, L = 75 mm, k

= 1.0 W/m.K; Pt catalyst 69

Figure 3.24 Temperature distribution along the bed for different bed lengths;

(a) L = 25 mm, (b) L = 50 mm, (c) L = 100 mm, 4) = 0.35, u=

1.0 m/s, a = 0.4, 750 K; homogeneous-heterogeneous

model, Pt catalyst 71

Figure3 .25 Temperature distribution for the gas (solid line) and the solid

(broken line) along the packed bed for various inlet temperature;

4) = 0.35, u= 1.0 mIs, k = 1.0 W/m.K, a 0.45, L = 100 mm;

homogeneous-heterogeneous model, Pt catalyst 73

Figure 3.26 Temperature distribution for the gas (solid line) and the solid

(broken line) for various bed porosities with 4) = 0.35, u = 1.0

m/s, ks = 1.0 W/m.K, Tin = 800K, L = loomm; homogeneous-

heterogeneous model, Pt catalyst 73

Figure 3.27 Temperature distribution for the gas (solid line) and the solid

(broken line) for various bed porosities with 4) 0.35, u = 1.0

xl"
m/s, k = 1.0 W/m.K, Ti
ff = 800 K, L = 50 mm; homogeneous-

heterogeneous model, Pt catalyst 74

Figure 3.28 Calculated methane conversion, as a function of inlet

- , temperature for various solid thermal conductivities; 4) = 0.35, u

1.0 m/s, L = 50 mm, s= 0.4; Pt catalyst 76 -

Figure 3.29 Methane mass fraction along the bed , at various inlet

temperatures for different solid thermal conductivities, 4) = 0.35,

u= 1.0 m/s; heterogeneous reaction, Pt catalyst 76

Figure 3.30 Methane mass fraction along the packed bed for different values

of thermal conductivities of the solid bed; 4) = 0.35, u= 1.0 m/s,

s= 0.4, Tffi = 750 K; homogeneous-heterogeneous reactions, Pt

catalyst 78

Figure 3.31 Temperature distribution for the gas (solid line) and the solid

(broken line) phases along the bed ,for different solid thermal

conductivities; 4) =0.35, u = 1.0 m/s, c = 0.4, Ti


ff = 750 K;

homogeneous-heterogeneous reactions, Pt catalyst 78

Figure 4.1 Schematic diagram of the experimental apparatus 81

Figure 4.2 A photographicview of the experimental apparatus ' 82

Figure 4.3 Cross-section view of the reactor 84

Figure 4.4 A photographic view of the' experimental packed bed 85

Figure 4.5 A photographic view of the choked nozzle nieters assembly 87

Figure 4.6 Exhaust gas sampling system with gas chromatograph 91

Figure 5.1 Comparison between experimental and numerical methane

conversion within the inert bed for 4) = 0.35, u= 1.0 m/s 100

Figure 5.2 Exhaust concentrations of CH4,CO2 and CO as a function of

inlet temperature within the inert bed for 4) = 0.35, u = 1.0

m/s 100

xiv
Figure 5.3 Effect of deactivation of fresh platinum catalyst on methane

conversion, = 0.35, u= 1.0 m/s 102

Figure 5.4 Exhaust concentrations of CH4 and CO 2 as afunction of inlet

temperature; fresh platinum catalyst, 0.35, u= 1.0 m/s 102

Figure 5.5 Exhaust gas temperature and methane conversion as afunction

of inlet temperature; fresh platinum catalyst, 4) = 0.35, u = 1.0

M/ -s 105

Figure 5.6 Comparison of methane conversion over fresh and aged

platinum catalyst, 4) = 0.35, u= 1.0 m/s 105

Figure 5.7 Exhaust gas temperature and methane conversion as afunction

of inlet temperature; fresh (5.0 h) platinum catalyst, 4) = 0.35, u

1.0 m/s 106

Figure 5.8 Methane conversion as a function of inlet temperature for

different equivalence ratios, u= 1.0 m/s; aged Pt catalyst 108

Figure 5.9 Methane conversion versus equivalence ratio at different inlet

temperatures, u= 1.0 m/s; aged Pt catalyst 108

Figure 5.10 Methane conversion as a function of inlet temperature at

different approach velocities with 4) = 0.35; aged Pt catalyst 110

Figure 5.11 Effect of approach velocity on methane conversion at different

inlet temperature with 4) = 0.35; aged Pt catalyst 110

Figure 5.12 Methane conversion as a function of inlet temperature within

the Co 3O4ICr2O3 catalytic bed, 4) = 0.35 and u= 1.0 m/s 112

Figure 5.13 Exhaust concentrations of CH4,CO and CO 2 as afunction of

inlet temperature; C0304/Cr2O3 catalyst, u = 1.0 m/s and 4) =

0.35 112

Figure 5.14 Methane conversion as a function of inlet temperature for

different equivalence ratios; C0304/Cr2O3 catalyst, u= 1.0 m/s 113

xv
Figure 5.15 Exhaust concentration of CO as afunction of inlet temperature

for different equivalence ratios; Q0304/Cr2O3 catalyst, u = 1.0

m/s 113
Figure 5.16 Comparison between Pt and C0304/Cr2O3 catalysts for Methane

conversion at different inlet temperatures, u= 1.0 m/s, 4) = 0.35 115

Figure 5.17 Exhaust concentration of CH4 and CO 2 as a function of inlet

temperature; Pt and C0304/Cr2O3 catalysts, u= 1.0 m/s, 4) = 0.35 115

Figure 6.1 Experimental data and predicted methane conversion as a

function of inlet temperature for various pre-exponential factors

and activation energies; platinum catalyst, u= 1.0 mIs, 4) = 0.35 120

Figure 6.2 Experimental and calculated methane conversion as afunction

of inlet temperature, platinum catalyst, u= 1.0 m/s, 4) = 0.35, E

= 46.0 k.T/mol for T1≥ 750 K, E= 42.0 kJ/mol for Tm< 750 K,

A3=1.03x10 3 kmol/m2.K 121


Figure 6.3 Experimental and calculated methane conversion as afunction

of inlet temperature for two equivalence ratios, platinum


'

catalyst, u= 1.0 m/s 121


Figure 6.4 Experimental and calculated methane conversion as afunction

of inlet temperature for two approach velocities, platinum


.

catalyst, 4) = 0.35 122

Figure 6.5 Experimental and calculated methane conversion as afunction

of inlet temperature, C0304/Cr2O3 catalyst, u = 1.0 m/s, 4) =

0.35, E = 42.8 kJ/mol for 775 K, E = 46.3 kJ/mol for Tin

> 775 K, A = 2.2 x i0 kmol/ m2.K 123


Figure 6.6 Experimental and numerical methane conversion as afunction

of the inlet temperature for two equivalence ratios, C0304/Cr2O3

catalyst, u= 1.0 m/s 124

xvi
Figure 6.7 Experimental and calculated methane conversion as afunction

of inlet temperature for two approach velocities, C0304/Cr2O3

catalyst, 4) = 0.35 124

Figure 6.8 Calculated methane conversion as a function of inlet

temperature for various bed porosities within the inert bed, 4) =

0.35,u= 1.0 m/s, k = 1.0W! m.K, L = 50 mm 126

Figure 6.9 Calculated methane conversion as a function of inlet

temperature for different bed lengths; s 0.4, 4) = 0.35, u= 1.0

mls,k5= 1.OW/m.K 126

Figure 6.10 Calculated methane conversion as a function of inlet

temperature for various solid thermal conductivities; 4) = 0.35, u

1.0mls,L=50 mm, c=0.4 127

Figure 6.11 Calculated methane conversion as a function of inlet

temperature for various bed porosities; 4) = 0.35, u= 1.0 m/s, k

= 1.0 Wlm.K, L = 50 mm; Pt catalyst 128

Figure 6.12 Calculated methane conversion as a function of inlet

temperature for different bed lengths; s= 0.5, 4) = 0.35, u= 1.0

m/s, k = ,I .0 W/ m.K; Pt catalyst 129

Figure 6.13 Calculated methane conversion as a function of inlet

temperature for various solid thermal conductivities; 4) = 0.35, u

= 1.0mls,L=50 mm, c=0.4;Ptcatalyst 129

Figure 7.1 Calculated fuel conversion as a function of inlet temperature

within an inert (* symbol) and catalytic packed beds, 4) = 0.35, u

1.0 m/s 135

Figure 7.2 Calculated ethane conversion as afunction of inlet temperature

within an inert bed for different equivalence ratios, u = 1.0 m/s,

k1.0W/m.K,L=50mm 135

xvii
Figure 7.3 Calculated ethane conversion as afunction of inlet temperature

for different equivalence ratios, u= 1.0 m/s, k = 1.0 W/m.K, L

= 50 mm; Pt catalyst 137

Figure 7.4 Calculated emperature distribution for the gas (solid line) and

the solid (broken line) along the bed for various equivalence

ratios of ethane, u= 1.0 m/s; Pt catalyst 137

Figure 7.5 Calculated ethane conversion as afunction of inlet temperature

within the catalytic (thin lines) and inert (*, thick lines) beds for

different approach velocities, 4=0.35 138

Figure 7.6 Calculated propane conversion as a function of inlet

temperature within the inert bed at different equivalence ratios,

u=1.0mIs,k1.0WIm.K,L=50min 140

Figure 7.7 Calculated propane conversion as a function of inlet

temperature at different equivalence ratios, u= 1.0 m/s, k = 1.0

WIm.K, L = 50 mm; Pt catalyst 140

Figure 7.8 Calculated temperature distribution for the gas (solid line) and

the solid (broken line) along the bed for various equivalence

ratios of propane, u= 1.0 m/s; Pt catalyst 142

Figure 7.9 Calculated propane conversion as a function of inlet

temperature within the catalytic (thin lines) and inert (*, thick

lines) beds for different approach velocities, = 0.35 142

Figure 7.10 Calculated fuel conversion as a function of inlet temperature

within catalytic (thick lines) and inert (*, thin lines) beds, 4 =

0.35, u= 1.0 mIs 143

Figure 7.11 Calculated carbon monoxide oxidation as a function of inlet

temperature within the inert bed for various equivalence ratios,

u1.0m/s 145

Figure 7.12 Calculated carbon monoxide oxidation as a function of inlet

temperature for various equivalence ratios, u = 1.0 m/s; Pt

catalyst 145

xviii
Figure 7.13 Temperature distribution for the gas (solid line) and the solid

(broken line) along the catalytic bed for various equivalence

ratios of carbon monoxide; u= 1.0 m/s; Pt catalyst 146

Figure 7.14 Effect of approach velocity on calculated, carbon monoxide

conversion as a function of inlet temperature within the inert

bed, 4)= 0.35 147

Figure 7.15 Effect of approach velocity on calculated carbon monoxide

conversion as a function of inlet temperature, 4) = 0.35; Pt

catalyst 147

Figure 7.16 Calculated hydrogen conversion as a function of inlet

temperature at different equivalence ratios within the inert bed,

u=1.0m/s.... 149

Figure 7.17 Calculated hydrogen conversion as a function of inlet

temperature at different equivalence ratios, u = 1.0 m/s; Pt

catalyst 149

Figure 7.18 Comparison between calculated hydrogen conversion within the

catalytic and inert beds (*; thick lines) versus inlet temperatures

for various ëquivalénce ratios, u= 1.0 m/s 151

Figure 7.19 Calculated temp'eràture distribution for the gas (solid line) and

the solid (broken line) ,along the bed for various equivalence

ratios of hydrogen; u= 1.0 m/s; Pt catalyst 151

Figure 7.20 Calculated temperature distribution for the gas (solid line) and

the solid (broken line) for CH4,C2H, C3H, CO, and H2 along

the bed, 4) = 0.35, u= 1.0 m/s; Pt catalyst 152

Figure 7.21 Calculated hydrogen conversion as a function of inlet

temperature within the inert bed for different approach

velocities, 4) = 0.10 . 152

xix
Figure 7.22 Calculated hydrogen conversion as a function of inlet

temperature within the inert bed for different approach

velocities, = 0.35 . 153

Figure 7.23 Calculated hydrogen conversion as a function of inlet

temperature for different approach velocities, = 0.35; Pt

catalyst 154

Figure 7.24 Comparison between calculated hydrogen conversion within the

catalytic and inert beds (*; thick lines) versus inlet temperature

for various approach velocities, = 0.35 154

Figure 8.1 Calculated and experimental propane conversion as a function

of inlet temperature within platinum bed, 4i = 0.35, u= 1.0 m/s 156

Figure 8.2 Calculated and experimental propane conversion as afunction

of inlet temperature within C0304/Cr2O3 bed, 4i = 0.35, u= 1.0

rn/s 156

Figure 8.3 Calculated and experimental carbon monoxide conversion as a

function of inlet temperature within platinum bed, = 0.35, u=

1.0 mIs 158

Figure 8.4 Calculated and experimental carbon monoxide conversion as a

function of inlet temperature within C0304/Cr2O3 bed, 4= 0.35,

u=1.0m/s 158

Figure 8.5 Calculated and experimental hydrogen conversion as afunction

of inlet temperature within platinum bed, = 0.35, u= 1.0 mIs 160

Figure 8.6 Calculated and experimental hydrogen conversion as afunction

of inlet temperature within C0304/Cr2O3 bed, = 0.35, u= 1.0

rn/s 160

Figure 8.7 Comparison of calculated propane conversion as afunction of

inlet temperature within platinum bed, 4 = 0.35, u= 1.0 m/s 164

xx
Figure 8.8 Comparison of calculated carbon monoxide conversion as a

function of inlet temperature within platinum bed, 4) = 0.35, u=

1.0rnls 164

Figure 8.9 Comparison of calculated hydrogen conversion as afunction of

inlet temperature within platinum bed, 4) = 0.35, u= 1.0 m/s 165

Figure B. 1 Calibration curves for 0.09mm diameter nozzle 192

Figure B.2 Calibration curves for 0.20 mm diameter nozzle 193

Figure B.3 Calibration curves for 0.34 mm diameter nozzle 193

Figure B.4 Calibration curves for 0.61 mm diameter nozzle 194

Figure B.5 Calibration curves for 1.49 mm diameter nozzle 194

Figure C. 1 Temperature distribution 'across the reactor,Tin 41 °C ' 195

Figure D. 1 Pressure drop across the packed bed reactor as afunction of the

flow rate at Tin=20 °C and P = 88.1 kPa 196

Figure D.2 Pressure drop across the packed bed reactor as afunction of the

approach velocity at Tin,


= 20 °C and P = 88.1 kPa 197

xxi
LIST OF SYMBOLS

Symbols Descripton

A pre-exponential factor (kmollm3.$)

a catalyst activity

catalyst spepific surface area (m2/m)

C concentration

Cp g specific heat of the mixture at constant pressure (J/kg.K)

specific heat of pellets (J/kg.K)

D Diffusion coefficient (m2/s)

d pellet diameter (mm)

E activation energy (kJ/mol)

Hk specific enthalpy of species k(J/kg)

h heat transfer coefficient (W/m2.K)

hv volumetric heat transfer coefficient (W/m3 .K)

k thermal conductivity (W/m.K)

effective thermal conductivity (W/m.K)

L length of the bed(mm)

MW molecular weight (kg/kmol)

MW molecular weight of the mixture (kglkmol)

Nu Nusselt number

P pressure (kPa)

Pr Prandtl number

Re
Reynolds number

Ru universal gas constant (kJ/kmol.K) -

SP surface area of apellet (mm2)

T temperature (K)

t time (s)
Q volume flow rate (rn3/h)
u approach velocity (m/s)
V volume (m)

V 3
volume of apellet (mm)
-

V3 space velocity (m3/(m3.h))

X axial coordinate (m)

Y mass fraction

Greek symbols

/3 extinction coefficient (1/m)


e porosity

equivalence ratio

viscosity (kg/m.$)
V
kinematic viscosity (m2/s)

molar consumption rate per unit volume for gas phase reactions

(kmol/m 3 s)

molar cons
umption rate per unit area of the catalyst for surface
reactions (kmol/m2.$)

mixture density (kg/m3)

Stefan-Boltzman constant (W/m2.K4)

Subscripts

CH 4 methane

cony conversion

eff effective

f fuel

g gas

in inlet, initial
k species

o oxidant

0 2 oxygen

out outlet

p pellet

s solid, surface

Superscripts

h gas-phase

m, n constants(Eq.' 2.11)

I
s catalytic surface

xxiv
1

CHAPTER 1

INTRODUCTION

The present experimental and numerical work is intended to address the catalytic

combustion in apacked bed reactor. The developed mathematical model is based on one

dimensional flow, heat and mass transfer with global reaction, which is adequate enough
to produce dependable results for design purposes. Moreover, alab scale experimental set

up is constructed with necessary measuring instrumentations. The experiments were

conducted for two reasons; to validate the developed model and to assist in establishing

the kinetic data for global surface reaction over platinum and C0304/Cr2O3 catalysts for

the gaseous fuels employed within the packed bed reactor. Also, the model is used to

investigate the effects of controlling parameters, such as inlet temperature, inlet velocity,

bed porosity and bed thermal conductivity on the characteristics of the homogenous and

heterogamous combustion. The model provided areasonable study of the characteristics

of the catalytic reactor.

The following paragraphs provide abackground to the physics and mechanisms of the

catalytic process, which is essential in developing the mathematical model. It also

introduces terminologies used in the catalytic literature. Finally, the objectives of the
present work and outlines of the thesis are introduced.

1.1 Background

Most industrial syntheses and nearly all biological reactions require catalysts.

Furthermore, catalysis is the most important technology in environmental protection, i. e.,


2

the reduction of emissions. A well-known example is the catalytic converter for

automobiles. The term "catalysis" was introduced as early as 1836 by Berzelius and a

definition that is still valid today that of Ostwald (1985): "a catalyst accelerates a

chemical reaction without affecting the position of the equilibrium". While it was

formerly assumed that the catalyst remained unchanged in the course of the reaction, it is

now known that the catalyst is involved in chemical bonding with the reactants during the
catalytic process. Thus, catalysis is acyclic process: the reactants are bound to one form

of the catalyst, and the products are released from another, regenerating the initial state.

In theory, an ideal catalyst would not need replenishing, but this is not the case in
practice. Owing to competing reactions, the catalyst undergoes chemical and physical

changes, and its activity decreases (catalyst deactivation). Thus catalysts must be

regenerated or eventually replaced.

Apart from accelerating reactions, catalysts have another important property: they

can influence the selectivity of chemical reactions. This means that completely different

products can be obtained from a given starting material by using different catalyst

materials. Industrially, this targeted reaction control is even more important than the

catalytic activity (Hagen, 1999).

Catalysts can be in gaseous, liquid, or solid state. Most industrial catalysts are

liquids or solids. Nuhierous organic intermediate products required for the production of

plastics, synthetic fibers, pharmaceuticals, dyes, crop-protection agents, resins, and

pigments, can only be produced by catalytic processes. Most of the processes involved in

crude oil processing and petrochemistry, such as purification, refining, and chemical

transformations, require catalysts. Environmental protection measures such as automobile

exhaust control and purification of off-gases from power stations and industrial plants

would be much less effective without catalysts.

Catalysts can be classified, according to various criteria such as structure,

composition, area of application or state of aggregation. Figure 1.1 represents the

classification of catalysts according to the state of aggregation in which they act.


3

Catalysts

Homogeneous Heterogeneous Heterogeneous


catalysts homogeneous catalysts
catalysts
[i ncl
.biocatalysts
Acid/base (enzymes)] Bulk
catalysts catalysts

Transition Supported
metal catalysts
compounds

Fig. 1.1 Classification of catalysts (Hagen, 1999).

Heterogeneous catalysis takes place between different phases. Generally the

catalyst is a solid, and the reactants are gases or liquids. In supported catalysts, the

catalytically active substance is applied to asupport material that has alarge surface area
and is usually porous (Hagen, 1999).

The suitability of a catalyst for an industrial process depends mainly on the

following three properties: (1) Activity, (2) Selectivity, and (3) Stability.

The activity of a catalyst is a measure of how fast one or more reactions

proceed in the presence of the catalyst. Activity can be defined in terms of

kinetics. Mostly, the activity of the catalyst is determined by measuring the

reaction rates in the temperature and concentration ranges that will be present

in the reactor. However in practice, the activity of the catalyst e.g., for catalyst

screening, determination of process parameters, optimization of catalyst

production conditions, and deactivation studies, is measured by one of the

following activity measures:


4

o Conversion under constant reaction conditions

o Space velocity for agiven, constant conversion 0

o Space-time yield

o Temperature required for agiven conversion.

• The selectivity of acatalyst of areaction is the ability of the catalyst to give

the desired product, out of all possible products, e.g. the percentage of the

starting material (reactant) that is converted to the desired product. The

selectivity is of great importance in industrial catalysis, as demonstrated by

the example of synthesis gas chemistry, in which, depending on the catalyst

used, completely different reaction products may be obtained (Fig. 1.2).

• The stability (deactivation behavior) of a catalyst refers to its chemical,

thermal, and mechanical stability. This determines its operational lifetime in

industrial reactors. Catalyst stability is influenced by numerous factors,

including its decomposition, coking, and poisoning. Catalyst deactivation can

be estimated by measuring the activity or selectivity as a function of time.

Catalysts that lose activity during processes can often be regenerated before

they ultimately have to be replaced. The total catalyst lifetime is of crucial

importance for the economics of aprocess. The mechanisms involved in the

deactivation and regeneration of catalytic surfaces will be reviewed next in

order to better understand the concept involved in such processes.

Generally, it is difficult to choose which of the above functions (properties) is the

most important because the demands made on the catalyst are different for each process
or application.
5

Ni CH+HO
o Methanization

Cu/Cr/Zn oxide 11 CH, OH Methanol synthesis


CO/H2
Fe, Co CnH n+m + H20 Fischer-Tropsch
Synthesis
Rh cluster CH2-CH2
Glycol
'oH OH

Fig. 1.2 Reactions of synthesis gas (Hagen, 1999).

• Catalyst Deactivation

Catalysts have only alimited lifetime. Some lose their activity after afew minutes, others

last for more than ten years. The maintenance of catalyst activity for as long as possible is

of major economic importance in industry. Catalyst deactivation, also known as aging, is

expressed by the decrease with time in the catalyst activity defined as the ratio of the

reaction rate at agiven time to the reaction rate at the time (t = 0) and is presented as

follows:

)(t = 0)

where, ais the catalyst activity, t is the time and w is the reaction rate of -
the catalyst

under the same condition. Not only does the decreasing catalyst activity lead to aloss of

productivity, it is. also often accompanied by a lowering of the catalyst selectivity.

Therefore, in industrial processes great efforts are made to avoid catalyst deactivation or

to regenerate deactivated catalyst.


6

A decline in activity during the process can be the result of various physical and chemical

factors. The four most common causes of catalyst deactivation are:

1. Poisoning of the catalyst. Typical poisons are H2S, Pb, Hg, S. and P.

2. Blocking of the active sites by deposits on the catalyst surface and change of the

pore structure (e.g., coking).

3. Sintering due to over heating of the catalyst leading to aloss of active surface

area.

4. Loss of the catalyst (e.g. through, formation of volatile metal carbonyls with CO).

These processes are described below and are shown schematically in Fig. 1.3 (Hagen,

1999).

o Catalyst Poisoning

Catalyst poisoning is a chemical effect. Catalyst poisons form strong adsorptive

bonds with the catalyst surface, blocking active centers, Therefore, even very small

quantities of the catalyst poisons can influence the adsorption of reactants on the

catalyst. The term catalyst poison is usually applied to foreign materials in the

reaction system.

o Deposits on the catalyst surface

The blocking of catalyst pores by polymeric components, especially coke, is another

widely encountered cause of catalyst deactivation. In many reactions of

hydrocarbons, side reactions lead to formation of polymers. If these are deposited

near the pore opening, catalyst activity and selectivity can be influenced due to
impaired mass transport into and out of the pores. At high temperatures (above

200 °C) these polymers are dehydrogenated to carbons, aprocess known as coking.

Catalysts with acidic or hydrogenating/dehydrogenating properties (e.g. metal,

oxides, and sulfides catalysts in catalytic reforming) are especially susceptible to

coking. The extent of coke formation depends directly on the acidity of the catalyst.
.7

o Thermal processes and sintering

Sintering is defined as the loss of active surface area by the agglomeration of small

crystallites into larger ones in the case of supported metal catalysts, or the collapse of

pore structure and loss of internal surface area in the case of the supports or various

unsupported (normally oxide) catalysts. Sintering is athermally activated process and

is physical rather than chemical in nature. The transformation of 7-.A1203 into a-

A1203 with its lower surface area is one example of sintering of the support due to

phase transformation. The rate of sintering increases with increasing tempeiature,


decreasing crystallite size, and increasing contact between the crystallite particles.

Phase changes due to overheating can limit the catalyst activity or lead to catalyst-

substrate interactions. However, achange in selectivity can also occur, especially in

the case of structure-sensitive .reactions.

o Catalyst losses via the gas phase.

High reaction temperatures in catalytic processes cafi lead to the loss of active

components by evaporation via processes such as volatilization. This does not only,

occur with compounds that are known to. be volatile (e.g., P205 in H3PO 4,silica

gel/activated carbon), it also occurs when metals react to give volatile oxides,

chlorides, or carbonyls (Hagen, 1999).


8

Fig. 1.3 Mechanisms of catalyst deactivation (M = metal) (Hagen, 1999).

In general, acatalyst must be chemically stable under actual operating conditions

(over a wide range of temperatures, air/fuel ratios, flow rates etc.) and resistant to

poisoning. Catalysts must also pose no potential danger to the environment (such as the

emission of finely dispersed particles of metals or their toxic derivatives) (Prasad et al.,
1984).

Environmental pollution and air quality are certainly a worldwide concern.

Combustion devices and processes contribute very, significantly to pollution through

exhaust emissions (CO, NO,,, unburned hydrocarbons and particulates, SO2, etc)

promoting the formation of smog, and acid rain and contributing to the greenhouse effect,
9

which has been liable for a general rise in global temperature and increased weather

variations (Prather, and Logan, 1994). These pollutants can cause asignificant hazard to

human and animal health as well as agricultural and economical impact. Seinfeld (1986)

indicated four principal effects of air pollution in the troposphere: (1) altered properties

of the atmosphere 'and precipitation; (2) harm to vegetation; (3) soiling and deterioration

of materials; and (4) potential increase in morbidity and mortality -in humans. The public

has an increased awareness of the necessity for controlling emissions for preserving our

planet. Governments around the world, especially those of advanced industrial nations,

are responding with increased regulations for emission control from mobile and
stationary sources. Ambient air quality standards were legislated and were put in force in

the early 1970s in Canada, U.S.A., Europe and Japan. Amendments to these later on have

generally imposed more strict standards and have brought more and more sources of air

borne pollutants under scrutiny (Koshland, 1996). In fact, the 90s has being called the

"decade of the environment". A concept known as "polluter pays principle", (PPP), has

been established in Canada, U.S.A., Japan, Europe and other countries (Nakajima, 1991;

Heck and Farrauto, 1994).

There are two approaches to consider in reducing undesirable emissions. They

are: to treat the pollutant being emitted or, better still, not to produce the pollutant at all!

Often the former is more realistic because there is no easy and economically acceptable

approach which avoids producing the pollutant with present day technology.

Catalysts in emission control processes have to be adaptable to the conditions of

the effluent gas to be treated. It is difficult to employ pre-treatment devices for the

removal of catalyst poison and dust, because they make the processes complicated and
costly.

One approach used to enhance the reactivity of fuels and to decrease emissions is the use

of catalytic combustion. Catalytic combustion enables the oxidizing of fuels very

efficiently under lean conditions (i.e. below the lean flammablity limit for fuel-air

mixtures) minimizing the emissions of unburned hydrocarbons (LJHC) and CO. This

takes place at relatively low temperatures and results in decreased formation of NO,,.
10

Therefore, catalytic combustion offers aclear opportunity to provide solutions to many


environmental issues.

The drive to lower pollutant emissions from combustion processes has led not

only to increased research in catalytic combustion but also to increased research in

alternative sources of energy and fuels such as natural gas, hydrogen, CO, gasified

biomass, etc.. Natural gas, which mainly consists of methane, has received increased

attention as an alternative fuel for internal combustion engines, commercial vehicles, gas

turbines, industrial furnaces and boilers because of its environmental advantages.

However, methane tends not to react relatively easily with air, especially, at low
.

temperatures. The presence of catalysts can permits the oxidation of methane at lower

temperatures and with faster rates than in their absence. ,Because a great amount of

natural gas, is utilized in many practical energy conversion systems, a better

understanding of, its catalytic oxidation will benefit combustion systems employing

natural gas by achieving more effective fuel utilization, reducing pollutant emissions,

improving the production of synthesis gas and reducing safety hazards.

Cars and stationary combustion facilities such as power plants and industrial
boilers are identified as amajor source of emissions of HC, CO, SO 2 particulates and

NO, to the atmosphere. Nitrogen oxides are formed in practical combustion systems by

high temperature oxidation of atmospheric nitrogen and nitrogen bound in the fuel.

Catalytic oxidation at low temperatures is awell et.blished technique to limit NO

production (Cybulski, and Moulijn, 1998). The cost of the catalyst, which in most cases
contains noble metals, and its susceptibility to aging or damage caused by thermal stress,

sintering and poisoning, have so far slowed down the widespread use of catalytic

combustion. Despite this, increasingly severe regulations on combustion emissions have

motivated research oriented toward the development of cheaper, regenerative and durable

catalytic materials. Thus, many new catalysts have been discovered and we may expect

more to be 'discovered. The development of catalysts which can withstand high

combustion temperature and which are inexpensive is an important area of research


(Zwinkels et al., 1993; Bozo et al., 1999).
11

In many of the applications mentioned earlier, the questions that often remain to be

answered are:

1. How much catalyst is needed in order to achieve the required stability of catalyst,

desired emissions, and desired heat transfer characteristics?

2. Is the catalyst only needed to ignite the mixture and does it then remain unused?

3. If it remains active in the higher temperature range, what is its role and in which

part of the reactor it is needed?

The following chapters attempt to answer some of these questions;


12

1.2 Objectives of the Study

The present research is related to catalytic oxidation of gaseous fuel-air mixtures in a

packed bed reactor. This study was initiated and designed to determine the kinetic data

(activation energy and the pre-exponential factor) for catalytic oxidation of common

gaseous fuels over non-noble metals (binary mixtures of cobalt and chromium oxides) as

well as platinum catalysts. Therefore, amathematical model and experimental program

were developed to investigate the performance of these catalysts in the oxidation of lean
premixed common gaseous fuel-air mixtures in apacked bed reactor.

The specific tasks of the present research are as follows:

1. To develop atheoretical model for the catalytic oxidation of fuel-air mixtures in

apacked bed reactor.

2. To investigate experimentally the effectiveness of platinum as well as anon-

noble binary mixture ,of cobalt and chromium oxides as catalysts in the

oxidation of lean homogeneous mixtures of gaseous fuel and air at atmospheric

pressure and different operational conditions in apacked bed reactor.

3. To validate the theoretical model against the experimental data obtained in this
investigation as well as against other published data.

4. To obtain the kinetic data for abinary mixture of cobalt and chromium oxides

• catalyst as well splatinum in apacked bed reactor.

5. To investigate the effect of bed porosity s, length of the. bed L, and solid thermal

conductivity kon the conversion of methane in apacked bed reactor.


13

1.3 Thesis Organization

The outline of the remaining chapters of the dissertation is as follows:

- Chapter 2: Covers the literature review.

- Chapter 3: Covers the physical and mathematical descriptions of the problem

under consideration and the detailed steps used to develop the mathematical

model. It also presents preliminary results and discussions of the model's

prediction of the oxidation of methane in air and compares the model's prediôtion

with any available corresponding experimental data. -

- Chapter 4: Desctibes the experimental apparatus and procedure used throughout

the course of this study.

- Chapter 5: Presents the experimental results for methane in air within the inert

and catalytic (platinum or binary mixture of cobalt and chromium oxides) packed
bed reactors.

- Chapter 6: Presents the kinetic data obtained for methane oxidation in air within

the inert and catalytic (platinum or binary mixture of cobalt and chromiuth
oxides) packed bed reactor.

- Chapter 7: Considers the catalytic oxidation of other commOn gaseous fuels (i.e.

ethane, propane, carbon monoxide, and hydrogen) in air.

- Chapter 8: Covers comparison of model prediction with the experimental data of

common gaseous fuels (i.e., propane, carbon monoxide, and hydrogen) in air in

order to determine the surface kinetic data for these fuels within the platinum or

the binary mixture of cobalt and chromium oxides catalysts.

- Chapter 9: Summarizes the relevant conclusions drawn from the study and

suggests some recommendations for possible future work.


14

CHAPTER 2

LITERATURE REVIEW

2.1 Introduction

Catalytic combustion has been vigorously explored as an effective route to energy

conversion in view of its capability to oxidize different fuels with high efficiency at

relatively low temperatures and under lean conditions, thus simultaneously preventing the

formation of NOx and minimizing the emissions of unburned hydrocarbons (UHC) and

CO. Extensive research on catalytic combustion has been conducted experimentally and

theoretically. A review of some relevant research is presented below.

2.2 Experimental Work

Experiments have been carried out on catalytic combustion for over 160 years.

Interestingly, the first use of catalysis was to promote combustion. Davy discovered in

1840 that platinum wires could induce combustion of flammable mixtures. The resulting

reactions appeared to take place on the surface of the wire, "without flame" and with the

high radiative fluxes associated with the solid surface emittance rather than the low

emittance characteristic of gaseous combustion products (Kesselring 1986). Davy's

observations led to the development of numerous radiative heater devices and, thus, the

first practical use of catalysis (Kendall et al., 1992). This discovery stimulated numerous

experiments throughout the late 19th century which demonstrated that various materials

could catalyze heterogeneous combustion.


15

2.2.1 Fundamental Studies

Some of the studies and investigations which might be termed "fundamental studies of

heterogeneous catalysis in combustion" have attempted to: identify specific.

heterogeneous reactions, quantify reaction kinetics for these reactions, define the

mechanisms by which the solid surface promotes or accelerates oxidation reactions and

identify catalyst properties leading to high activity.

Hiam et al, (1968) obtained aformula for the oxidation of propane-air mixtures

over platinum catalyst. Trimm and Lam (1980) studied the oxidation of methane over

platinum supported on porous and non-porous alumina fibres. They observed permanent

catalyst deactivation due to sintering of platinum or alumina; the latter effect was

accelerated in the presence of steam. Fresh and aged catalysts were characterized in their

investigation. Trimm and Lam also proposed global rate expressions for the oxidation of

methane for two temperature regimes, below and abpve 813 K. They suggested that the

formation of intermediates may affect the reaction selectivity from one in which oxygen

adsorption is predominant to one in which methane adsorption is most important. Otto

(1989) investigated methane oxidation over platinum/7-alumina. Oxidation was found to

be structure sensitive; this was reflected in the change in the apparent activation energy.

They reported that methane oxidation kinetics were consistent with the existence of at

least two different platinum entities, termed dispersed and particulate platinum oxide,

which were characterized by differences in particle size and oxidation state. Otto

concluded that the reaction kinetics of methane oxidation depend strongly on the

pretreatment of the platinum samples since oxidation tends to disperse platinum while

reduction can cause sintering. Song et al. (1991) studied the ignition and extinction of

lean methane or propane in air on platinum foil in stagnation-point flow. They developed

an irreversible one-step reaction rate expression able to show that the heterogeneous

reaction dominates the system behavior at the lower temperature while both

homogeneous and heterogeneous reactions play important roles in the higher temperature

range. It was also predicted that the homogeneous ignition temperature was higher with

the surface reaction than without the surface reaction. This is because the homogeneous-
16

heterogeneous reactions expand the stabilized operating regions with high reaction rates

compared with either the heterogeneous or the homogeneous reaction alone.

Voltz et al. (1973) determined the rate constants and activation energies of carbon

monoxide and propylene oxidation on apelleted platinum-alumina catalyst with synthetic

gas mixtures (i.e. ablend, of pure CO, C3H8,02, NO, CO 2,and N2)between 477 and 644

K to simulate a real ,engine's exhaust conditions. The relative importance of catalyst

density and heat capacity, catalyst activity, initial bed temperature, inlet exhaust gas

composition and temperature, and converter location were studied and it was found that

they could change depending on the converter systems and test procedure (Kuo et al.,

1980). It was found also that the exhaust gas flow rate, heat loss from the converter,

oxygen concentration in excess of 2.0%, and gas channeling in the catalyst bed were of

minor importance. Kuo et al. estimated the kinetic parameters of the global reaction of

CO and HC from their experimental and theoretical investigations of the operational

catalytic converter and later used them in modeling the oxidation of CO.

Pfefferle et al. (1989a) investigated the effect of catalytic activity on gas ph ase

ignition in an ethane-air boundary layer by measuring OH concentration profiles over


heated platinum and quartz surfaces. They found that platinum catalytic influence on gas

phase ignition depends strongly on the fuel-air ratio. Depletion of fuel near the catalytic

surface by mass transfer-limited oxidation reactions can account for the catalyst

inhibiting effect. Also, their results indicated that catalytic surface-induced production of

active intermediates can promote gas phase ignition and offset the effect of reactant
depletion near the surface. Pfefferle et al. (1989b) showed that aplatinum catalyst acts to

promote the formation of oxygen atoms in the region adjacent to the surface. They

concluded that such effect was most apparent at lean equivalence ratios (4 <0.30 for

ethane). It was also reported that at higher equivalence ratios the ignition of ethane in air

at atmospheric pressure was inhibited by the platinum surface because of fuel depletion at

the surface due to the mass transfer-limited catalytic oxidation reaction. However, at low

pressures ranging from 5.3 to 18.7 kPa, ethane ignition was promoted by catalytic
reaction at all equivalence ratios studied (Griffin et al., 1992). Furthermore, it was shown

that N2O substitution for 02 strongly inhibited the gaseous ignition of ethane, hydrogen,
17

and acetylene over platinum surfaces, and suppressed the preignition level of OH radical
concentrations to avalue below their detection limits. Pitchai and Klier (1986) reported

that formaldehyde is an important combustion intermediate in both gas-phase and

catalytic-surface combustion. They suggested that the apparent formation of OH radicals


could result from formaldehyde production on the surface and its subsequent oxidation in

the boundary layer.

2.2.2 Applications

•Some studies reported in the literature as applied research could be called "reduction to

practice", these include attempts to construct and employ practical devices exploiting

heterogeneous combustion. In fact, there are many devices under development employing

catalytic combustion, such as domestic water heaters (Seonhi, and Scholten, 1997;
Ragaini et al., 1999), cookers (Scholten et al., 1999), and space and process heaters (Cho

et al., 1999; Kang et al., 1999); some of them are already on the market. Moreover, gas

turbines with catalytic combustors are currently in the final stages of testing for durability

such as in the U.S. Kang, 1999; Beebe et al., 1999; Smith et al., 1999) and Japan
(Tochihara and Ozawa, 1999).

Unlike radiant catalytic heaters, adiabatic catalytic oxidation systems have been

developed relatively recently, within the list 40 years. With the ability to burn very low

concentrations of combustibles at low temperatures in the presence of comparably low

concentrations of oxygen (outside of the homogeneous flammable limits), catalytic fume

abatement reactors, automotive catalytic converters, and nitric acid plant tail gas reactors

have found widespread use. In fact, such applications to date present one of the largest

single markets for catalysts. A major advantage of these devices is their ability to burn

fuels with negligible NO formation. Catalytic reactors have also been used to destroy

preexisting NO (three-way automotive catalytic reactors are one example).

Will and Bennett (1992) studied the effect of flow maldistributions on emission

control in automotive converter canisters for inlet velocities from 1to 15 m/s and catalyst
18

specific space velocities ranging from 11.1 to 22.2 (us) at ambient pressure. Their results

show anoticeable effect of the velocity on the emission conversion. P'attas, et al. (1994)

used aCFR engine to measure NO and matched it to the predicted results of their model

in order to obtain the kinetic data for NO reduction. The formation and destruction of

NO in acatalytic combustor of methane was studied by Agarwal et al. (1994) using a

series of Lai..Sr-
.Sr,, (La is Lanthanum; Sr is Strontium where both are catalyst

promoters) catalysts. The emission of NO was less than 3.0 ppm under lean combustion

conditions (3.8% CH4 in air) at temperatures as high as 1000°C. They attributed the low

NO emission during the lean catalytic combustion to the low temperature. Li and Li

(1995) studied three-way reactions to remove CO, BC, and NO,,. They showed that the

three-way performance depended not only on thç reaction of CO and HC with 02, but

also on the number and strength of NO adsorption sites. They concluded that the three-

way noble metal catalysts reactions for the removal of CO, UHC, and NO Were

controlled by internal diffusion at high space velocity of 44.4 (its). However, with non-

noble metal catalysts (La, Ce, Cu, Co, Mn, and Fe) there are 'internal diffusion controls at

space velocities less than 11.1 (its) and kinetic control at above there space velocity.

Moreover, Li and Li (1996) used rare earth (RE)-transition metal oxides for
simultaneously removing three major pollutants, CO, BC, and NO in automotive

emissions. They also showed that acatalyst of RE-transition metal with the following

composition RE(3%), Ni(2%), Co(3%), Mn(1%) has awider "Three-Way Window" than

acatalyst with acomposition of RE(3%)with Cu(4%) and Mn(2%) or Fe(2%). Schlegel

et al. (1996) investigated experimentally as well as numerically the NO emissions from


both catalytic and non-catalytic, lean premixed combustion of methane at atmospheric

pressure. This study reported that the higher the catalytic conversion of methane was, the

lower the NO emissions were. The NO level in catalytic combustor was significantly

lower than those of non-catalytic combustor. Corma et al. (1998) studied NO reduction
with propane using copper or cobalt on metal exchange zeolites. It was reported that this

catalyst has agood stability.


19

2.2.3 Non-Noble Metal Catalysts

Non-noble metal catalysts have gained agreat deal of interest in catalytic combustion

research as alternatives to very costly noble metal catalysts. Jones (1997) tested different

transition metal oxides (MoO 3,Cr2O3,Co3O4,Fe2O3,Y203,CeO 2,m-ZrO2,and MgO) in

the combustion of propane-air mixtures and showed that their light-off temperature

values were comparable with that of platinum catalyst. The temperature at which a

catalyst begins to operate is called the light-off temperature where the combustion

process changes from kinetically controlled regime (corresponding to 10% conversion

of the fuel) to mass transfer controlled regime (corresponding to —90% conversion of the

fuel) (Cybulski, and Moulijn, 1994). It was reported in the literature that metal oxides of

cobalt and chromium and their mixture'


shave showed high activity for the oxidation of

light hydrocarbons (Prasad et al., 1980; Depiak, 2003). The main advantage of metal

oxide catalysts over noble metal catalysts is the lower cost and high thermal stability over

arange of temperatures, which are typical for catalytic combustors (Sokolovskii, 1990;

Zwinkels et al., 1993). Also, interest in the catalytic combustion of alternative fuels such

as heavy distillates, low Btu gas and synthetic fuels directed investigations into the

oxidation of fuel-bound nitrogen (fuel-N) in acatalytic combustor (Prasad et al., 1981).

Under lean conditions, catalytic combustors using noble metal catalysts produce very

high levels of fuel-N conversion to NO (50 - 90%). It is evident from aliterature survey

that the formation of nitrogen oxides from fuel-bound nitrogen may be suppressed by

using metal oxide catalysts (Ismagilov and Kerzhentsev, 1990). Prasad et al. reported

1981) that transition metal oxide (C0304/Cr2O3) catalysts yield very low levels of fuel-N

conversion (10 - 20%), in the lean fuel range of operation (4 = 0.4). Non-noble catalysts

(such as MoO3,Cr2O3,Co 3O4,Co/ZSM-5, CuZSM-5, LaMnO 3,V205,NiO, CuO, and


C0304/Cr2O3) were efficient in the oxidation of very lean methane-air mixtures

(Blazowski and Walsh, 1975; Quinlan et al., 1986; Cybulski and Moulijn, 1994; Li and

Li, 1996; Depiak and Wierzba, 1999; Cerri et al., 2000), and have shown enhanced

activity for CO oxidation, and NO reduction (Bharadwaj and Schmidt, 1994; Adelman et

al., 1996; Desai et al., 1999). These investigations focused on methane oxidation over
20

various non-noble metals catalysts. There is aneed to examine further the effectiveness

of such catalysts in relation to their applications to the combustion of natural gas, carbon

monoxide, hydrogen, etc. in order to obtain kinetic coefficients for these fuels over non-

noble metal catalysts.

2.2.4 Catalytic Oxidation of Some Common Gaseous Fuels

The ability of catalysts to oxidize various fuels to products of combustion has been

studied extensively. High combustion efficiency results in both minimum fuel


consumption and low pollutant emissions of CO through the complete oxidation of the

fuel to CO 2 and 1120. However, catalytic combustion efficiency depends on flow

conditions of the premixed reactants such as stoichiometry, pressure, inlet temperature

and velocity as well as catalyst configuration (monolithic structure, fibre burners, ceramic

cylinders, and pellet beds, etc.) and activity. The parametric study of Prasad et al. (1981)

on the performance characteristics of binary transition metal oxides (Co304,Cr203)

supported on alumina pellets using lean propane-air mixtures showed that the conversion

efficiency at the exit was very sensitive to small changes in the inlet conditions. During

the experiment they measured CO, UHC, and CO 2 at the exit from the reactor and

calculated the combustion efficiency on the basis of carbon balance. It was found that the
combustor performance was more sensitive to temperature than it was to the inlet

velocity or the equivalence ratio; a 1.0 percent change in the inlet temperature could

cause a50% change in overall propane conversion.

CH4 Oxidation: Cullis and Willatt (1983) found that the conversion of methane was

lower over aplatinum catalyst than over apalladium catalyst. They also showed that the

conversion efficiency -of methane over platinum or palladium could be affected by the

catalyst supports materials. Oh et al. (1991) conducted aset of experiments to show the

activity of agroup of noble metal catalysts (Rh, Pd, and Pt) in methane oxidation with or

without cerium (Ce) additives. They reported that the complete methane oxidation

activity ranking was Pd > Rh > Pt in the absence of Ce and was Rh> Pd Pt in the
21

presence of Ce. However, under reducing conditions in the absence of Ce, methane

oxidation activity decreased in the order Pd > Rh > Pt, with the tendency to form CO

decreasing in the order Rh > Pd > Pt. The activity of the catalyst depends on the fuel type

too as illustrated by Depiak and Wierzba (2003). The oxidation of several fuels such as

CR4,C2R4, C3H8,
CO, and H2 over platinum as well as over binary C0304/Cr2O3 catalysts

was studied. It was shown that aplatinum catalyst was superior to aC0304/Cr2O3 catalyst
for all the fuels employed in their study. Moreover, the conversion efficiency of the fuel

based, on the bed temperature on both catalysts (i.e. platinum or C0304/Cr2O3), decreased

in the order H2 > C2H4 > CO > C3H8 > CR4 (with H2 is the most reactive fuel).. The

conversion efficiency,
of the fuel is defined as the ratio between the initial and exit fuel

concentration to that of the initial concentration of the fuel.

The activity of acatalyst is also subject to variation with time as high combustion

temperatures physically change the catalyst materials. Cullis and Willatt (1983) reported

that prolonged exposure of the precious metals to oxygen at the temperature of 823 K
caused structural changes in palladium catalyst, although this did not occur with

platinum. The different behavior of palladium and platinum catalysts was attributed to the

difference in the ability of these precious metals to adsorb oxygen prior to reaction. They
also found that the catalytic activity for methane oxidation was independent of the size of

palladium particles dispersed on ?-aluminum oxide support. Hicks et al. (1990)

investigated the effect of palladium particle size on its atal,rtic activity in methane

oxidation at 260 to 370 °C. They showed that the catalytic activity of palladium depends

on its particle size. Their results suggested that the methane oxidation activity of the

dispersed PdO was significantly lower than that of the small palladium crystallites. It was

also shown that the catalytic activity of platinum depends on the distribution of the
dispersed metal and crystalline phases.

C2116 Oxidation: The influence of the addition of asecond fuel on the performance of a
given catalyst was investigated by L. D. Pfefferle et al.(1989a, 1989b and 1992). They

found that addition of small amount of ethane to amethane-air mixture allowed easier
22

ignition and flame propagation through aflat plate boundary layer system with platinum

as the catalyst.

ff8 Oxidation: The effect of the addition of propane on the ignition-extinction behavior
of methane in air over platinum foils in stagnation flow at atmospheric pressure has been

studied by Song et ál. (1990) and Balakrishna et al. (1994). Propane is the lightest

straight-chain hydrocarbon whose homogeneous oxidation kinetics appears to be typical

of higher molecular weight hydrocarbons (Glassman, 1980; Hautman et al., 1981). It was

observed that the combined mixture of methane-propane in air with sufficiently high

propane contents ignited at much lower temperature and power input than in the case of a

methane-air mixture. Moreover, coking was observed for fuel rich mixtures in cases in

which methane and propane were the fuel. An early investigation of Wampler et al.

(1976) was concerned with evaluating emission indexes and combustion efficiencies of

propane-air mixtures at inlet temperatures of 700-730 K as afunction of C3H8/air ratios

by volume on a platinum coated monolith. The maximum combustion efficiency

(99.99%) was achieved at C3H8/air = 0.0152 (4 = 0.24) with inlet temperature of 719 K

and the emission indexes, (g pollutant/kg fuel), of NO., and CO were 0.042, and 0.222,

respectively. The differences between platinum itself and very dilute platinum-in-gold

alloys on the dehydrogenation of propane to propene have been investigated by Biloen et

al. (1977). They found that propane dehydrogenation over platinum and over the alloys

occur via the same mechanism which includes the conversion of propyl radicals to ii-

bonded propene, using two adsorption sites located on one platinum atom, one of which

carried a hydrogen atom. Bruno et al. (1983) studied propane conversion at inlet

temperatures of 650 to 800 K, 110 kPa .pressure, velocities of 10 to 40 m/s, and

equivalence ratios of 0.19 to 0.32 with ahoneycomb platinum catalyst with 7-alumina

washcoat. It was found that the propane is mainly converted to CO 2 and H20 at the

surface. They also concluded that under the above conditions; gas-phase reactions tend to

become rapidly more important as temperature and equivalence ratio are increased and as

flow velocity decreased. Also, the surface propane conversion rate was much faster than

the rate of propane diffusion to the surface, resulting in adiffusion-controlled oxidation.


23

The oxidation of propane in air over transitional metal oxide catalysts was investigated

by Prasad et al. (1980) as mentioned earlier. The activities of oxides of Cr, Mn, Mg, Co,

and Fe in single or in -binary mixtures were tested for combustion of C3H8/air in an

adiabatic tubular reactor at various inlet temperatures (563-728 K) and equivalence ratios

(0.07-0.66). They found that the binary mixture of cobalt oxide and chromium oxide were

the .most effective in propane oxidation among the catalysts examined. Moreover, it was

concluded that the complete conversion of propane over C0304/Cr2O3 catalyst started at

an equivalence ratio of 0.196, an inlet velocity of 2.6 m/s, and an inlet temperature of 58Q
K; the NO emission in for these conditions was 0.03 g/kg fuel and was dependent

solely on the value of the exit temperaiure.

CO Oxidation: Carbon monoxide has its share of attention in research in the field of

catalytic combustion not only because of being part of emission products but also as an
important commodity in petrochemical industry. In fact, the oxidation of carbon

monoxide on platinum is one of the oldest known catalytic reactions and one of the most

thoroughly studied, yet it continues to pose difficulties and challenges to both


experimentalists and theoreticians, mostly due to the frequent observations of hysteresis,

associated with steady-state multiplicity, and self-sustained oscillations in the reaction. In

fact, there is no generally accepted theoretical explanation at hand, even though alarge

number of studies have addressed the subject. These studies were summarized by Razon

and Schmitz (1986). It is suggested that such hysteresis is due to the interaction between
reaction processes on the surface and transport processes between the surface and the gas

phase producing unstable phenomena (i.e., steady-state multiplicity and oscillations). Gas

phase impurities can produce spurious behavior. However, gas phase impurities were

dismissed as being the sole cause of this behavior after great care had been taken to
eliminate impurities from the system. This leads to the conclusion that the unstable

phenomenon must also be related to processes that occur on the surface (catalyst

morphology), of which the kinetic mechanism is the most important factor.

As early as 1959, Khitrin and Solovyeva investigated the role of surface reactions

in the combustion of carbon monoxide to obtain kinetic data inside narrow platinum and
24

copper tubes at temperatures ranging from 500 to 1400 °C. The global rate expression
obtained for small carbon monoxide concentrations was independent of oxygen

concentration. The mechanism of dry carbon monoxide oxidation over precious metals

was reported by Dwyer (1972). Employing platinum gauze as a catalyst and. using C

tracer measurements, the stoichiometric number of the carbon monoxide oxidation

reaction was measured. It was concluded that the reaction scheme and most probable
mechanism over platinum surface is: -

02 at the surface) (2.2)

CO + °(adsorbed at the surface) > CO 2 (2.3)

The investigation of carbon monoxide oxidation from synthetic gas mixtures on a


pelleted platinum-alumina catalyst at atmospheric pressure and in atemperature range of

200-370 °C was studied by Voltz et al. (1973) in order to establish arate expression fpi

CO (with 10% H20 on a wet basis) oxidation. A complex kinetic equation was

formulated with rate constants and activation energies. They reported that the rates of

oxidation were increased with increasing oxygen concentration and were inhibited by

carbon monoxide, propylene, and nitric oxide. Aging of pelleted-alumina and platinum

monolithic catalysts was also addressed in their investigation. However, this study was

intended for automotive emission control systems as was the work of Kuo et al. (1980). A

much simpler rate expression form was proposed by the later study of humid carbon

monoxide on platinum-alumina pellets. It was reported by Kuo et al. that lowering the

catalyst density or heat capacity by 25% enabled the inlet gas to warm up alarger portion

of the catalyst, and therefore increased the CO and HC conversion during warm-up.

Boehman et al. (1992) obtained an overall reaction rate for carbon monoxide oxidation

for temperatures from 180 to 300 °C and CO concentrations from 0.1 to 4.0%. The rate

expression was based on the dual-site Langmuir-Hinshelwood mechanism with empirical

modifications to account for the effect of high CeO 2 content over platinum. Langmuir-

Hinsheiwood mechanism is based on the assumption that both reactants are adsorbed
25

without dissociation at different free sites on the catalyst surface. This is then followed by

the actual surface reaction between neighboring chemisorbed molecules' to give the

product adsorbed on the surface. In the final step the product is desorbed away from the

surface. Dekker et al. (1994) studied the oxidation of dry car12on monoxide over an

alumina supported copper-chromium oxide catalysts for automotive pollution control

using the step response method. It was found that the catalyst goes through an
oxidation/reduction cycle when the gas feed composition is changed between oxidizing

and reducing. In the oxidation cycle, the catalyst temperature was raised to 773 K in an

oxygen stream for half an hour followed by cooling to room temperature. While in the

reducing cycle, ahelium gas stream was used. Also, the activity of the catalyst is higher

than the steady-state activity after the gas composition has changed from reducing to

oxidizing. It was also shown that 27.0% of the surface was covered with CO and CO 2 in a

reducing reaction mixture and 18.0% in an oxidizing reaction mixture. Other

investigations dealt with different issues indirectly involving carbon monoxide such as

the catalytic combustion of alow heating value fuel (biomass). Magnus and Jaras, (1997)

noticed-an improved oxidation of methane over LaMAl 11 O19 (La is acatalyst promoter;

M = Cu, Mn, Ni, Co, Al) catalysts by the presence of CO (with 10-12% H2O) and H2.

112 Oxidation: Catalytic oxidation of hydrogen has also been the subject of many

investigations. Hydrogen has many attractive features. Apart from being an excellent fuel

itself,it is needed increasingly for the upgrading of fossil fuels, the production of

alternative fuels, such as methanol, and awide range of other products. It is used in fuel

cells and is an important raw material in the petrochemical industry. Boudart et al. (1977)

studied the effect of pressure on reactions between hydrogen and oxygen on platinum
foils supported on silica gel in mixtures with an excess of oxygen. They found that

catalytic' activity was much less at high pressure than that at low pressure. Mori et al.

(1977) showed that the oxidation of hydrogen (2-7 mol %) at high pressure can be

improved by employing high temperatures up to 1100 K, so as not to produce nitrogen

oxide for hydrogen-air. Hanson and Boudart (1978) investigated the effect of platinum

dispersion and particle size on the reaction between H2 and 02 over Pt/SiO 2 catalysts at
26

temperatures between 273 and 373 K. Their results show that mixtures with an excess of

hydrogen reveal that the catalyst with the highest dispersion was seven times more active

at 273 K than was the catalyst of lowest dispersion. Whereas the reaction was first order

in oxygen and zero order in hydrogen. In the mixtures with an excess of oxygen, the

catalytic activity was independent of particle size, and the reaction was first order in

hydrogen and zero order in oxygen. Schefer et al. (1980) studied interactions between the

gas-phase combustion of hydrogen-air mixtures and surface reactions over aplatinum flat

plate in high-temperature combustion. The results obtained for arange of equivalence

ratios from. 0.05 to 0.30 and platesurface temperatures from 470 to 1300 K showed a
significant surface heat release for all mixtures at plate temperatures as low as 470 K. The

investigation of Schefer et al. was continued by Brown et al. (1983). It was devoted to

developing acatalytic combustion system in which the surface kinetics of various fuel-air

mixtures could be studied. Moreover, the system was intended to enable the study of the

high temperature surface oxidation rate for lean H2/air mixtures on aplatinum catalyst

over arange of surface temperatures from 450 to 1070 K and equivalence ratios from

0.05 to 0.20. They found that in the laminar boundary layer of aheated catalytic plate the

surface reaction rapidly became diffusion limited downstream of the plate leading edge.

This enabled them to obtain the surface reaction kinetics under such conditions where

surface reaction rates were significantly influenced by transport effects. The production

of hydrogen from direct oxidation of methane to synthesis gas over aNi/Al2O3 catalyst at

low temperatures was the subject of research by Choudhary et al. (1993, 1998). Their

results reveal that the performance of the reduced catalyst (reduction by H2 at 500 °C)

was better. It was also concluded that the partial oxidation of methane over Ni/A1203

under nonequilibrium conditions using high space velocity of 138.8 (its) results in a

much higher selectivity and productivity of CO and H2 than that obtained at the

equilibrium, with achange in the reaction path.


27

2.3 Theoretical and Computational Studies

Many mathematical models of catalytic reactors have been recently developed in order to

study their behavior and improve their performance. The performance of a catalytic

combustor is dependent upon operational conditions of the combustor i.e. the flow,

temperature, and composition of the mixture, the combustor design, and the properties of

the catalyst used. To find the optimal combinations of these factors by building and

testing actual catalytic combustors (reactors) would require avery elaborate ,and costly

experimental program. A desirable alternative is to develop and use apredictive model of

the combustor so that various combinations can be screened to determine which should

be built and tested.

The mechanisms of fuel conversion in catalytic combustion are very complex

phenomena. These to be' accounted for in catalytic combustion modeling include

heterogeneous and homogeneous chemical reactions, energy, mass and momentum

transfer by convection and diffusion in the gas phase as well as at the gas-solid interface.

These phenomena are strongly interactive and nonlinear because of the thermal effects

associated with the chemical reactions. Considering the strong nonlinearity associated

with temperature effects, the numerical solution of the governing equations is quite

difficult and requires the making of anumber of approximations. However, there are

successful attempts to model the transport, phenomena and chemical reaction

mathematically (Corma et al., 1998). The subject of mathematical modeling of catalytic


combustors is well reviewed by different authors. The most recent review is published by

Groppi et al (1999). They concluded that mathematical models represent apowerful tool

for the design, analysis and operation of catalytic combustors for gas turbines. The

complex physical and chemical phenomena affecting the behavior of a catalytic

combustor are numerous and the identification of reasonable approximation levels is

crucial. Computational studies reported in the literature present theories and analyses

which attempt to mathematically investigate the following:'


28

2.3.1 Bulk Diffusion

The diffusion of premixed hydrogen-air mixture over ahot platinum and non-catalytic

(quartz) surface (flat plate) has been, modeled by Markatou et al. (1991). They have

shown that the production of OH radicals during the surface reaction enhanced the

combustion of lean mixtures. Also, the thermal and mass boundary layers of hydroxyl

radical concentration became thinner for hydrogen combustion at = 0.10 and aplate

temperature of 1170 K. Markatou et al. attributed the thinner thermal and mass boundary

layers in the case of catalytic walls to the depletion of the fuel due to the surface reaction.

Their findings agree with the mass boundary layer thicknesses observed experimentally

for catalytic combustion of CH4IC2H4/air and H2 combustion under other conditions.

Goraiski and Schmidt (1999) have modeled the effect of catalyst pore size, fuel

concentrations, temperatures and pressures on the tubular plug-flow reactor exhaust

compositions in the high-temperature (T = 1473 K) combustion of methane on platinum.

They pointed out that the homogeneous chemical reaction is significantly inhibited by the

heterogeneous reactions. This inhibition is aresult of the adsorption of radical species

from the gas phase at the surface of the catalysts, which prevents initiation of gas-phase

free-radical chain reactions. The resuls of Goraiski and Schmidt (1999) showed that at

high temperatures and pressures the adsorption of the radical species 0, H, and OH on

the surface ihhibits the homogeneous chemical reaction enough that negligible amounts

of NO and CO are formed.

2.3.2 Intraparticle Diffusion

Mathematical models have also been developed to investigate the effects of an

intraparticle difftisional resistance (diffusion within the catalyst pores) on the ignition

performances of acatalyst. Trimm and Lam (1980) numerically studied the oxidation of

methane over platinum supported on porous and non-porous alumina fiber. The kinetics

of the reactions have been evaluated at temperatures above and below 815 K, where a

change in the apparent activation energy was found to occur. It was pointed out that pore
29

diffusion should be significant between 950 and 1000 K from their calculation of the

effectiveness factors (defined as the ratio of the rate of reaction when diffusion is

important at high temperatures to the rate of reaction when constant concentration

prevails throughout the catalyst at low temperatures). Nakhjavan et al. (1995) have

investigated the effect of catalyst surface area on monolith' combustor performances by

using a1-D model which accounts for heterogeneous and homogeneous reactions and for

axial heat conduction in the solid phase. Their analysis was applied to laminar flow

conditions and the results indicated that the porosity of the bed has an effect on the

ignition of the catalytic combustor. It was shown that larger catalyst surface areas will

benefit catalytic ignition especially during start-up. The results of Nakhjavan et al. also

show that the effect of porosity is more important in the transient state than in the steady

state. Moreover, the effect of porosity in their simulation was found to depend on the type

of fuel used for the combustion (i.e. slow oxidizing fuel such as methane, fast oxidizing

fuel such as propane',and extremely fast oxidizing fuel such as hydrogen). Groppi et al.

(1999) investigated the role of surface area on steady-state ignition performance of the

segmented catalyst section of ahybrid combustor. They developed a 1-D mathematical

model of a single monolith channel fueled by natural gas. The reaction-diffusion

phenomena within the catalyst poreswere accounted for by calculation of the catalyst

effectiveness factor based on the Thiele modulus .(defined as the ratio between the

intrinsic reaction. rate to the effective diffusivity of the limiting reactant or product), using

the analytical solution for first-order kinetics and for isothermal catalysts. Their

simulation results showed that a minimum level of catalyst specific surface area is

required to achieve best performance of th6 combustor.

2.3.3 Coupled Heterogeneous and Homogeneous Reactions

A number of theoretical and experimental studies have been conducted to acquirea better

understanding of the interaction between homogeneous and heterogeneous reactions. The

significant effect of this interaction demonstrates the importance of considering


30

homogeneous-heterogeneous reactions on various configurations such as the flat plate

boundary layer, the stagnation-point flow, and monolithic reactor.

Schefer and co-workers (1980, 1982, and 1983) conducted fundamental studies to

understand the role of catalytic activity at the boundary layer. They were among the first

to investigate the relation between catalyzed heterogeneous reactions at the wall and

homogeneous reactions in the adjacent boundary layer. The results of their simulation
showed that aplatinum wall acted as asource of free radicals only near the plate leading

edge. Further downstream, it appeared to be asink for free radicals, unlike quartz which

does not promotes fecombination of OH radicals, at the given conditions of 1 = 0.1 at

1170 K. The roles of heat and mass transfer and their effect on the combustion process

were also investigated. Their results indicate the existence of an initial region near the

plate leading edge in which the radical concentration increases with little associated heat

release, and adownstream region where heat release due to gas-phase combustion results

in a significant increase in thermal boundary layer thickness. The main difference

between Markatuo et al. (1991) study and of Cattolica and Schefer (1982) is that the later

did not observe the thinner thermal and mass boundary layers due to the surface

combustion of hydrogen. Treviflo (1981) and Treviflo and Peters (1 985) also used a

boundary layer model to establish criteria for gas-phase ignition in catalytic systems with

external heat loss. They assumed amass-transfer-limited surface fuel oxidation and one-

step homogeneous reaction. Markatou et al. (1993) used detailed gas-phase reaction in

their boundary layer model for the combustion of methane-air mixtures over heated

catalytic and non-catalytic surfaces maintained at constant temperature. They found

computationally that 0, H, or OH radicals can significantly accelerate the induction

period for ignition of light hydrocarbon mixtures in air. Their results illustrate the

sensitivity of the predicted methane flame propagation on the surface boundary

conditions used, which is more pronounced than for ahydrogen flame studied earlier..

The interaction between the heterogeneous and homogeneous reactions was also

investigated in different flow configurations (stagnation point flow). A computational

study of.the combustion of mixtures of hydrogen and air (4 from 0.15 to 0.3) impinging

on aplatinum plate at an elevated temperature of 1173 K was presented by Ikeda et al.


31

(1993). It was shown that, at relatively low surface temperatures and low concentrations

of hydi'ogen, heterogeneous reactions dominate while at higher temperatures and higher

concentrations homogeneous reactions tend to do so. However, their model was not

validated experimentally. Vlachos and co-workers (1993, 1995, and 1996) computed the

ignition and extinction of the homogeneous combustion of lean H2 in air from room

temperature up to 1500 K with stagnation point flow and detailed chemistry numerically.

They concluded from their studies that H atoms are, the most important in shifting both

ignition and extinction points away from the surface, especially near a heated wall at

sufficiently large equivalence ratios (about 12% H2). The HO 2 radical is the next most

important in ignition, but the least important in extinction. Moreover, the presence of

H20 in the initial mixture results in aslightly enhanced combustion of 112. Machos et al.

also pointed out that ignition and extinction can be influenced by heat transfer, especially
at low pressures. At high pressures flame instabilities are mainly driven by reaction heat

generation and are strongly affected by the enhanced third body efficiency of 1120 in the

reaction H + 02 + M --> HO2 + M. A more rigorous multi-component model of the

catalytic ignition of CH4, CO, and 112 oxidation on platinum and palladium at

atmospheric pressure was studied by Deutschmann et al. (1996) for two simple

configurations. The simulation includes detailed reaction mechanisms for the gas-phase

and the surface. A quantitative agreement between measured and, calculated ignition

temperatures was obtained. It has been shown that the ignition temperature of CH4

decreases with increasing CH4/0 2 ratio up to an equivalence ratio of 0.9.

For monolithic tubes geometry in high temperature catalytic processes, the

behavior of reactors in which homogeneous and heterogeneous reactions occur

simultaneously is modeled and investigated by Harrison and Ernst (1978). A steady-state

model of heat and mass transfer for the combustion of four mole percent dry carbon

monoxide in an adiabatic laminar tubular flow reactor was considered. However, axial

heat conduction and diffusion were not included. Three reaction models were

investigated: homogeneous phase reactions only, heterogeneous reactions only, and

combined heterogeneous and homogeneous reactions. Comparison of results from these

three models demonstrates that it is necessary, to include both homogeneous and


32

heterogeneous reactions when modeling adiabatic high temperature "catalytic converters"

because at temperatures between '650 and 1150 K homogeneous reactions ,begin to


contribute significantly to the intermediate pool of radicals. However, they concluded

that for design purposes, for sufficiently high inlet temperatures (i.e. near the upper end

of this range, which is close to 1150K), acatalyst may be necessary only near the reactor

inlet to light-off the homogeneous reaction. Bensalem and Ernst (1982) modified the

previous model (l'aminar flow) and applied it to atwo-chamber, reactor (laminar and

turbulent flow) of the same length in order to improve its performance at high

temperature. The first chamber was asmall diameter,


channel with acatalytic wall which

was' mainly controlled by amass transfer surface reation. The chamber was not coated

with catalyst and had alarger diameter. In this terminal portion of the reactor, the bulk

gas temperature was high enough to support the homogeneous combustion of the fuel.

Bensalem and Ernst showed that this particular deign minimizes considerably the

pressure drop without altering the overall conversion at relatively ahigh temperature and

turbulent flow. Such atwo-chamber reactor might be used when there is an interest in

minimizing the pressure-drop (e.g. combustors, stack clean-up reactors). Their results

also confirmed the earlier results of Harrison and Ernst indicating that both the

homogeneous and heterogeneous reactions are important at the intermediate range of

'temperatures from about 650 to 1100 K. T'ien (1981) developed atransient model to

investigate the behavior of catalytic monolith combustors including both heterogeneous

and homogeneous reactions in fuel-lean operation. He used aone-dimensional model that

allowed for temperature transients in the thin monolith wall and aquasi-steady-state gas

thermal balance. T'ien's results also indicate that higher combustor temperatures and

adiabatic flame temperatures shorten the response time to a change in the upstream
conditions. Alin et al. (1986) developed a one-dimensional transient model similar to

T'ien's model, which accounted for the effect of thermal conductivity in the wall. They

confirmed that cold start-up was astrong function of the monolith's thermal inertia, its

material and configuration. Metallic monoliths were found to respond much more quickly

to changes in operating conditions when compared with ceramic monoliths. Higher inlet

temperatures and operating pressures, and lower gas velocities were shown to result in
33

faster response times. Performance of reactors of square channel geometry was better

than that of triangular or sinusoidal geometry because of the better heat and mass
characteristics of the square channel geometry.

2.3.4 Transport Mechanisms

Chen et al. (1988) developed athree-dimensional model to analyze the transient response

of an axisymmetric ceramic monolith car converter during its warm-up and engine

misfiring. Their model simulates the thermal and conversioh characteristics of anon-

adiabatic monolithic converter operating under flow maldistribution conditions. Their

model accounts for convective heat and mass transport, gas-solid heat and mass transfer,

heat conduction, chemical reactions and their attendant heat release, and heat loss to the

surroundings. The results of their simulations indicated that variation in the exhaust flow

rate had aminor effect on conversion efficiency during warm-up. However, the noble

metal loading and the flow distribution of the exhaust gas were shown to have astrong

influence on the solid temperature profile during the converter warm-up. Flow

maldistribution and radiant heat loss were identified as the major sources for the

temperature gradient. The predicted temperature profiles provided abasis for the analysis
of thermal stresses and fatigue in the monolith converter assembly.

Groppi et al. (1995, 1997, 1999) have reported that 1-D models for avarious

channel geometry (triangular) lead to predictions of gas exit temperatures for asingle
segmented monolith in CH4-fueled catalytic combustors that are in fair agreement with

those of more rigorous 2-D and 3-D models. This conclusion is obtained under the

assumption that light-off occurs immediately downstream of the catalyst inlet.

Bond et al. (1996) developed a numerical model of the catalytic combustion

process based on CHEMKIN code (Coltrin et al., 1991) for detailed gas-phase and

surface kinetics. The effect of equivalence ratio (ranging from 0.11 to 0.44) and inlet

velocity (from 5.5 to 10.0 m/s) at atmospheric pressure were considered. The predictions

of the numerical model of combustion of natural-gas over platinum supported on ceramic


34

honeycomb monoliths were in fair agreement with


' the experimental data at temperatures

below 900 K. However, at higher temperatures the predicted fuel conversion was two

orders lower than the measured data. NO emissions were shown below 1ppm and CO

was observed at the ppm level.

The application of catalysts in radiant burners was considered by Rumminger et

al. (1999). Simulations were performed to investigate the effect of platinum catalyst

loading, location of the catalyst layer, and its thickness on the performance of burners

with single and bi-layered porous media. It was shown that only athin layer of catalyst

placed at the downstream edge of the porous medium was required for efficiency

improvements. The model predictions were compared with experimental data obtained on

catalytic radiant burners.

The effects of operating conditions on combustion efficiency and pollutant

formation were reported by Moallemi et al. (1999). The numerical and experimental

investigations were carried out using palladium and platinum based catalytic monoliths

for methane combustion (for heating purposes). The model for a well-stirred reactor

included detailed homogeneous/heterogeneous chemical kinetics for methane-air

mixtures of equivalence ratios ranged from 0.3 to 06 and space'velocities between 400

and 1209 (1/mm). Dupont et al. (2000) extended the above model to investigate the effect

of the catalyst (Pt and, Pd) loading, flow rates ranging from 40 I/min to 94 1/mm, and

monolith length. It was found that apalladium catalyst provided alarger range of stable

operation than platinum ones for equivalence ratio ranging from 0.44 to 0.80. Their

results showed that the pollutant emissions (i.e., CO, NO,,, and UHC) were "near zero"


for the operating conditions employed in their investigation. The catalyst role was proved

not only necessary to enable the ignition of fuel mixtures below flammability limits, but

also to ensure the complete oxidation of the fuel' to CO 2 via surface reactions in the

steady state. Also areduction of 70% in the original length (of 50 mm) of the monolith
was found possible.

Accordingly, in view of the key role that mathematical modeling can play in

future developments of catalytic combustor technology, joint efforts from experimental


and theoretical work groups devoted to model validation are strongly recommended.
35

In general, the above mentioned models were developed for specific applications

and cannot be applied directly to other conditions. In addition some of these models were

not validated experimentally. Recently, amodel of atransient plug-flow reactor which

includes bed heat conduc'tion, radiative heat transfer, axial diffusion, and convective heat

transfer between the solid and gas in addition to the homogeneous and heterogeneous

reactions was pr
oposed by Younis and Wierzba (2002, 2004) and is the subject of the

present thesis. The effect of operating conditions such as inlet temperatures, approach

velocities, and equivalence ratios on methane conversion efficiency were investigated:

2.4 Summary

Catalytic combustion is employed in a wide range of applications to utilize various


gaseous fuels efficiently a:nd suppress exhaust emissions without major increase in the

cost. A substantial research effort has resulted in some interesting options for materials
utilized in catalytic combustion. Especially, the incorporation of non-noble metal

catalysts on high thermal resistance alumina support seems to provide an opportunity to

circumvent loss of surface area and catalytic activity even at high temperatures. Several

single and complex transition metal oxides such as cobalt and chromium oxides have

ben shown to have high oxidation activity, approaching the noble metals. However, it

appears to be difficult to combine the two toughest demands for combustion catalysts,

namely, sufficient low temperature activity and high-temperature stability. Even though

positive results have been achieved, attention should be focused on metal oxides catalysts

because of the advantages they offer such as lower cost and potentially higher maximum

operating temperatures. The catalytic combustion within packed bed reactors is not we ll

pursued specially for non-noble metal catalysts, most probably due to its high pressure

drop even though, packed bed reactors have a wide application in the .petrochemical

industry. Furthermore, there is alack of kinetic data for non-noble, metal oxide catalyst

which is important for modeling purposes. Therefore, modeling and experimental

research are necessary to generate realistic models to improve the reactor design and
36

operation. Moreover, the combined heterogeneous and homogeneous reactions are

important, especially for high temperature combustion and in some cases for ignition.'

The present study attempts to develop a mathematical model for the catalytic

combustion in apacked bed reactor and thus to obtain kinetic data for abinary mixture of

cobalt and chromium oxides catalyst. The developed model incorporating the kinetic data

obtained during this investigation should allow more time and cost efficient parametric

study of such catalyst.


37

CHAPTER 3

ANALYSIS AND MATHEMATICAL MODELING

3.1 Introduction

The development of the present model followed the philosophy of constructing a

catalytic packed bed reactor mathematical model which includes only the essential

parameters of the reactor. These parameters were determined by beginning with asimple

model that contained only the major transport processes and physical and chemical

parameters and then comparing the performance of the model with values measured in an

experimental catalytic reactor. When any significant discrepancy arose, the model was

refined by including more details (e.g., the model initially assumed constant gas

properties, then the model was refined for these properties to be afunction of the local

temperature), but without including any parameters that are arbitrarily adjusted to

improve the model predictions.

Accordingly, in the present investigation, amathematical model was proposed for


the oxidation of lean fuel-air mixtures within inert and catalytic packed bed reactors. The

transient model developed includes heat and mass transfer (i.e. axial heat conduction in

the gas and solid phases, radiative heat transfer, axial diffusion, and convective heat
transfer between the solid and gas) as well as both homogeneous and heterogeneous

reactions. The developed model is general in applicability and can be applied to any

gaseous fuel and catalyst but the calculations are mainly made for methane with platinum
as the catalyst. The present model is used to obtain kinetic data for other fuels over non-

noble metal catalysts such as the binary mixtures of cobalt and chromium oxides within

the packed bed reactor.


38

Thus, in this chapter, a detail description of the steps and efforts employed in

developing the mathematical model including homogeneous reactions will be addressed.

In order to validate the mathematical model, an experimental program was conducted in


parallel. The mathematical analysis starts with the physical description of the system

under consideration.

3.2 Description of the Physical Model.

A premixed, preheated homogenous fuel-air mixture enters into acylindrical packed bed

reactor (catalytic or non-catalytic), initially at auniform temperature. The temperature of

the packed pellets and the fuel-air mixture changes with time and distance due to heat

transfer and chemical reaction processes. The schematic diagram of the physical model

under consideration is shown in Fig. 3.1.

Fig. 3.1 Schematic diagram of the physical model.


39

The transport processes taking place in the packed bed reactor are very complex, as they

involve several modes of heat transfer, as well as homogenous and heterogeneous

chemical reactions. When the preheated fuel-air mixture flows through the reactor bed

the heat transfer processes involve convective heat transfer between the gaseous mixture

and pellets, axial heat conduction in the pellets as well as heat radiation. Radiative heat

transfer in the porous bed is an extremely complex process. It can take place between the

solid pellets themselves, between the solid pellets and gases, as well as between the

pellets, gases and the walls. Although, radiative heat transfer is not very significant at low

temperatures, it can b'ecome a dominant mechanism of heat transfer at elevated

temperatures. The most significant mode of radiation is the radiation in the solid pellets.

Therefore, the radiative heat transfer in only the solid pellets is accounted for in the

model prediction. Since the gas temperature is intimately coupled to the solid via internal

convection, it also exhibits transient behavior until steady state is reached. Moreover,

heat conduction takes place in the solid, in both directions from the location in the packed

bed where chemical reactions commence, as that location will be heated first. The heat

transfer modes mentioned above take place simultaneously in addition to the

homogeneous and heterogeneous chemical reactions. Some of these heat transfer modes

such as radial heat conduction and radiative transfer in the gas need not to be accounted

for in the present one-dimensional analysis, due to the numerical. complexity involved in

solving the resulting nonlinear partial differential equations.

3.3 Description of the Mathematical problem

3.3.1 Governing Equations

Governing equations for ageneral packed bed reactor are presented here for the oxidation

of fuel-air mixtures within atransient system. In the system described, the following
assumptions were made:

1. The velocity of the flow is uniform, i.e., plug flow.

2. The reactor is operated adiabatically.


40

3. The gas and solid are not in local thermal equilibrium, i.e., separate energy
equations are employed for these two phases.

4. Gaseous radiation is negligible compared to the pellet radiation.

5. Radiative heat transfer in the packed bed was modeled using the concept of so-

called effective thermal conductivity (Eq. 3.4) of the pellets which is considered

to be afunction of their temperature.

6. The density, the specific heat and the thermal conductivity of the gases are

considered to be variable and are evaluated for the instantaneous local


temperature and mixture composition.

7. The thermal properties such as the density, thermal conductivity, specific heat,

and porosity of the pellets are uniform and constant.

It should be noted that the thermal conductivity of the solid (pellets) is not the same as

the effective conductivity of the material. The effective conductivity constitutes the

pellet's solid thermal conductivity and flow conditions (i.e., the porosity of the bed) as
well as the radiation effect.

The resulting laminar, axisymmetric, unsteady, one-dimensional (plug flow) conservation

equations of mass, energy and species for the transport phenomena can be written as:

Continuity

7(P g s) + ax (P g "i) 0 (3.1)

where, p is the density of gaseous mixture, tis the time, xis the axial distance along the

reactor, uis the axial velocity, 6 is the porosity, and subscript 'g' refers to gas.

Energy equation for the gaseous phase:


41

a a a aT
—(pc sT)+—(p usc T)
at g pg g ax g pg g 8x g ax
(3.2)
-(1--s)h -1'3- ethHkAiflf

In the above equation, T is the local temperature, kg is the gas thermal conductivity and

c, is the constant pressure specific heat of the mixture, k is the volumetric heat transfer

h
coefficient per unit volume, Hk is the specific enthalpy of species k, (Dk is the molar

production rate per unit volume of species k due to homogenous reaction in the pore,

MJ4 is the molecular weight of species k, where subscript 's' refers to solid and, 'k'

refers to species k (i.e., the fuel or the oxidant). The gas property data are listed in

Appendix A.

The first term of the left-hand side of equation (3.2) represents the storage of

energy in the gas per unit volume. The second term accounts for the energy advected by

the gas per unit volume. The first term of the right-hand side of the equation (3.2)

accounts for heat conduction in the gas phase per unit volume. The second term

represents internal gas/solid convection per unit volume. The third term represents the

amount of energy released in the as per unit volume due to homogeneous reactions.

Energy equation for solid phase (pellets):

a = H15)keff) + (1—e)h(2--I)
ax ax
(3.3)
+ (1—e)th HkMWk acat

where cs is the specific heat of the solid, Wk is the molar consumption 'rate per unit area

of species kdue to catalytic reaction (to be discussed in the next section), and acat is the

catalyst specific surface area. -


42

The left-hand side of equation (3.3) represents the storage of energy in the solid per unit

volume. The right-hand side first term accounts for heat conduction per unit volume. The

second terñi represents internal heat convection between the gas and pellets per unit

volume. The third term stands for the amount of energy released. at the surface per unit
volume due to heterogeneous reaction.

In equation (3 .3), 'kW is the effective thermal conductivity of the pellets calculated using

aRosseland diffusion approximation as follows (Modest, 1993):

keff = k + 16a7 or keff = f(T) (3.4)


38

where o is the Stefan-Boltzmann constant, 5.67 x1 W/(m2.K4), and 0is the extinction

coefficient.

The volumetric heat transfer coefficient h equation (3.2 and 3.3) between the gaseous

mixture and pellets is calculated as:

hV (3.5)

where S, and V, are the surface area and volume of apellet correspondingly, h is the

convective heat transfer coefficient calculated as follows (Wakao and Kaguei, 1982) for

gaseous laminar flow:

hd 0. 6
Nu ,
= - 2.0 + 1.1Pr Re (3.6).
kg

In the above equation, Nu, is the averaged Nusselt number, d, is the pellet diameter, Pr,

is the Prandtl number, and Re p ,is the Reynolds number.


43

where

/1 Cpg
Pr = (3.7)
kg

where p is the viscosity of the mixture.

and

ud
Re = —a (3.8)
Vg

where v9 is the kinematic viscosity.

Chemical species balance:

o a a
.(pg eY)
j + — Co euYk) = ,. — )
—(pg .- D l + a'd)MWk (3.9)
ax g OX Ox

where Yk is the mass fraction of species k, Dk is the mixture diffusion coefficient of


species k.

In equation (3.9) the first term of the right-hand, side represents the molecular diffusion

of the species kper unit volume in the mixture. The second term stands for the generation

of the species kdue to homogeneous chemical reaction per unit volume. The first term of

the left-hand side of the equation (3.9) represents the rate of gain of mass of species kper

unit volume in the gas mixture. The second terms accounts for the mass flow of species k
out due to advection by bulk flow per unit volume.

The gaseous mixture is treated as an ideal gas at constant atmospheric' pressure.

Therefore, the mixture density is calculated from:

pMW
1
09 (3.10)
R Tg
44

In equation (3. 10), p is the pressure, MW is the mean molecular weight of the mixture

and R is the universal gas constant.

3.3.2 Kinetic Modeling

• Gas Phase (Homogeneous Reactions)Kinetic Modeling

The energy released due to gas-phase reactions is calculated by employing an

irreversible one-step overall homogeneous reaction of the Arrhenius type as

follows:

h = Ag Cm C exp [j Eg (3.11)

where Ag is the pre-exponential factor for homogeneous reaction. Cf and C0 are

molar concentrations of the fuel and oxygen respectively. Eg is the activation

energy for, the gas-phase reaction. m, and n are the exponents of the molar

concentration of the fuel and oxygen.

• Catalytic Surface (Heterogeneous Reactions) Kinetic Modeling

The surface reaction rates over the catalytic pellets are modeled as an irreversible,

one-step reaction rates using the empirical expressions proposed by Soiig et al.

(1991) and Markatou et al. (1993):

A Cf C' 5
ex[
ES
R J (3.12)
45

where A is apre-exponential factor for the heterogeneous reaction, Cf and CO

are molar concentrations of the fuel and oxygen respectively, E8 is the activation

energy for the heterogeneous reaction.

This form of the rate epression is consistent with all experimental results cited

under conditions relevant to this study (high oxygen concentration). This would be

consistent with aLangmuir-Hinshelwood type reaction mechanism assuming low

overall surface coverage and dissociative adsorption of oxygen.

The molar concentration of the species, Ck, is related to their mass fraction, Yk,
by:

Ck =P g Yk /MWk (3.13)

3.3.3 Initial conditions:

The gaseous mixture is assumed initially to be at uniform initial temperature and mixture

composition:

At t=O and O≤x≤L

U= Ujn, Tg= Tin, Yi = 0 (3.14)

3.3.4 Boundary Conditions:

For the system described in section (3.2), the following boundary conditions are

considered:

atx=0 uu; Tg Ts Tin; Yk.= Ykin (3.15)

at x= L au =0; -=0 g=
(3.16)
ax' 5x ax
46

3.4 Numerical Model

3.4.1 Method of Solution

The SIMPLE algorithm for solving the equations is discussed in detail by Patankar

(1980). The methodology used to solve equations (3.1, 3.2, 3.3, and 3.9) with the

associated boundary conditions is based on the implicit, control volume, finite-difference

technique. The unsteady finite-difference form of the energy equations is obtained for

each node location, by integrating equations (3.2) and (3.3) over space in the x-direction

and over time. These algebraic equations are then solved using aline-by-line iterative

method. Due to the nonlinearity, the gas and the solid temperatures are underrelaxed. The

underrelaxation of temperatures helps to prevent the solution from diverging. .The time

dependence is handled by stepping in time using fully Euler method and retaining a

converged solution at each time step. The grid used in the finite-difference scheme

establishes the positions of the nodes; the control volume boundaries are shown in Pig.

3.2, where the dashed vertical lines are the control volume boundaries. The spacing

between the boundaries is the control volume thickness. The solid dotted lines denote the

node position. There are atotal of M nodes in the porous bed of length L. The number of
nodes used to generate the results is 101 nodes in a50'mm bed length.

3.4.2 'Computer Code Development

A 'step-by-step approach was pursued in developing a computer code to simulate the


flow; heat and mass transfer in the packed bed reactor. Each stage of the program was

tested carefully and evaluated. A summary of the steps followed in the process of
.developing the computer code is listed below:

1. The energy equations for the fluid and the solid were solved. The sensitivity of the

program with respect to volumetric heat transfer coefficient was tested by setting the

initial temperature of the fluid and solid at different values. The solution converged
47

to the same temperature for the fluid and solid after acertain period of time. The

period to reach equilibrium was dependent on the velocity.

2. All the above steps were applied first for constant thermophysical properties of the

fluid and the solid. Then variable .propèrties were implemented in the computer -
program.
(

3. The species conservation equation was added with a global irreversible one-step
reaction (Arrhneius type).

4. The global one-step heterogeneous reaction was added to the solid energy equation

and was implemented in the numerical code.

A flow chart of the numerical code is shown in Fig. 3.3.

Fig. 3.2 Control volumes and mesh generation.


48

(
Start
Input data
(Tj, uj, and )

I Calculate
1 thermophysical
properties of the gas

Set boundary and


initial conditions
S.

Solve energy
equations for gas
and solid phases
I

I
Solve species
equation

Update the
solution

If the time
> max. time

Yes

Output

Fig. 3.3 Flow chart of the numerical model.


49

3.5 Preliminary Results of Numerical Simulation of Methane

The conversion of afuel in lean fuel-air mixtures depends on various parameters, such as

inlet temperature, equivalence ratio, 4) (defined as the actual fuel/air ratio divided by the

stoichiometric fuel/air ratio), approach velocity, and u(defined as the average velocity of

the flow in the conduit without packed pellets). The effects of these parameters on the

extent of conversion of afuel were investigated numerically using the above model for
the range of operating conditions summarized in Table 3.1.

Table 3.1 Values of operating parameters employed in the numerical calculations.

Length of reactor (mm) 50.0; 75.0; 100.0

Type of packed pellets 0.5% Pt on y-Al2O3; y-A1203

Apparent average density of the packed bed, Ps (kg/m3) 100.0

Thermal conductivity of the packed bed, k5(W/m.K) 1.0-4.0

Specific heat of the packed bed, CPs (J/kg.K) 837.0

Porosity, s 0.2 - 0.9

Inlet temperature, Tin (K) 293 - 1300

Approach veloèity, u(m/s) 1.0 - 3.0 -

Equivalence ratio, 4) 0.05 - 0.50

Fuel CR4;C2H6;C3118;CO; H2

The fuel conversion was calculated as:

% fuel cony = 4fue1 100 = fuel, fuel out 100


(3.17)
-

fiiel 1 fuel,

Moreover, the effects of bed porosity, solid thermal conductivity, bed length, and fuel

type were also investigated.


50

Methane was employed as the primary fuel in the present study. The influences of the

numerous parameters identified above on the conversion rate of the methane were

studied. Three different types of calculations were conducted with methane as afuel. In

the first case only gas-phase (homogeneous) chemistry was considered as in the case of

an inert packed bed reactor. In the second case only heterogeneous (surface) reactions

were taken into account, as in the case of very low inlet temperatures. In the third case
both homogeneous (gas-phase) and heterogeneous (surface) reactions were considered, as

in the case of oxidation at higher temperatures.

The homogeneous (gas-phase) reaction rates for methane oxidation over the inert

pellets were modeled using an irreversible one-step reaction rate (Eq. (3.11)) in the
general form:

Eg
= Ag C 4C ex[ RTgJ (3.18)

It is readily acknowledged that this formulation for the reaction rate is in adequate to

represent its variation during the course of the oxidation process and sufficiently detailed

reaction kinetics need to be employed. However, for the overall modelling being pursued

here with the objective of obtaining a parametric study, this assumed form was

considered to be adequate for the purpose. The values of the pre-exponential factor "Ag"

and the activation energy "E.g"in the above equation (Eq. (3.18)) were first chosen on the

basis of the values available in the literature for methane-air mixtures in premixed flame

and were: A. = 7.90 x 1010 kmol/m3.s, and Eg = 202.60 kJ/mol (Westbrook, 1981). These

data have been used by other researchers and in some cases the value for power of the

fuel and oxidant concentration as well as the pre-exponential factor have been changed to

fit the conditions under.consideration. For example, Hautman et al. (1981) started with •

the above equation (Eq. 3.18) and developed amultiple step kinetic for the oxidation of

hydrocarbons within flow reactors and shock tubes. They developed rate expressions

which encompass an equivalence ratio range 0.12 to 2. 0, atemperature range from 960 to
51

1540 K, and apressure range 1to 9atm. Duterque et al. (198 1) altered the above reaction

to apply it intheir well-stirred reactor for the combustion of light hydrobarbon in air for

equivalence ratios from 0.65 to 1.2 and temperatures from 1450 to 1960 K. However,

Coffee (1985) used Westbrook's rate expression to obtain kinetic rate expressions and

data for stoichiometric and rich methane-air mixtures at atmospheric pressure. A wide

range of other formulation and corresponding data also can be found in the literature with
various degrees of agreement with experimental observations. Figure 3.4 shows methane

conversion as afunction of the inlet temperature for the equivalence ratios of 4) = 0.35

and u= 1.0 m/s using the kinetic data of Westbrook (1981). It can be seen that such data

cannot predict the experimental data of Depiak and Wierzba (1999) for methane

conversion in a packed bed reactor within reasonable accuracy. Therefore, kinetic


coefficients were obtained by matching the model prediction with the experimental data

of Depiak and Wierzba within the packed bed reactor. These acquired values which were

used throughout this investigation are: A. = 6.6 x 1010 kmollm3.s, Eg= 180.0 kJ/mol.

The heterogeneous (surface) reaction rates for methane over catalytic pellets were

modeled as an irreversible one-step reaction rate using the equation suggested by Song et
al., 1991:

CO AS CcH4 CexP[
Ts
] (3.19)

where the ',alues of the parameters A3 1.03 x i0 kmollm2.s, and E3= 86.0 kJ/mol were

chosen from (Trimm, and Lam, 1980 for V > 813 K) for methane-air mixtures with

platinum as the catalyst (i.e., platinum supported on porous alumina fibre). These

coefficients were used at the early stages of this research to investigate the behavior of

the packed bed reactor under different operating conditions. These results will be

discussed in this chapter. However, when using the above mentioned kinetic data in the

model could not predict to areasonable accuracy the experimental data of the packed bed

reactor. Therefore, new kinetic data were derived for the packed bed reactor in the

following chapters.
52

100 -

— Experimental (Depiak and Wierzba, 1999),


80 -. - Calculated (Westbrook kinetic data) ,

- 1
60
I

40 - I

I I

I
-

20 - I

A
AW

goo 900 1000 1100 1200


Inlet Temperature (K)
Fig 3.4 Comparison between predicted methane conversion using the kinetic data of

Westbrook (1981) and experimental data for methane conversion from Depiak and.

Wierzba (1999); u= 1.0 rn/s, 4) = 0.35, k3= 1.0 W/m.K, inert bed.

3.5.1 Inert Bed

The effect of the inlet gas temperature on methane conversion in areactor with an inert

bed due to gas-phase reactions is shown in Fig. 3.5 for the equivalence ratio (4)) of 0.35

and an approach velocity of u= 1.0 m/s (calculated at reference-unheated conditions) It

can be seen that below 900 K there is no significant conversion of methane. Rapid

chemical reaction occurs at an inlet temperature of -j 1080 K. It can also be seen that the

predicted results are in fair agreement with the corresponding experimental data of

Depiak and Wierzba (1999) for the '


same operating conditions. However, the predicted

conversion of methane at lo w
' temperatures is somewhat less than that obtained
experimentally.

As the equivalence ratio of these lean mixtures increases the conversion ofthe methane is

completed at lower temperatures, as shown in Fig. 3.6. The complete conversion of


53

methane in the mixture with an equivalence ratio of 0.50 takes place at an inlet

temperature of 860 K, while a temperature of 1200 K is needed for the complete

conversion of methane in the mixture with an equivalence ratio of 0.15.

The effect of approach velocity is shown in Fig. 3.7 for amixture with of 0.35. An

increase in the approach velocity causes a significant decrease in the conversion rate,

which was expected due to adecrease in the residence time. For example, at the inlet

temperature of 1088 K, the methane conversion was 100.0%, 12.2%, and 6.1% at the

approach velocities of 1.0, 2.0, and 3.0 m/s respectively.

100 A

- - - - Calculated (present work)


80 - A Experimental (Depiak and Wiex2ba)

.9 60

40

20

A AAA ±A A #
goo 900 1000 1100 1200
Inlet Temperature (K)

Fig. 3.5 Comparison between calculated and experimental results of methane conversion;

u= 1.0 m/s, = 0.35, ks= 1.0 W/m.K, inert bed.


54

100

----0.15 I
80— - --0.2
1 i
II
0.35

II
0.4
0.5
I
'-,

60 -
A 0.35,Exp.
.-
-

ii
40

ts

20— 1/
- I I
.0

I I
00 ' 700 800 900 1000 ' 100 1200
Met Temperature (K)
Fig. 3.6 Calculated methane conversion as afunction of inlet temperature at different

equivalence ratios; u = 1.0 m/s, k = 1.0 W/m.K, inert bed, (*) Depiak and Wierzba

(1999).

100

u1.0m/s
80 u=2.0 m/s
u3.Om/s

c40

20

-
- -
•__7 I I

900 1000 1100 1200


Inlet Temperature (K)

Fig. 3.7 Calculated methane conversion as afunction of inlet temperature at different

approach velocities; 4) = 0.35, k = 1.0 W/m.K, inert bed.


55

3.5.2 Heterogeneous Reactions

In order to examine the combustion processes under the idealized artificial condition of

the combustion process is assumed to be controlled by the surface reaction only (purely

heterogeneous reactions) in this case. It can be seen (Fig. 3.8) that, as in the first case of

homogeneous reactions (inert bed), methane conversion increases with an increase in the

inlet temperature. The temperature at which complete conversion of methane occurs is

higher than the corresponding temperature found for the case of solely gas-phase

reactions. This is areflection of the values for activation energy and the pre-exponential

factor (A = 1.03 x 10"5 kmol/m2.s, and Es = 86.0 kJ/mol; (Trimm and Lam, 1980))

chosen with platinum as catalyst and methane as the fuel that were obtained for a

different reactor configuration (see page 118 for the sensitivity of methane conversion

due to changes in the pre-exponential factor and activation energy).

It can also be seen that the increase in the equivalence ratio from 0.20 to 0.50 has

little effect on methane conversion. At the inlet temperature of 1075 K, a maximum

difference of 4.83% was obtained for methane conversion as the equivalence ratio

increased from k= 0.35 to 4= 0.50 at the light-off regime. The light-of regime

corresponded to inlet temperature ranges from 1050 to 1125 K. The most noticeable

conversion difference occurred within this regime, otherwise, it was negligible. Since,

there is no data available in the literature for methane conversion along the packed bed

for = 0.35, the range of equivalence ratio was extended to 4 0.71 in order to compare

the simulated methane conversion with the predicted data of Goralski and Schmidt (1999)

in aplug flow reactor. The conversion of methane as afunction of the axial position in

the reactor at = 0.71, u= 1.0 m/s and T = 1200 K is shown in Fig 3.9. Methane begins

to react immediately upon entering the catalytic bed and continues to react along the bed.

Goralski and Schmidt surface reactions 'are modeled with detailed kinetics using an

elementary 19-step mechanism for combustion of methane on platinum. The results are in

reasonable agreements. However, methane conversion obtained from the present

investigation is higher than that of Goralski and Schmidt taking into consideration that

the predicted methane conversion from the present study is calculated using one-step

irreversible surface reactions.


56

100

80

• 60

0
40

20

900 1000 1100 1200 1300


Inlet Tem perature (K)
Fig. 3.8 Calculated methane conversion as a function of inlet temperature for various

equivalence ratios; u= 1.0 m/s, k = 1.0 W/m.K, Pt catalyst.

100

80
"p

0
.-

20

00 0.01 0.02 0.03 0.04 0.05


Distance along the bed (m)

Fig. 3.9 Calculated methane conversion along the bed; 1i = 0.71, u= 1.0 m/s, Ti. = 1200

K, Pt catalyst.
57

The effect of the approach velocity on methane conversion is presented in Fig. 3.10 for 4
= 0.35. As in the case of inert beds the conversion of methane decreases with an increase

in the velocity. The maximum decrease in the conversion of-methane with an increase in

velocity is obtained in the light-off regime of the catalyst i.e. within the temperature

range from 1000 to 1200 K. At higher approach velocities, ahigher inlet temperature is

required to obtain the same conversion of methane (e.g., for a 50% conversion of

methane, the inlet temperatures were 1100, 1150, 1175 K at the approach velocities of

1.0, 2.0, and 3.0 m/s respectively.

100

80

• 60

rj 40

20

900 1000 1100 1200 1300


Inlet Temperature (K)
Fig. 3.10 Calculated methane conversion as afunction of inlet temperature for different

approach velocities; 0.3.5, k3= 1.0 W/m.K,Pt catalyst.


58

3.5.3 Coupled Homogeneous-Heterogeneous Reactions

In an actual catalytic bed both homogeneous and heterogeneous reactions are taking

place. Therefore, it is important to include both these reactions in the model using the

kinetic data for the gas-phase reactions of Ag = 6.6 x 1010 kmo11m3.s, Eg = 180.0 kJ/mol

and for the surface reactions A = 1.03 x i0 kmol/m2.s, and E = 86.0 kJ/mol.

The effect of the inlet temperature and mixture equivalence ratio on methane

conversion in the catalytic bed is shown in Fig. 3.11. As the equivalence ratio of the

mixture is increased, the conversion of the methane is completed at a much lower

temperatute. At temperatures lower than the light-off temperature fuel conversion is


negligible and once the light-off temperature is reached, the conversion rises rapidly. It

can be seen (Fig. 3.11) that the light-off temperature increases with decreasing

equivalence ratio. For example, the light-off temperature for avery lean mixture of j =

0.15 is -4000 K while that for the somewhat richer mixture of = 0.35 is -'740 K.

The effect of the approach velocity on methane conversion within the catalytic

bed is shown in Fig. 3.12. Similarly to the previous two cases (with the inert or

heterogeneous reaction only beds), methane conversion decreases with the increase in the

flow velocity (i.e. as the residence time decreases). For instance, at the velocity of 1.0

m/s, complete methane conversion occurs at an inlet temperature of - 750 K, while at

velocities of 2.0 and 3.0 m/s it occurs at much higher inlet temperatures of— 895 and 987
K respectively. In contrast to the case of heterogeneous reaction only (Fig. 3.10), it can

be seen that this effect of the approach velocity is less profound as the value of the

velocity increases. The same trend was observed experimentally by Depiak and Wierzba

(1999) for different catalysts such as cobalt oxides, chromium oxides and their binary

mixtures. The effect of approach velocity for equivalence ratio of 4= 0.1 on fuel

conversion is given in page 150.

It appears that the gas-phase reactions occurring alongside surface reactions in the packed

bed enhance the methane oxidation rate and allow it to occur at much lower inlet

temperatures than those achieved by the heterogeneous or homogeneous reactions alone.


59

800 900 1000 1100 1200


Inlet Temperature (K)
Fig. 3.11 Calculated methane conversion as afunction of inlet temperature at different

equivalence ratios; u= 1.0mJs,k = 1.0. W/m.K, Pt catalyst.

100 I

80

.2 60

0
40

U=1.0 mIs
20
u2.Om/s.
u3.Om/s

OS —_J I,
700 800 900 1000 1100
Inlet Temperature (K)
Fig. 3.12 Calculated methane conversion as afunction of inlet temperature at different

approach velocities; = 0.35, k5= 1.0 W/m.K, Pt catalyst.


60

3.5.4 Effect of Bed Porosity

The porosity of the bed, defined as the ratio of the volume of the void (i.e. gas volume) to

the total volume of the bed, depends on the solid particles size and the method of their

packing in the reactor. For example,. in conventional packed beds, such as distillation and

adsorption columns, the particles are of a relatively large size, typically 1-2 in. in

diameter (Cybulski and Moulijn, 1998). In catalytic beds, however, much smaller
particles are used, with typical diameters between I.and 3mm, and the porosity is much

lower, typically around 0.4. Enlargement of the catalyst particles will lead to apoor

utilization of the catalyst particles. Smaller pore sizes Swill result in higher catalyst

surface-area-to-volume ratios and will increase the relative rate of heterogeneou

reactions, whereas larger pores will tend to allow relatively more gas-phase reactions.

The large particle size also implies that apoor use is made of the reactor space, which is a

significant cost factor. For the same reason, a very loose packing with a very high

porosity will be unacceptable in such aprocess.

The effect of the bed porosity for the same geometrical shape of particles on

methane conversion due to heterogeneous reaction only within a bed 50 mm long is

shown in Figs. 3.13 and 3.14. Methane conversion at the exit of the bed decreases as the

bed porosity increases at the same temperature. Generally, such atrend can be expected

since an increase in the porosity means a smaller catalytic surface area. The largest

differences in methane conversion occur within the inlet temperature range of 1000-1150

K (Fig. 3.14), corresponding to the light off regime. Methane conversion at an inlet

temperature of 1100 K is about 27% higher for bed porosity of 0.20 than that for the bed
porosity of 0.45. For the higher porosity of 0.50 only negligible amounts (-1.0%) of

,methane conversion were ,obtained.


61

100
9.

1 Heterogeneous reaction only


80
S =0.30
8 - 0.35
. 60
4,)

20

=O5O

- 0
700 800 900
-9 .

1000 1100
_f:::.:L — —
1200
_.
1300
Inlet Temperature (K)

Fig. 3.13 Calculated methane conversion as afunction of inlet temperature for various

bed porosities; 4) = 0.35, u= 1.0 m/s, k = 1.0 W/(m.K), L = 50 mm.

100
Heterogeneous reaction only .

80 - - - - - T•m=1000 K
- T1=1050K
Tm=1100K
- T=1150K
. 60
4,)

C
40

20

I I I
00
0.1 0.2 0.3 0.4 05
Porosity

Fig. 3.14 Calculated methane conversion as afunction of bed porosity for different inlet

temperatures; 4) = 0.35, u= 1.0 m/s, k = 1.0 WIm.K, L = 50 mm.


62

The performance of the catalytic reactor model was different when both the

homogeneous and heterogeneous reactions were accounted for. It can bee seen (Fig. 3.15)

that even at the relatively low temperature of 750 K complete methane conversion could

be obtained within abed 50 iiim long for arange of the bed porosities (c = 0.20 - 0.40).

Moreover, with an increase in the bed porosity from 0.20 to 0.35, complete conversion

took place at an earlier location within the reactor. However, further increase in the bed

porosity to 0.40 resulted in the full consumption of methane at alocation farther from the

inlet to the reactor. And in the case of the porosity of 0.45 there was no methane
conversion at all within abed 50 mm long. Such abehavior is most probably due to the

following competing factors. An increase in the bed porosity results in an increase in the

bed temperture (Fig. 3.16) due to the reduced bed solid mass and its heat capacity. This

enhances the oxidation reaction. Moreover, the increase in the porosity results in a

decrease in the local velocity and an increase in the residence time within the bed. On the

other hand, the increase in the porosity decreases the catalytic surface area which is

essential in promoting the methane oxidation at such alow temperature (750 K). For the

bed porosity of 0.40 the later factor is probably dominant, which results in shifting the

oxidation reactions to alater (downstream) location within the bed. And it appears that

for the bed porosity of 0.45 the catalytic surface area is not enough to initiate the
oxidation process.

It is also worth to mention that the maximum value in the gas temperatures (peak)

for the mixture of = 0.35 are generated by the thermal energy released by methane. This

maximum value in the gas temperature diminishes as the mixture reaches the equilibrium

state due to heat exchange between the gas and solid phases by convection. Also, it is

worth emphasizing that the maximum gas temperature for the methane-air mixture is

below 1700 K. Thus, a rule-of-thumb, the thermal NO mechanism is usually


unimportant at temperatures below 1800 K.
63

0.02

0.0175

0.015

0.0125

0.01

. 0.0075

0.005

0.0025

0 •..__I I I I I a I
0 0.01 0.02 0.03 0.04 0.05
Distance Along the Bed (m)
Fig. 3.15 Effect of bed porosity on methane mass fraction along the packed bed for

=O.35, u=1.O m/s, k=l.O W/m.K, and T= 750 K; homogeneous-heterogeneous model.

1700

1600

1500

1400
N
c=0.20 =O.3O
N
s=O.35

1300
I e=O.40
1200 I
I
1100 I
I
1000 1
/
/
900 /

/ 0.45
800

700
0 0.01 0.02 0.03 0.04 0.05
Distance Along the Bed (m)
Fig. 3.16 Temperature distribution for the gas (solid line) and the solid (broken line) for

various bed porosities with 4i = 0.35, u = 1.0 m/s, ks = 1.0 Wlm.K and Tin = 750 K;

homogeneous-heterogeneous model.
64

3.5.5 Effect of Bed Length

It is expected that the bed length would influence fuel conversion. The effect of the bed

length on the methane conversion due to surface reactions only is shown in Fig. 3.17 for a

bed porosity of 0.5. It can be seen, that although no methane conversion was obtained

within a shorter bed of 50 mm long, the complete conversion of the methane was

obtained for abed of increased length (75 and 100 mm). Increasing the bed length while

other parameters kept constant increases the surface area of the catalyst as well as the

residence time, i.e. more time is available for the completion of the oxidation reactions.

The maximum methane conversion is obtained for bed porosity of c= 0.2, therefore, the

effect of the length of the bed at different inlet temperatures and bed porosity of 0.2 is

illustrated in Fig. 3.18. The maximum increase in the conversion of methane as the bed

length increased is at the inlet temperature of 1050 K which corresponds to the start of

the light-off regime of the catalyst. For example at this temperature, the increase in

methane conversion is 35.7 % with an increase in the bed length from 50 to 100 mm.

However, this increase in the methane conversion decreases to avalue of 17.2 % with an

increase in the bed length from 50 to 75mm. It can be seen also that increasing the inlet

temperature further diminishes the effect of the bed length (e.g., at the inlet temperature

of 1200 K). At such relatively high temperatures this behavior is expected because the

mass diffusion processes are controlling the combustion process. The results suggest that

the increase in the catalytic surface area as well as the heat capacity of the longer beds do,

not have a significant influence on methane conversion; such behavior is in agreement

with the reactor performance at the mass diffusion regime.

The effect of the length of the bed on methane conversion can also be seen in Figs. 3.19'

and 3.20 for different bed porosities. Comparison of the results (Figs. 3.13, 3.19 and

3.20) shows that at the same inlet temperatures and for the same porosity the methane

conversion values are significantly higher for the longer bed of 100 mm than those for the

shorter bed of 50 mm as expected. For instance, at the inlet temperatures of 1100 K for

the bed porbsity of 0.2 and bed lengths of 50 and 100 mm, methane conversion was

66.9% and 93.6% respectively. Moreover, for the longer bed of 100 mm complete

methane conversion can be achieved even with-a very high value of bed porosity of 0.9.
65

100
Heterogeneous reaction only

80
L=75min
100 mm

. 60

40

20
L=50mm

0-r - - ,-I J - - - l_I - - I_I -- + -,-

700 800 900 1000 1100 1200 1300


Inlet Temperature (K)
Fig. 3.17 Calculated methane conversion as afunction of inlet temperature for different

bed lengths; c= 0.50, 0.35, u= 1.0 m/s, k5= 1.0 W/ m.K; Pt catalyst.

100

80

g
• 60
qu
C
40

20

1000 K
Heterogeneous reaction only

0
50 75 100
Length of Bed (nun)

Fig. 3.18 Calculated methane conversion as a function of bed length for different inlet

temperatures; u= 1.0 m/s, 4= 0.35, s= 0.2, k 1.0 W/m.K; Pt catalyst.


66

100

80

• 60

0
40

20

700 800 900 1000 1100 1200 1300


Inlet Temperature (K)

Fig. 3.19 Calculated methane conversion as afunction of inlet temperature for different

bed porosities; L= 75 mm, 4= 0.35, u= 1.0 m/s, k = 1.0W! m.K; Pt catalyst.

800 900 1000 1100 1200 1300


Inlet Temperature (K)

Fig. 3.20 Calculated methane conversion as afunction of inlet temperature for different

bed porosities; L = 100 mm, 4= 0.35, u= 1.0 m/s, k = 1.0 W/ n.K; Pt catalyst.
67

It can also be seen (Fig. 3.21), that at the relatively high inlet temperature of 1300 K the
effect of porosity is rather small and that complete oxidation of methane is achieved

already at ashort distance from the inlet to the reactor for both porosities. However, at

the lower inlet temperatures, the effect of the porosity becomes significant. Few factdrs

contribute to the observed trends. As the catalytic surface area is increased so the surface

reactions increased, which is important to the heating of the solid. Expectedly, the

exothermic catalytic surface reaction will increase the solid temperature and this increase

in temperature will enhance methane conversion. Moreover, for the longer bed, its heat

capacity is larger which also facilitates an increase in the temperature of the bed. The

difference in methane mass fraction at the inlet temperature of 1100 K, is greater than at

1000 or 1200 K for bed porosity of 0.2 aiid 0.9 due to the high difference of methane

conversion at the inlet temperature of 1100 K.

Figure 3.22 illustrates the effect of both the bed length and bed porosity on

methane conversion at the inlet temperature of 1100 K. The highest methane conversion

is obtained for the bed length of 100 mm followed by the bed length of 75 mm for all

porosities employed in this investigation ranging from 0.2 to 0.9. For the bedlength of 50

mm, methane conversion is negligible for the bed porosity of s≥ 0.5 for the reasons

mentioned earlier. The results suggest that for abed length of 50 mm with bed porosities

of s≥0.5, the catalytic surface area has the key role in methane oxidation which seems as

if it is not large enough to initiate the oxidation of methane.

The temperature of the solid at the exit of the bed versus the bed porosity for

different inlet temperatures with abed length of 75 mm is presented in Fig. 3.23. It can be

seen from Fig. 3.23 that the solid temperature exceeds the inlet temperaure due to

exothermic reaction at the catalytic surface. This increase in the solid temperature is

always beneficial in enhancing the methane conversion. The smaller the bed porosity is
the higher the difference between the inlet and the solid temperature The maximum

effect occurs for inlet temperatures ranging from 1050 to 1100 K. (i.e., the light-off

regime where most of the methane conversion takes place) with the bed porosities of s<

0.5 where more than 57% of methane has been reacted.


68

0.02
t ..

0.0175 L
I .. .

I
0.015 + ..
a ••

0.0125 ..

Th •\*

0.01
:tt
-
•\
. 0.0075
T )
0.005 : --- ---- 1000
1100

- — 1300

0.0025 : -.-.- 1000


----.- 1100 ••_••_.._.._..

- - 1300 ,

I I

0 0 0.02 0.04 0.06 0.08 01


Distance along the bed (m)
Fig. 3.21 Methane mass fraction along the bed for various inlet temperatures for c= 0.2

(thick line) and a= 0.9 (thin line; *), k = 0.35, u = 1.0 m/s; heterogeneous reaction, Pt

catalyst.

100

Heterogeneous reaction only


..-

80
.5

"p . .5 -

. 60
S ..

C
cj 40

Tin = 1100 K
20 L =100nun
L=75mm
L=50mm

0 I I I
00 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10
Bed Porosity
Fig. 3.22 Calculated methane conversion as afunction of bed porosity for different bed

lengths; u=1.0 m/s, = 0.35; Ti,, = 1100 K, ks= 1.0 W/m.k; Pt catalyst.
69

1400
T in Heterogeneous reaction only
1300K

1200 K

1150K
1100K
1075 K
1050 K

_ . .-. - .-.
- . _ -

- 1000K - --------

I IIIIIIIII1Ii 11 I1111I1111I
I I IIIIIIIIIIIII i

00 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 10


Bed Porosity
Fig. 3.23 Bed solid temperature at the bed outlet as afunction of porosity for various inlet

temperatures; u= 1.0 m/s, 4) = 0.35, L = 75 mm, k = 1.0 W/m.K; Pt catalyst.

Furthermore, Fig. 3.23 shows that the solid temperature decreases as the bed porosity

increases due to the corresponding decrease in the catalytic surface area per volume.

Therefore, alower catalytic surface reaction rate will lead to lower solid temperatures and

lower methane conversion as a consequence. At much higher temperatures (i.e. Tin ≥

1175 K), methane conversion reaches 90 - 97% for bed porosities ranging from 0.2 - 0.9.

Therefore the heat release due to the surface reaction is much smaller than at the lower

inlet temperature. Thus, the increase in the solid temperature at the exit is very small.
70

In the case of coupled heterogeneous-homogeneous reactions, the effect of bed length is

less profound, as can be seen in Fig. 3.24 for = 0.35, u= 1.0 m/s at the inlet temperature

of Tir,= 750 K and for various bed lengths. Methane conversion for the bed porosity of s

= 0.40 is not enhanced by increasing the bed length from 50 to 100 mm. The location of

the complete oxidation of methane occurs at the same distance into the bed for both bed

length of 50 and 100 mm as shown in Fig. 3.24 (b) and (c). However, no methane

conversion for the shorter bed of 25 mm long was obtained (Fig. 3.27 (a)) at such low

temperature (750 K). It appears that the bed length is not enough to induce the oxidation

processes in the bed due to the smaller catalytic surface area and shorter residence time.

The effect of inlet temperatures on methane conversion for the longer bed of 100 mm is

illustrated in Fig. 3.25 in the gas and solid phases temperature distribution along the bed

for abed porosity of è= 0.45. No methane conversion is obtained at the inlet temperature

of 750 K. However, at the higher inlet temperature of 760 K, complete conversion of

methane is obtained at the exit of the bed. Increasing the inlet temperature further

allowed the complete conversion to take Olace at an earlier location into the reactor (e.g.,

methane oxidation completed at the 9mm for the inlet temperature of 800 K). The results'
suggest, that although the longer bed has larger catalytic surface area and longer residence

time, it is still not enough to initiate methane oxidation at the inlet temperature of 750 K.

Expectedly as the inlet temperature increases, the role of the gas-phase reactions
increases too and so does the conversion of methane.
71
1800

1600
(a)

1400

1200

E- 1000 Gas
- - - - Solid

800

I,I I II,, I-
0 0.005 0.01 0.015 0.02 0.025
Distance Along the Bed (m)
1800

1600 (b)

1400

1200

E
E
- 1000

Gas
- - - - solid
800

I I I r I I I
0.01 0.02 . 0.03 0.04 0.05

Distance Along the Bed (m)


1800

1600
(c)

1400

1200

B
E
- 1000

Gas
Solid
800

0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 01

Distance Along the Bed (m)


Fig. 3.24 Temperature distribution along the bed for different bed lengths; (a) L = 25

mm, (b) L 50 mm, (c) L = 100 mm, = 0.35, u = 1.0 m/s, F, = 0.4, T= 750K;

homogeneous-heterogeneous model, Pt catalyst.


72

The influence of both the bed length and bed porosity on methane conversion is presented

in Figs. 3.26 and 3.27 at the inlet temperature of Tin = 800 K for bed lengths of L = 100

and 50 mm respectively and for various bed porosities. The complete conversion of

methane is obtained for the bed porosities of s= 0.40, 0.45, 0.50, and 0.53 for L = 100

mm. A negligible amount of methane (i.e., less than 1.0%) is converted for bed porosities

of ≥ 0.55 as it can be seen from Fig. 3.27. Furthermore, the results suggest that the

optimum value of the bed porosity at this temperature (T m = 800 K) is the bed porosity of

= 0.40. Although the location of the complete oxidation of methane is negligible

between the bed porosities of 8 = 0.40 and 0.45, the temperature of the exhaust gas is

lower for the bed of s= 0.40. For the bed length of L = 50 mm (Fig. 3.27), the value of

the maximum porosity is c= 0.50 for the complete conversion of methane to take place.

As mentioned earlier for shorter beds the catalytic surface area is smaller and less

residence time is available for the fuel to oxidize.

The performance of the coupled heterogeneous-homogeneous reactions is different than

the case of the heterogeneous reaction only. It appears more likely in this case (eyen at

such moderate temperatures) that assuming the presence of gas-phase reactions occurring

alongside surface reactions in the bed, will offer a more effectively predictive and
sensitive model.-
73
1800

1600

a 1400

I
1200

1000

800
I I
0.00 0.02 0.04 0.06 0.08
Distance Along the Bed (m)
Fig. 3.25 Temperature distribution for the gas (solid line) and the solid (broken line)

along the packed bed for various inlet temperature; 4 = 0.35, u = 1.0 m/s, k8 = 1.0

W/m.K, s= 0.45, L = 100 mm; homogeneous-heterogeneous model, Pt catalyst.

1800

1600

a 1400

I 1200

ENO
1000
74

1800

1600 /

Tm =800K
1400
&= 0.45

1200 /
/
= 0.50
/
1000 /
I
0.52
800
I II I I I I 1I I , I
0.00 0.01, 0.02 0.03 . 0.04 0.05
Distance Along the Bed (m)

Fig. 3.27 Temperature distribution for the gas (solid line) and the solid (broken line) for

various bed porosities with 0.3 5, u= 1.0 m/s, k = 1.0W/m.K, Ti,= 800 K, L = 50

mm; homogeneous-heterogeneous model, Pt catalyst.


75

3.5.6 Effect of Thermal Conductivity of the Bed of Solids

Calculations show (Fig. 3.28) that an increase in the solid thermal conductivity of the bed

from 0.5 W/(m.K) to 4.0 W/(m.K) has no effect on methane conversion within abed of

50 mm long when only heterogeneous reaction is considered.

The methane mass fraction along the bed is plotted in Fig. 3.29 for the above

values of thermal conductivity at various inlet temperatures. The. effect of solid thermal

conductivity is negligible at the kinetic control regime of the catalyst where the methane

conversion is about 10% or less at the inlet temperature of Tj 1000 K. However, as the

inlet temperature increases and reaches into the light-off regime the effect of solid

thermal conductivity is more apparent on the mass fraction along the bed (i.e. at Tin =

1100 K). The increase in the inlet temperature further reduces the effect of solid thermal

conductivity on the mass fraction of methane along the bed. This corresponds to the

range of temperature between Ti,, = 1200 to 1250 K where most of the methane is already

consumed. It can be seen from Fig. 3.29 at an inlet temperature of Tin = 1100 K, that the

higher the conversion of the methane along the bed the more profound the effect of

thermal conductivity. At higher inlet temperatures (1200 and 1250 K), the influence of

the thermal conductivity is larger at the beginning of the bed where most of the methane

is oxidized. This effect diminishes as the methane conversion reaches completion along

the bed.
76

100

Heterogeneous reaction only

80 k0.5W/mK
k: = 1.0 W/mK
k =2.0 W/mK
. 60 k5 =3.OW/mK
k=4.0W/mK

(40

20

I I I I I i I r I

00 800 '900 1000 1100 1200 1300


Inlet Temperature (K)

Fig. 3.28 Calculated methane conversion as afunction of inlet temperature for various

solid thermal conductivities; 4) = 0.35, u = 1.0 m/s, L = 50 mm, a= 0.4; Pt catalyst.

0.02

0.0175

0.015
0

•' 0.0125

z 0.01

k1=0.5 W/mK
.0.0075
k8= 1.OW/niK
kg=20W/niK
0.005 k4.0W/mK

0.0025

0
0.01 0.02 0.03 0.04 0.05
Distance Along the Bed (m)
Fig. 3.29 Methane mass fraction along the bed at various inlet temperatures for different

solid thermal conductivities, 4) = 0.35, u = 1.0 m/s; heterogeneous reaction, Pt catalyst.'


77

The effect of the thermal conductivity when the coupled homogeneous-heterogeneous

reactions are considered in the model is shown Fig. 3.30. It can be seen that the value of

the thermal conductivity affected the location of complete methane conversion within the

bed. The higher the value of ksis, the shorter the distance needed to complete the methane

conversion. Figure 3.31 represents the temperature profiles for the gas and solid for the

above cases. It can be seen that an increase in the solid thermal conductivity will increase

the solid temperature significantly at the inlet part of the bed due to enhanced heat

conduction in the bed, and this will accelerate the oxidation process. Thus, the full

consumption of methane will take place at an earlier location within the catalytic reactor.
Also, it can be seen that higher values of thermal conductivity result in lower, exhaust gas

temperatures due to increased heat dissipation rates. It was also found that there will be

no conversion of methane for a solid thermal conductivity below avalue of k = 1.0

W/m.K. This is in agreement with the results of Groppi et al. (1999). This is the case

where heat conduction is negligible such as for ceramic material. However, for all values

of k = 1.0 - 4.0 W/m.K, complete conversion was achieved within abed 50 mm long at

Tin =750K.
78

0.020

0.018

to
0.015

M 0.013

0.010

0.008

0.005

0.003

0.01 0.02 0.03 0.04 0.05


Distance Along the Bed (m)
Fig. 3.30 Methane mass. fraction along the packed bed for different values of thermal

conductivities of the solid bed; = 0.35, u= 1.0 m/s, e= 0.4, T= 750 K; homogeneous-

heterogeneous reactions, Pt catalyst.

1600

1400

1200

1000

800

0.01 0.02 0.03 0.04 0.05


Distance Along the Bed (m)
Fig. 3.31 Temperature distribution for the gas (solid line) and the solid (broken line)

phases along the bed for different solid thermal conductivities; 0.35, u = 1.0 m/s,

s=0.4, T1 =750 K; homogeneous-heterogeneous reactions, Pt catalyst.


79

CHAPTER 4

EXPERIMENTAL APPARATUS AND PROCEDURE

4.1 Motivation

The approach of the present study has been to examine catalytically supported

combustion in awell characterized system in which most of the important physical and

chemical processes found in larger-scale catalytic combustors are present. The geometry

chosen is apacked bed reactor. Studies of catalytic combustion reported to date tend to,

stress either experimental or theoretical aspects, and few make the comparison between

theoretical and' experimental results necessary to understand the essence of surface

combustion (Mori et al., 1977; Groppi, 1999). The parallel development of anumerical

modeling program was necessary for the analysis of the experimental results obtained in
this study.

One of the objectives of this study was to obtain experimental data on fuel

conversion at different operational conditions (inlet temperature, equivalence ratio, and

approach velocity) and compare the data with the corresponding numerically predicted

data to determine the reaction rate constant and activation energy for aone-step surface

reaction. Moreover, the effectiveness of the catalyst for complete combustion of lean

fuel-air mixtures at relatively low temperatures 'and atmospheric pressure was to be


determined experimentally.
80

4.2 Experimental Apparatus

The apparatus required for this investigation had already been designed and was

operational (Mehta, 1990; Khalil, 1993; Depiak, 2003). However, some modifications

(and repairs) were made to meet the objectives of the present investigation. A schematic

diagram of the experimental apparatus is shown in Figure 4.1. The apparatus consists

mainly of a catalytic packed bed reactor, a compressed air supply and main heater

assembly capable of providing a wide range of air velocity and temperature, and

instrumentation (choked nozzle flow meters, thermocouples, gas analyzers for exhaust

gas diagnosis, etc.). A photographic view of the apparatus layout is shown in Fig. 4.2

which displays the above mentioned main components of the set-up. The main parts of

the apparatus are made of stainless steel (#316L) for durability as well as to withstand

continuous operation at high temperature.


2
Gas Cylinder Fuel Injector

4r T
r Thermocouple Read-out
System Exhaust Gas
Cooling System
1

I Main Heater
Catalytic Reactor
2

Temp. Controller Flame Trap

Auxiliary Heater
Gas Analyzers
( Holding,
Dryer
Tank

Legend
1. Thermocouples
2. Chocked nozzles flow meters

Compressor

Fig. 4.1 Schematic diagram of the experimental apparatus.


Fig. 4.2 A photographic view of the experimental apparatus.
83

4.2.1 Catalytic Reactor

The reactor is aquartz tube which has a28 mm inside diameter and is 300 mm long.

Figure 4.3 shows a' cross-section view of the reactor. An external electrical ceramic

radiant heater has been installed to provide a'100 mm long section of uniform

temperature (± 3 °C) along the reactor. This auxiliary heater is connected to adjustable

powerstats to obtain the desired uniform temperature. Within this section there is a50

mm long catalytic bed made of 3.2 mm diameter 7-alumina base pellets held in place by

two 25.4 mm long honeycomb monoliths (ceramic cordirite-Celcor; manufactured by

Corning Glass) as illustrated in Fig.4.4. An opening in the center of the honeycomb

monolith piece at the end of the packed pellet is made in order to slide in and out asheath

thermocouple measuring the exhaust gas temperature.

A high temperature heating tape (Omega series FGS high temperature heating tape of 627

Watts and 25.5 x 152.4 mm) was wrapped around the pipe system (including aflame trap
and fuel injector) connecting the main heater with the reactor to minimize the heat loss.

Also, aspecial insulation that can withstand high temperature was wrapped around the

piping system. The reactor can be seen in Figs. 4.2 - 4.4.


Radiant Heater
/
1

r Heating Tape / Packed Bed


/ / /
/ / I Insulation
/
/
4,
z. ••
Exhaust gas
/ sampling
/
probe

/
/
/

t
i i/I

1 / 1
/

_ Honeycomb
I
Monolith

• NN N \\
1/
Thermocouples V

Fig. 4.3 Cross-section view of the reactor.


Fig. 4.4 A photographic view of the experimental packed bed.
86

4.2.2 Compressed Air and Main Heater Assembly

Compressed air is supplied to the apparatus from the laboratory compressed air line. This

line is connected to an air filter to remove any moisture or oil residue in the air at the inlet

to the apparatus. A pressure regulator is installed after the air filter to regulate the air

supply into the choked nozzle assembly. This pressure regulator is able to provide

compressed air with maximum pressure of 100 psi. The compressed air at 'the required

mass flow rate is passed through achoked nozzle flow meter to determine the air flow

rate into the main heater. The main heater assembly consists of six 1.5 kW ceramic

heating elements made of Ni-Cr wire. These elements are mounted in astainless steel

rectangular box with gas inlet and outlet ports. They are controlled with aprogrammable

microprocessor-assisted temperature controller system. The main heater is connected to

high voltage power supply of 208 V. The details of the main heater assembly can be seen

elsewhere (Mehta, 1990). A maximum temperature of 625 °C was set in order to ensure a

longer running life for the heater's elements. The main heater assembly was controlled by

aPD temperature controller with aprogrammed microprocessor which maintained the

desired air temperature value for each running condition with accuracy of ± 0.2% of

reading. Temperature was measured by aK tye thermocouple placed at the exit of the

heater and connected to the temperature controller. This assembly is capable of supplying

air over awide range of temperatures (293 - 898 K).

4.2.3 The Choked Nozzle Meter System

Flow rates of the fuel and air entering the reactor were metered by apre-calibrated set of
choked nozzles. Figure 4.5 shows aphotographic view of the control panel of the choked

nozzle metering system used throughout the course of this study. A wet test meter and/or

orifice plates were used for calibration of the chocked nozzles depending on the flow

rate. Typical calibration figures are shown in Appendix B.


87

Fig. 4.5 A photographic view of the choked nozzle meters assembly.


88

4.2.4 Fuel Injection System

The system for fuel injection was designed to achieve ahomogeneous fuel-air mixture at

the inlet to the reactor. The radial-jet fuel injector has an inner chamber of 12.7 mm

diameter. The chamber contained 12 radially drilled holes of 0.4 mm diameter each, and

equally spaced around the inner periphery of the injector. The injector was positioned

approximately 500 mm upstream from the reactor. The fuel-air mixing was further

enhanced by the presence of aflame trap positioned downstream of the injector.

4.2.5 Exhaust Gas Analysis System

Gas samples were drawn during experiments using awater cooled 1/16-in (1.58-mm)

inside diameter stainless steel probe placed at the center line of the reactor at its end (Fig.

4.3). Use of the water cooled probe allowed the experimentor:

1. to quench the reactions within the gas sample quickly so that further chemical

reactions were halted (i.e. freezing the reactions),

2. to condense the water vapor in the exhaust gas,

3. to cool down the exhaust gas before it entered the gas analyzers. The exhaust

gas enters the probe at atemperature of up to 1400 K (this temperature depends

on the inlet mixture's temperature; equivalence ratio of the mixture and

percentage of the converted fuel), but enters the gas analyzers at atemperature
around 300 K.

Provision was made to process a gas sample using either gas analyzers or a gas

chromatograph. A valve was installed upstream of the probe connected to two

polypropylene tubing lines of 6-mm outside diameter. One line is connected to aset of

gas analyzers while the second line is connect to agas chromatograph for gas analysis.

The exhaust gas samples were withdrawn from the reactor with the aid of a sampling

system which includes asuction pump, awater trap, apenna drier, apressure regulator,

and a flow meter. In the present investigation the sample was delivered to the gas

analyzers at aflow rate and pressure of 800 ml/min and 112 kPa respectively. Species
89

concentrations in the exhaust gas were measured using Beckman gas analyzers (Model

864): an infra-red analyzer for hydrocarbons such as CH4 (0.05% accuracy), and infra-red

analyzers for CO and CO 2 (0.01% and 0.06% accuracy correspondingly). A magneto-

penu'matic analyzer (Model 755) was used to measure 02 concentration (0.1% accuracy).

The gas analysis system is shown in Fig. 4.2. A multi-channel data acquisition system
was connected to the gas analyzers to monitor the concentration of each constituent of the

exhaust gas.

Samples could also be analyzed using a3800 Varian gas chromatograph (GC) as
illustrated in Fig. 4.6. The sampling system in this case includes asuction pump, purge

line, and asample container that can hold up to six liters of sampled exhaust gas. This

sampling system is connected directly to the GC. The GC analysis was performed using

two columns in series. A flame ionized detector (FJD-GC) was employed to detect the

hydrocarbons such as CH4,C2H6, and C3H8. A thermal conductivity detector (TCD-GC)

was also used for detecting CO, CO 2,02, and N2 in the exhaust gas. The feed sample.

flow rates as well as the calibration gas flow rates into the GC were set at 100.0 ml/min.

The gas chromatograph was calibrated using standards for all the species listed above

except H20. From the preliminary test, the total time required for full analysis of the

sample, including a10-minute stabilizing period and 2-minute of retention period for the

FIDD column, as well as a 15-minute stabilizing period and a11-minute retention period

for TCD column, was 37 minutes. However, the total time required for full sample

analysis using the set of analyzers was only 4-5 minutes. Table 4.1 shows the value of

met
hane measured using the two apparati mentioned above for = 0.35 at different inlet

temperatures. The maximum difference in the value of measured methane between the

reading of the gas chromatograph and the gas analyzer was 0.71 % of reading. Thus it

was mostly the gas analyzers which were used throughout this investigation for
consistency and efficiency.
90

Table 4.1 Comparison between the measured methane using the gas chromatograph and

the gas analyzer for 4= 0.35 at different inlet temperatures.

Measured Methane
Measured Methane
Inlet Temperature using the Gas Difference
using the GC
(K) Analyzer (%)
(/o)
(%)

292 3.49 3.5 0.29

459 . . 3.5 0.0

524 3.46 . 3.45 0.29

573 3.44 3.44 0.0

651 3.41 3.40 0.0

726 3.33 3.35 0.60

773 3.19 3.18 0.31

798 2.91 2.90 0.34

869 2.78 2.8 ' 0.71


Fig. 4.6 Exhaust gas sampling probe with gas chromatograph.
92

4.2.6 Instrumentation and Measurements

• Temperature Measurement

The temperatures of the fuel-air mixture at the inlet, the outlet, and within the reactor

were measured by aset of Inconel 600 sheathed K-type (chromel-alumel, i.e., Ni-

Cr/Ni-Al) thermocouples with 1/16-in OD.

Thermocouples were also used to monitor the temperature of the air at the exit from

the main heater and the temperature of the wall at the inlet to the reactor. In order to

measure the gas temperature at the exit from the reactor, axial thermocouple was

placed through the honeycomb monolith as shown in Fig. 4.3. Two other
thermocouples were used for monitoring the wall temperature of the reactor catalytic

bed. The loëations of the above thermocouples are shown in Figs. 4.1 and 4.3.

The thermocouples were connected to athermocouple readout through aconnector

block which can accommodate six independent thermocouples. The temperature

distribution in the radial direction at the outlet from the reactor was measured during

the preliminary tests using atransverse thermocouple. This distribution is shown in


Appendix C Fig. C.1).

Measured temperatures can be different than actual gas temperatures due to heat

losses via conduction in the wires, radiation from the wires, and conduction and
convection from the gas to the wires. This analysis for the thermocouple correction is

shown in Appendix C. However, temperature values f'or, the experimental data

represent uncorrected measured values. The accuracy in the thermocouples

measurements for K-type is 2.2 °C or 0.75% of the reading whilst, an uncertainty of±

0.5% of the measured values is expected in the thermocouples readout.

• Pressure Measurement

The pressure drop in the packed bed reactor was measured using differential pressure

transducer. The pressure taps were mounted at the inlet and outlet of the 300 mm long

reactor. Figures D. 1and .D.2 (Appendix D) show the pressure drop along the packed
93

bed reactor which varied from 0.08 to 0.54 kPa ±0.5 Pa for the flow rates of 0.03 and

4.8 m3/h respectively. Such small pressure drops mean that the reactor effectively

operated under atmospheric pressure (the local atmospheric pressure is 89 kPa).

4.3 Experimental Procedure

The apparatus was regularly checked for leaks before taking measurements. The steps

required to prepare the experimental apparatus (i.e. calibration of equipments, check for
leaks, etc.) before acquiring each set of data were:

1. Opening of the fuel cylinder (i.e., standard calibration gases, fuel and nitrogen)

valves and setting the pressure gages to the desired pressure.

2. Calibration of all the gas analyzers; starting with the oxygen analyzer followed by

carbon dioxide; carbon monoxide and methane analyzers.

3. Calibration of the GC was checked once aweek only, in cases when the GC was
used.

4. Switching-on all the equipment and instrumentation being used (see section 4.2 of
this chapter).

5. Switching-on the water supply for the exhaust gas sampling probe.

6. Opening the valve to the gas analyzers or to GC.

7. Tuning the choked nozzle valve to the required pressure; calculated for adesired air
flow rate.

8. Setting amethane flow rate to the required value by adjusting the choked nozzle

pressure corresponding to the desired value of the fuel-air mixture equivalence


ratio.

9. Opening the valve for methane injection allowing it to flow through the reactor.
10. Analyzing the gas sample at room temperature. This step is used to ensure the

• required mixture equivalence ratio was obtained properly. Also, it is amethod for

eliminating the possibility of leaks throughout the system.


94

11. If the required equivalence ratio was obtained correctly, the methane injector valve
will be switched-off.

12. An extra precaution was employed throughout the experimental part of this

research. In cases of compressed air shut off due to compressor failure or leaks etc.,

a standby nitrogen cylinder was connected to the apparatus. The, compressed

nitrogen supplies aflow through the reactor which will cool,the main heater and the

catalytic bed as well to prevent damage to either part.

The following procedural steps were followed in obtaining the experimental data in the

order in which they appear.

1. Setting the main heater controller to the required air temperature.

2. Switching-on the main heater. The time needed to reach desired steady state

temperature depends on values of the air velocity and temperature. In those cases

when high inlet gas temperature experiments were conducted special attention was

paid to setting the required main heater temperature. A rate of 50 °C per five

minutes was adopted in raising the main heater temperature in order to prolong the

life of the elements of the main heater and to avoid more frequent repairs.

3. Switching-on the power of the auxiliary radiant and tape heaters and adjusting their
powerstats to the desirable temperature.

4. Recording the temperatures from all the thermocouples.

5. Opening the valve of the methane injector once the steady state temperature was
reached (usually, it took five to 15 minutes to reach asteady state temperature for

values between 400 and 900 K) and recording the temperature at the inlet and exit of
the reactor.

6. Analyzing exhaust gas sample withdrawn from the reactor after 30 seconds of the

methane injection. This time corresponds to the response -of the first gas analyzer

(i.e. 02 analyzer).

7. Recording the readings of all the analyzers as well as the thermocouples after a
steady conversion of methane is reached (usually , to 5minutes).

8. Switching-off the methane injection valve.


95

9. Resetting the main heater controller to ahigher inlet air temperature for the next run'.

10. Repeating steps 3to 8.

After covering the required range of inlet temperatures within the capability of the

apparatus, the following shut-off procedure was adopted:

1. Switching-off the power to the radiant and tape heaters.

2. Reducing the main heater set' temperature 50 °C per five minutes while only air is

flowing through the reactor until atemperature of 100 °C was reached. This measure

was adopted to help to prolong the life of the elements used in the main heater
assembly.

3. Switching-off the power to the main heater and allowing the cooling process of the

reactor.' -

4. Closing the upstream valve of the gas sample probe.

5. Shutting-off the water supply to the cooling sampling probe.

6.. Closing the methane cylinder valve as well as the standby nitrogen cylinder value.

7. Switching-off all other instrumentation.

4.4 Pellet Materials for the Packed Bed Reactor

Non-catalytic and catalytic pellets were employed in the reactor. For a non-atalytic

packed bed (inert packed bed) Gamma alumina oxide' (y-'A1203) pellets manufactured by

Johnson Mattey Inc. with aspecific surface area of about 300 m2/g were used (Alfa Aesar

catalog, A Johnson Matthey Company, 2001-02).

For catalytic bedpellets, two different catalysts were used, namely platinum and anon-

noble metal oxide catalyst (a binary mixture of cobalt oxide and chromium oxide,

Co 3O4JCr2O3,3.13 ratio by mass). These catalysts were deposited on the same gamma-

alumina oxide ('y-A1 203)pellets as the non-catalytic pellets. The loading of the platinum

catalyst, "manufactured by Johnson Mattey Inc., was 0.5% Pt by weight with aspecific
surface area of about 90 m2/g, while the loading of C0 304/Cr2O3 catalyst, made in house

(Depiak, 2003), was 5.33% by weight. ,


96

A set of preliminary tests was conducted to measure the apparent density and the porosity

of the packed beds. The average apparent density which includes the effect of the bed

porosity was found to be 100.0 kg/m3,the porosity of the bed was measured by the
investigator on the basis of several trials as being between 0.38 and 0.42 depend on the

packing of the pellets. These values were consistent with the values cited in the literature

for packed beds (Cybuiski, and Moulijn, 1998).

Experiments were conducted with methane as the primary fuel in this study which was

supplied from high pressure cylinders. Methane of high purity grade of 99.9% by volume

(<0.2% nitrogen balance) was used.

4.5 Apparatus Limitations and Operational Parameters

A series of preliminary tests were conducted which assisted in establishing operating

ranges and apparatus capabilities. The operating range of the apparatus was set by these

preliminar
ytests.

Experiments were conducted with lean methane-air mixtures of different equivalence


ratios (0.2 - 0.6) at different inlet temperatures up to 920 K and different approach

velocities up to 3.0 m/s, calculated at the reference temperature of 293 K and an

atmospheric pressure of 88 kPa.

It is customary in chemical industries to use the parameter "space velocity" instead of an


approach velocity. The space velocity "V5" is defined as the ratio of the volume flow rate

of the mixture "Q" (m3/h)to the volume of the reactor "V" (m) (Eq. (4.1)).

M /(rn 3h)) (4.1)

The value of the space velocity used in the present investigation based on the above

definition is between 54,000 and 216,000 (m3/(m3h)) corresponding to the approach

velocities between 0.75 and'3.0 m/s respectively.


97

The maximum inlet temperature ( 1040 K) for the inert packed bed was higher than the

above limited temperature value of 920 K. The maximum inlet temperature (1040 K) was

used at the early stage of this study and alarge number of difficulties and problems were

encountered. Thus, a decision was made to limit the operating inlet temperature to a

lower and safer operational value of 920 K.

Another limitation was discovered during the experiments. The temperature inside the

reactor could not exceed 1300. °C. Beyond such atemperature, the quartz could start to

soften and deform (melting point of quartz 1600 °C (1873 K). Thus, the temperature

inside the reactor was monitored and kept below 1200 °C for all the experiments

conducted in this investigation.


98

CHAPTER 5

EXPERIMENTAL RESULTS'

5.1 Introduction

The effect of various operational parameters, such as inlet temperature, equivalence ratio,

and approach velocity was investigated experimentally using the apparatus described in

Chapter 4. Three different beds were employed: inert (y-A1203 pellets); and two catalytic

('y-A1203 with 0.5% Pt and C0304/Cr2O3 (3.13 ratio by mass) with 5.33% loading by

weight). The majority of the experiments were conducted using the platinum catalytic

bed. The data obtainea were used to validate the mathematical Imodel and to obtain

kinetic data (activation energy and pre-exponential factor) for the surface reaction on Pt

and C0304/Cr2O3 catalysts.

The concentrations of the measured species in the exhaust gas were based on the

dry volumetdc percentage. They were corr


ected to account for the moisture (1120) that is

removed from the exhaust prior to analysis. This correction yields so-called wet
concentration. All the experimental data presented for species conversion were expressed

on wet basis: The exhaust gas correction is discussed in detail in Appendix E.

The value of the conversion of methane was calculated as:

CH 4 LCH 4 CH4,,,, - CH4,0.


(5.1)

where, CH4,
in represents the inlet concentration of methane in the mixture and CH4, 0 t

represents the corrected measured methane concentration at the exit of the reactor.
99

5.2 Inert Packed Bed

The experimental data obtained during the present work as well as the experimental data

of Depiak and Wierzba (1999) have been used to find the kinetic data for the numerical
simulation of the gas-phase reactions within the inert packed bed. The same kinetic data

for the gas-phase reactions have also been used in the numerical simulation within 'the
catalytic beds.

The results of the methane oxidation in air at equivalence ratio 4 = 0.35 and approach

velocity u = 1.0 m/s are shown in Fig. 5.1 for the inert bed. It can be seen that the

noticeable conversion within the inert bed starts at arelatively high temperature 800 K.

The conversion increases slowly up to atemperature of 1078 K when rapid conversion

(ignition) takes place. For comparison the predicted results (using the model described)

and experimental data of Depiak and Wierzba (1999) were also ihown in Fig. 5.1. It can

be seen that they were in afair agreement, especially at temperatures greater than 1000

K. However, the predicted conversion of methane at low temperatures is somewhat.

smaller than that obtained experimentally.

The exhaust gas concentrations of CH4,CO 2,and CO within the inert bed over the entire

inlet temperature range for the above condition of 4= 0.35, u= 1.0 m/s are plotted in Fig.

5.2. There is anoticeable CO concentration detected in the exhaust at the middle inlet

temperature range. The maximum measured concentration of CO is obtained at an inlet

temperature of- 1020 K. Beyond this temperature (1020 K), the methane concentration

becomes significantly low due to higher reaction rates and consequently the CO
concentration begins to fall producing essentially a zero concentration at the inlet

temperature of 1060 K and above.


100

100

90
Inert Bed
80
—A Experimental data (present work)
— — — — Calculated data (present work)
' 70
—a Experimental data Depiak & Wierzba (1999)

• 60

50

40

30
(-)
20

10

fi
200 400 600 800 1000 1200
Temperature (K)
Fig. 5.1 Comparison between experimental and calculated methane conversion within the

inert bed for 4 = 0.35, u= 1.0 m/s.

4 0.4

3.5 0.35

0.25

Inert Bed
0.2
—n---- CH
—e--- Co,
.A Co 0.15

6 0.5 0.05

0 400 600 I 800 ioo' 0


Temperature (K)
Fig. 5.2 Exhaust concentrations of CH4,CO 2 and CO as afunction of inlet temperature

within the inert bed for = 0.35, u= 1.0 m/s.


101

5.3 Catalytic Packed Bed

5.3.1 Effect of Catalyst Aging

The results of methane combustion obtained with fresh platinum catalyst are presented in

Figs. 5.3 and 5.4 with 4 = 0.35 and u = 1.0 m/s. It can be seen that at temperatures

between 673 and 723 K, the conversion of methane over the fresh catalyst dropped

sharply and is consistent with CO 2 concentration. Such behavior could be attributed to the
following factors:

1. Initially anti-oxidant s
ealant (LPS Anti-Seize) was used between the flanges as

well as on the screws connecting the reactor to the apparatus. This sealant

contains lead and zinc, the vapors of which can poison the platinum catalyst

(Hagen, 1999; Cybulski, and Moulijn, 1998). At higher temperatures the vapors

of the sealant could have been deposited on the external surface of the catalytic

pellets and have block access to their interior. Thus, the deactivation of the

catalyst could have occurred by pore blockage. This conclusion was drawn after a

sudden stop and resumption of the conversion with inlet temperature between 673

and 723 K. Since with an increase in the inlet temperature above 700 K the

conversion of methane also increased as shown in Fig. 5.4. This is probably due

to evaporation of the sealant and the consequent regeneration of the catalyst.

Moreover, the lead poisoning is reversed athigher temperatures (T> 1000 K) due

to the decomposing and volatilization of the lead compounds (Kesselring et al.

1979). It can be seen in Fig. 5.5 that the temperature at the exit from the reactor

reached - 1275 K. The reactor was dismantled and cleaned of the sealant. Another

set of experimental data was obtained after reassembling the reactor with the same

fresh catalyst which had been used for five hours only.
102

50

40 - ----- Fresh Pt Catalyst -

0"

© 30

C_) 20

10

I I 1 9 1 I
00 400 500 600 700 800
Inlet Temperature (K)

Fig. 5.3 Effect of deactivation of fresh platinum catalyst 'on methane conversion, k =

0.35, u= 1.0 m/s.

4.0

(3.5
0 •'
1- 1

0 3.0

2.5
Pt Catalyst
2.0
: —a--- CH Concentration
—4----- CO, Concentration

1.0

0.5

0 0 cJ
bo
- -

400 500 600 700 800 900 100


Inlet Temperature (K)
Fig. 5.4 Exhaust concentrations of CH4 and CO 2 as afunction of inlet temperature; fresh

platinum catalyst, 1= 0.35, u= 1.0 m/s.


103

2. Sintering of platinum catalyst caused by thermal stress. It can be seen (Fig 5.5)

that at the inlet temperature of 673K, 40% of methane was oxidized and the

exhaust gas temperature reached 1275 K.

It had been reported in the literature (Prasad, 1984) that platinum sinters

rapidly at temperatures of 500 - 900 °C (i.e., 773 - 1173 K). It was also found.

that platinum at temperature above 585 °C will form large metallic crystállites.

This change in crystallite size is known as the agglomeration of platinum. The

formation of large crystallites of platinum results in much reduced activity (Lee,

and Ruckenstein, 1983; Ruckeiistein, and Dadyburjor, 1983; Butt, and Petersen,

1988; Hagen, 1999; Cybuiski, and Moulijn, 1998).

A change in the color of the catalytic pellets was observed after

conducting experiments for more than 130 hours. The color of the fresh catalytic

pellets changed from metallic medium/dark gray to alighter gray after using them

for 19 hours and to avery light gray (almost white) after 131.5 hours. Other

researchers have also observed similar changes -in catalys


t color and the

corresponding activity of their catalysts (Bharadwaj, and. Schmidt, 1994;

Torniainen et al., 1994). They explained that the changes in color of the catalyst

are due to sintering of the platinum metal into largei particles with smaller

surface-to-volume ratios, and to the exposure of Al2O3.Large crystals of platinum

metal (- 100 j.tm) were observed on their used catalysts. .

3. A series of phase changes of the catalyst support, y-A1203,occur when it is heated

to highertemperatures. y-Al2O3 with a specific high surface area of about 300

m2/g can be transformed into cc-A1203 with its specific surface area of less than 10

m2/g. Such transformation takes place at temperatures around 1173 K. Sintering

thus results in pore closure and aburying of active catalytic sites in the alumina

support (Prasad, 1984; Richardson, 1989; Cybulski, and Moulijn, 1994 and 1998.-

Hagen, 1999; Alfa Aesar catalog, 2001-02).


104

A significant change in the performance between the fresh catalyst used for 5.0 h and

used catalyst for 131.5 h(aged) was also observed (se Fig. 5.6). At temperatures below

750 K, the highest conversion obtained was over the'catalyst which was 5.0 h old. At

these temperatures the difference in methane conversion over the catalysts aged for 19.5
and 131.5 hwas very small.

The exhaust as temperatures were measured at the exit of the packed bed with the 5.0 h

platinum catalyst with u = 1.0 m/s and 4 = 0.35. They are plotted against the inlet

temperatures in Fig. 5.7 together with the corresponding values of the conversion of

methane. The increasein the exit gas temperature is linear and slow for inlet temperatures

below 625 K. The exhaust gas temperatures at the inlet temperature above 625 K (which

correspond to the start of the light-off iegime) increase significantly due to asubstantial

increas
e in the rate of methane oxidation. A corresponding increase in exhaust

temperature was observed. For example at inlet temperature of 748 K the exhaust

temperature was 1350 K. This value is in afair agreement with the value of 1364 K

obtained by Bond et al. (1996) for 38% of methane conversion. The corresponding
calculated adiabatic flame temperature (for 40% of converted methane) is 1429 K which

is rather in good agreement taking into consideration that the measured temperature was

not corrected for the heat loss. However, the error in the thermocouple measurements due

to radiation and convection is calculated (Appendix C) and the r


esults are shown in Fig.

5.7. The repeatability in the experimental methane conversion is also shown in Fig. 5.7
with amaximum error of 4.8%.
105

1400 100

1200
80 '

1000 0

60

800
0
0
40
600

20
400

200 1 0, I I I I I I I I o
400 500 600 700 . 800
Inlet Temperature (K)
Fig. 5.5 Exhaust gas temperature and methane conversion as a function of inlet

temperature; fresh platinum catalyst, 41 = 0.35, u= 1.0 m/s.

100

Pt Catalyst
80
• Fresh catalyst (5.0 h)
—4k----- Aged catalyst 19.5 h
---- Aged catalyst l3l.5h

260

20

I t I I I In I I II II I I I I I

400 500 600 700 800 900 1000


Inlet Temperature (K)
Fig. 5.6 Comparison of methane conversion over fresh and aged platinum catalyst 41 =

0.35, u= 1.0 m/s.


106

1400 100

Pt Catalyst
1200 Temperature

a
- 80
conversion

1000 0

- 60

800

- 40
600

20
400

200
400 500 600 700 800 0
Inlet Temperature (K)
Fig. 5.7 Exhaust gas temperature and methane conversion as a function of inlet

temperature; fresh (5.0 h) platinum catalyst, = 0.35, u= 1.0 mi's.


107

5.3.2 Combustion of Methane over Aged Platinum Catalyst

Aging of the platinum catalyst affected the conversion of the fuel drastically. Therefore,

aged catalysts were used to obtain trends on the effects of inlet temperature, equivalence

ratio, and approach velocity independent of the current activity of the catalyst. All the

results below were obtained, with the aged catalyst. Experiments were performed over the

range of equivalence ratios from 0.2 to 0.6 with approach velocities varied from 0:75 to

3.0 m/s. Howevef, -some experiments were carried with fresh platinum catalyst for

comparison purposes.

It can be seen from'


Fig. 5.8 that.noticeable methane conversion starts at much lower inlet

±emperatures (500 K) than in the case of the inert bed. It was also illustrated that methane

conversion increases with increasing inlet temperatures. Generally, at low temperatures,

conversion is controlled by reaction kinetics ("kinetic limitation") at the catalytic surface


(Bond, et al., 1996; Pfefflerle and Pfefflerle, 1987), while at higher temperatures, surface

reactions are fast compared to the rate of mass diffusion to the surface ("diffusion

limitation"). Between these two regimes exists an unstable operating region, where the

methane conversion depends very strongly on the inlet temperature. For temperatures

above 800K this sudden increase in conversions is observed for equivalence ratios of

>0.35.

The effect of equivalence ratio on methane conversion for the approach velocity of u

1.0 m/s,is illustrated in Figs. 5.8 and 5.9. An increase in the methane equivalence ratio

increased the conversion of methane for the entire inlet temperature range employed. At

temperatures below 500 K methane oxidation for all equivalence ratios is negligible and

at relatively low temperature about 650 K, the fractional conversion is approximately

independent of the equivalence ratio. Methane conversion -


is much higher in the mixture

with the equivalence ratio of c,f = 0.6 at temperatures above 700 K probably due to the

fact that such ejuivalence ratio is within the flammability range of methane. The

conversion improves significantly with increasing inlet temperature (Fig. 5.9).


108

300 400. 500 600 700 800 900 1000


Inlet Temperature (K)
Fig. 5.8 Methane conversion as afunction of inlet temperature for different equivalence

ratios, u= 1.0 m/s; aged Pt catalyst.

100

Pt Catalyst
80
----- 1573K
-,--- T=698K
—a---- T,=74SK
—4-- T823K
• 60 ' 1848K
41

40

20

0.2 0.4- 0.6 08


Equivalence Ratio
Fig. 5.9 Methane conversion versus equivalence ratio'at different inlet temperatures, u=

1.0 m/s; aged Pt catalyst.


109

As expected, the strong effect 'of inlet temperature is apparent throughout the entire range

of equivalence ratios tested. For instance, the conversion increases from 12.9% to 36.4%

as the equivalence ratio increases from 0.20 to 0.60, respectively at an inlet temperature

of 823 K. A similar trend, i.e. the increase in methane conversion with an increase in the

equivalence ratio over platinum catalysts supported on ceramic honeycomb monoliths

was also reported by Bond et al. (1996). However, their measured methane conversions at

850 K inlet gas temperature are higher than those obtained in this investigation, 13.5%

and 38% for equivalence ratios of 0.18 and 0.30, respectively. It appears that at higher

temperatures the performance improved for all equivalence ratios although above 850 °C

(i.e. 1123 K) the platinum catalyst exhibited achange in characteristics due to thermal

stress; in addition to the phase change of the catalyst support. It is worth mentioning that

the experiment with the equivalence ratio of 0.35 was conducted first. It was followed by

experiments with the equivalence ratios of 0.20, 0.50, and. 0.60. As pointed out, the

results suggest that aging of the platinum catalyst may influence the conversion curves
for higher equivalence ratios as discussed in the previous section.

. Experiments were also performed to determine the effect of approach velocity on the

reactor performance. The velocity was calculated based on the reference temperature of

•293 K and the atmospheric pressure of 88 kPa. Approach velocities were varied from

0.75 to 3.0 m/s for the mixture equivalehce ratio of = 0.35. The results are shown in

Figs 5.10 and 5.11. The expected decrease in methane conversion was observed at higher

velocities. It can be seen that as the approach velocity decreases, the light-off position is

progressively shifted to lower inlet temperatures. At higher inlet 'temperatures the effect

of velocities is somewhat more pronounced, e.g. at the hilet temperatures of 823 atid 598

K, the methane conversion decreased from 28.2% to 13.7% and from 5.9% to 0.0%

correspondingly, when the velocity increased from 1.0 to 3.0 m/s. Higher approach

velocities correspond to reduced residence time (contact time between the catalysts and

the fuel). This trend has also been observed experimentally and numerically by other

researchers (Dupont et al., 2000; Moallemi et al., 1999). -


110

50

Pt Catalyst
40
—A---- uO.75m/s
X u1.Om/s
— y u2.Om/s
•— u- 2.5m/s
• 30 • u3.Om/s

C20

10

400 500 600 700 800 900 1000


Inlet Temperature (K)

Fig. 5.10 Methane conversion as afunction of inlet temperature at different approach

velocities with = 0.35; aged Pt catalyst.

50

Pt Catalyst
40
—a--- T=598K
A T698K
T,=748K
T=823K
30 -.---

I
c320

10

%0 0.5 1.0.. 1.5 2.0 2.5 3.0 35


Approach Velocity (m/sec)
Fig. 5.11 Effect of approach velocity on methane conversion at different inlet

temperature with = 0.35; aged Pt catalyst.


111

5.4 Oxidation of Methane over Binary (Co 3O4ICr2O3)Catalyst

The performance of C0 304/Cr2O3 (3.13 ratio by mass) within the reactor in methane

oxidation was investigated and was compared to those of Depiak (2003) in Fig. 5.12 at

various inlet temperatures for 4, = 0.35 and u= 1.0 m/s. In Depiak's work the conversion

of methane plotted against measured temperatures which were taken at the middle of the

reactor. Therefore, these measured temperatures were adjusted based on his temperature

distribution along the bed to reflect the inlet temperatures corresponding to this work and

compared to the values of the present investigation. Moreover, they exhibited high

temperature stability over a range of temperatures, which are typical for catalytic

combustors. The corresponding exhaust concentrations of CH4,CO, and CO 2 with

C0304/Cr2O3 catalyst at the above mentioned operational conditions are shown in Fig.

5.13. At low temperatures, the oxidation was relatively small' However, the oxidation of

methane at the inlet temperature of 750 K was increased significantly. At the inlet

temperatures above 600 K a noticeable amount of CO was measured. With further

increase in the temperature value the CO concentration in exhaust gases increased

significantly. At an inlet temperature of 880 K, the detected CO concentration was

0.095%.

The effect of equivalence ratio on methane conversion with the C0304/Cr2O3 catalyst is

illustrated in Fig. 5.14 at different inlet temperatures. The trends observed are similar to

those for Pt catalyst i.e. as the equivalence ratio increased so does, the conversion of

methane. It was also found that the concentration of CO increased with an increase in

equivalence ratio, especially in the region of light-off temperatures, for example from

0.063% for 4, = 0.35 to 0.165% for 4, = 0.50 at inlet temperature of 798 K as shown in

Fig. 5.15.
112

500 600 700 800 900 1000


Inlet Temperature (K)
Fig. 5.12 Methane conversion as afunction of inlet temperature within the C0304/Cr2O3

catalytic bed, = 0.35 and u= 1.0 m/s.

4.0 0.40

0.35

Co 3O4ICr2O3 Catalyst

—4---- CH4
CO'
• co

- 0.05

300 400 500 600 700 800 900 i00%-oo


Inlet Temperature (K)
Fig. 5.13 Exhaust concentrations of CH4,CO and CO 2 as afunction of inlet temperature;

C0304/Cr2O3 catalyst, u= 1.0 m/s and 4i = 0.35.


113

400 500 600 700 800 900 1000


Inlet Tempemtwie (K)
Fig. 5.14 Methane conversion as afunction of inlet temperature for different equivalence

ratios; Co 3O4JCr2O3 catalyst, u= 1.0 m/s.

0.20

C0 304/Cr2O3 Catalyst

0.15
• 0.35
0.50

C
.,

0.10

0.05

0.00
P & I
300 400 500 610 700 800 900 1000
Inlet Temperature (K)
Fig. 5.15 Exhaust concentration of CO as afunction of inlet temperature for different

equivalence ratios; C0304/Cr2O3 catalyst, u= 1.0 rn/s.


114

Furthermore, the performance of C0 304/Cr2O3catalyst (aged for 17.0 h) was compared to

that of the platinum catalysts (fresh and aged) as shown in Fig. 5.16. In fact, platinum and

palladium are the most commonly used catalysts in catalytic combustion applications for

their high specific activity in comparison to base metal oxides. However, practical

limitations such as high volatility and limited supply restrict their use. Therefore, finding

alternative catalysts of base metal oxides is of great importance. Metal oxide catalysts

typically have lower catalytic activity and higher light-off temperatures than noble metal

catalysts. This is demonstrated in Fig. 5.16, in which the performance of afresh (5.0 h)

platinum catalyst is much better than that of Co 3O4JCr2O3.The light-off temperature for a

binary catalyst of cobalt and chromium oxides is around 750 K while avalue of 610 K

was identified for fresh (5.0 h) platinum catalyst. At an inlet temperature of 723 K, the

difference in the conversion of methane between these two catalysts is very large about

78.57%. However, Co3O/Cr2O3 catalysts were more stable at high temperatures. Als
o,

Fig. 5.16 illustrates clearly that the effect of the catalyst aging on methane conversion for

platinum catalyst in comparison to C0 304/Cr2O3 catalyst, which is more thermally stable

at high temperature. It has been shown in the literature (Prasad et al., 1981) that the

C0304/Cr2O3 catalysts have been tested for up to 400 h of operation at maximum


temperatures of up to 1373 K without significant deactivation. Furthermore, Depiak

(2003) has shown insignificant effect of aging for 50 hof the C0 304/Cr2O3 (0.33 ratio by

mass) catalysts exposed to the .temperatures of 800 K and above on methane conversion

for 4
= 0.35 and u= 1.0 m/s. It can be seen that the fresh (5.0 h) Pt catalyst and one aged

for 19.5 hperform much better than C0 304/Cr2O3 catalysts. However platinum catalysts

are not able to sustain their activity over longer exposures to high temperature. Thus,

C0 304/Cr2O3 catalysts showed better performance in methane conversion than Pt ones

aged for 131.5 h. Both catalysts show similar performance before the light-off

temperatures were reached. The exhaust concentration of CH4 and CO 2 for fresh (5.0 h)

Pt catalyst was compared with that of C0 304/Cr2O3 catalysts in Fig. 5.17 under the same

conditions of tfi = 0.35 and u = 1.0 m/s. At all temperatures no CO was detected in the

exhaust with Pt catalyst. Both catalysts exhibited similar behavior with shifted curves

toward the higher temperature region for C0304/Cr2O3 catalysts.


115

100

• Fresh Pt catalyst (5.0 h)


80
—'v'---- Aged Pt catalyst 19.5 h
-4----- Aged Pt catalyst 131.5 h
0'
—*----- Co,04/Cr,O, catalyst

• 60

Q 40

20

0 ej
400 500 600 700 800 900 1000
Inlet Temperature (K)
Fig. 5.16 Comparison between Pt and C0304/Cr2O3 catalysts for methane conversion at

different inlet temperatures, u= 1.0 m/s, 4) = 0.35.

4.0

3.5

CH4 (Pt catalyst, 5.0 h)


• CO 2 (Pt catalyst, 5.0 h)
o CH 4 (Co,O/Cr,O, catalyst)
o CO 2 (Co,O 4ICr2O, catalyst)

400 500 600 700 800 900 1000


Inlet Temperature (K)
Fig. 5.17 Exhaust concentration of CI- 4 and CO 2 as afunction of inlet temperature; Pt

and Co 3O4JCr2O3 catalysts, u= 1.0 m/s, 4) = 0.35.


116

5.5 Summary

The effects of inlet temperature, equivalence ratio, and approach velocity on methane

conversion over inert and catalytic beds were investigated experimentally. It is found that

there is aneed for higher inlet temperature (Tj> 1078 K) to initiate methane conversion

over the inert bed, while relatively low temperature (T m > 620 K) is needed for methane

conversion by utilizing catalytic bed. In general, the conversion of methane increased

with increasing the inlet temperature and equivalence ratio or decreasing the approach

velocity. It was also found that platinum catalyst exhibited aging within five hours of
usage, which affect the conversion of methane drastically.

The main objective of this experimental work is to establish kinetic data for methane

conversion within the packed bed reactor needed for the mathematical modeling, which is

the topic of the next Chapter.


117

CHAPTER 6

NUMERICAL SIMULATION OF METHANE COMBUSTION IN A

PACKED BED REACTOR

6.1 Introduction

The model developed for the present investigation includes two overall reaction rates,

one .forthe catalytic reaction occurring at the surface and the other for ahomogeneous

reaction within the gas-phase. The rate expressions for asingle-step overall irreversible

homogeneous reaction (Eqs. (3.11) and (3.18)) and for asurface reaction (Eqs. (3.12) and

(3.19)) of the Arrhenius type were used. This research intended to investigate whether

such simplified rate expressions obtained from the literature can predict satisfactorily

some of the experimentally observed trends of such complex systems. It also intended to

derive from experimental observations kinetic data for methane oxidation in the packed

bed reactor. The effect of some of the operating parameters on the methane conversion
within the packed bed addressed in the preliminary results is reconsidered in this Chapter

using the kinetic data obtained from the present work.

6.2 Determination of Kinetic Data for Surface Reactions

6.2.1 Platinum Catalyst

Results of the numerical simulation obtained with kinetic data chosen from the literature

for similar conditions, specifically, Trimm and Lam (1980), Song et al. (1991) and

Markatou et al. (1993) do not match well our experimental results as was mentioned
118

earlier (Fig. 3.14). Therefore, the pre-exponential factor and the activation energy of the

surface reaction were obtained from the experimental data under the conditions in the
packed bed reactor, using the following procedure:

1. Substituting equation (3.12) for the surface reaction rate (i.e. heterogeneous reaction

rate) in the energy equation of the solid (pellets) in rder to include the surface

reaction effect on methane conversion within the catalytic bed.

2. Choosing as a first approximation a value for the activation energy and pre-

exponential factor for the surface reaction from the literature. Trimm and Lam
(1980) give two values of (E, = 86 kJ/mol. and B8 = 46 kJ/mol with As = 1.03x10 5

kmoll(m2.K)) for methane-air mixtures with platinum as the catalyst.

3. Calculating the methane conversion curve as afunction of the inlet temperature.


4. Comparing the predicted methane conversion curve with the experiment.

5. If the calculated and measured methane conversion curves in step (3 and 4) were far

apart, the value' of the pre-exponential factor was changed, while the value of the

activation energy was held fixed. The value of the pre-exponential factor was varied

until the model reasonably predicts the methane conversion. This procedure was

done only for one set of data (4 = 0.35, u 1.0 m/s).

6. Changing the value of the activation energy in order to reach even better matching

between the experimental and predicted methane conversion curves as afunction of


the inlet temperature.

7. Repeating steps (3) and (4) every time the activation energy value was changed until

a good agreement was reached between the two mentioned curves of methane
conversion.

8. The final values of the pre-exponential factor and activation energy with the best

match between one set of experimental data and predicted methane conversion will
be the adopted value of the kinetic data for the platinum catalyst within the packed

bed reactor.
119

Figure 6.1 shows the methane conversion curves corresponding to different chosen values

of the activation energy and the pre-exponential factor used to determine the kinetic data

for methane-air mixture of 4= 0.35 and at u= 1.0 m/s over the fresh platinum catalyst.

Furthermore, Fig. 6.2 shows the experimental data and thQ predicted methane conversion

obtained by using values of the pre-exponential factor and the activation energy obtained

from the present investigation for two temperature regimes. The resulting activation

energy of the surface reaction was found to be 46.0 ± 1.0 kJ/mol over the. temperature

range 750 - 925 K. Moreover, the obtained kinetics of methane oxidation shows achange

in the activation energy at temperatures below 750 K. The value of the activation energy

of 42.0 ± 0.8 kJ/mol at temperatures < 750 K was found to better match the experimental

results. The pre-exponential value was found to be 1.03 x 10 -3 kmol/(m2.


s) for the entire

range of temperature. The predicted results for methane conversion at different

temperatures using the above mentioned values of activation energy and pre-exponential

factor showed good agreement with the experimental data obtained during this

investigation as well as with data of Depiak and Wierzba (1999).

The conversion of the methane at = 0.50 using the pre-exponential factor and

surface activation energy obtained during this investigation was plotted in Fig. 6.3 with u

= 1.0 m/s. The oxidation of methane was enhanced by increasing the concentration of the

methane at the inlet.

The increase in the approach velocity from 1.0 to 3.0 m/s showed areasonable

agreement with the experimental data of Depiak (2003) using the above mentioned

kinetic data as illustrated in Fig. 6.4.


120

100 .-.--
/
I.
/
80 -

I /
I
60—,
I I
I I /
40-

/
20- I
I
A AA /
0 A' I It IttI Ii
400 600 800 1000 1200
Inlet Temperature (K)

A Experimental data (present work)


Experimental data (Depiak & Wierzba, 1999)
- E5 86 kJ/kmol; A5 1.03x10 5
E=46 kJ/kmol; A5=1.03x10 5
E3=46 kJ/kmol; A2.Ox1O5
E=46 kJ/kmol; A1.O3x1O 2

Fig. 6.1 Experimental data and predicted methane conversion as a function of inlet

temperature for-various pre-


exponential factors and activation energies; platinum catalyst,

u= 1.0m/s,=0.35.
121
100

Pt catalt
ys I ,
-

- A Experimental data (present work)


80 -- - - - Calculated data (present work)
- Experimental data (Depiak & Wierzba, 1999) I
0••

• 60

(540

20

0
200 400 600 800 1000
Inlet Temperature (K)

Fig. 6.2 Experimental and calculated methane conversion as a function of inlet

temperature, platinum catalyst, u= 1.0 m/s, 4) = 0.35, E8=46.0 kJ/mol for Tj≥ 750 K, E5

= 42.0 kJ/rnol for Tm< 750 K, A5=1.03x10 3 kmol/m2.K.

100

Pt catalyst /

A Experimental data (present work, 0.35) /


80 Calculated data (present work, 0.35) 1
j
- - - -

Experimental data (Depiak & Wierzba, 0.35)


- . - - - Calculated data (present work; 0.50) . /
/
. 60
-

. /
- !'
-

20 - I]
I'

0
200 400 600 800
Inlet Temperature (K)

Fig. 6.3 Experimental and calculated methane conversion as a function of inlet

temperature for two equivalence ratios, platinum catalyst, u= 1.0 m/s.


122

100
Pt catalyst
- - - -. Calculated data-
(present work)
80 - Experimental data (Depiak& Wierzba, 1999)
,0
- - - - Calculated data (present work)
0'•' 49 Experimental data (Depiak & Wierzba, 1999)

60

40
U = 1.0 rn/s ,

I.
I
U 3.0 rn/s
20

'tj.
0
0 200 4(J 600 800 1000
Inlet temperature (K)
Fig. 6.4 Experimental and calculated methane conversion as a function of inlet

temperature for two approach velocities, platinum catalyst, 4) = 0.35.

6.2.2 C0304/Cr2O3 Catalyst

The same sequence of steps in obtaining the kinetic data has been carried out for methane

conversion over the C0304/Cr2O3 catalyst at u = 1.0 m/s and 4) = 0.35. The agreement

between the experimental methane conversion data and the computed results is shown in

Fig. 6.5. The values of the activation energy and the pre-exponential factor for a

heterogeneous reaction were estimated to be 46.3 ± 0.3 kJ/mol and 2.2x10 3 kmol/(m2.$)

respectively for temperatures greater than 775 K. These have afair agreement with the

experimental data. However, at temperatures less than or equal 775 K, the activation

energy was 42.8 ± 0.8 kJ/mol.

Figure 6.6 illustrates the improved conversion with increasing the equivalence

ratio as a function of inlet temperatures using the kinetic data obtained through this

investigation. For instance, starting from methane conversion of 64.9% at 4) = 0.35,


1.23

methane conversion increased with increasing 4), attaining avalue of 85.11% at 4) = 0.50.

The predicted methane conversion is in reasonable agreement with the experimental, data

for 4) = 0.50. However at temperatures above 800 K, the predicted methane conversion is

as much as 20% higher than the measurements. These observations suggest adeficiency

in the surface chemistry mechanism, which may require a more detailed reaction

mechanism.

The same tendency as that in the case of platinum catalyst toward the dependence

of methane conversion on the approach velocity is observed for the C0304/Cr2O3 catalyst

(Fig. 6.7).

100

C0 364/Cr2O3 Catalyst
80 Experimental data (present work)
- - - Calculated data (present work)
Experimental data (Depiak & Wierzba, 1999) .../
/
.2 60

34o
16

20

in

400 600 800 1000


Inlet Temperature (K)
Fig. 6.5 Experimental and calculated methane conversion as a function of inlet

temperature, C0304/Cr2O3 catalyst, u= 1.0 m/s, 4) = 0.35, B5= 42.8 kJ/mol for ≤ 775

K, B5= 46.3 Id/mo! for Tin > 775 K, A = 2.2 x i0 kmoll m2.K.
124

100

Co 3O/CiO3 Catalyst
80
-
- - - • - - Experimental
Calculated data
data
(present
(present
work;4) 4)=
work;=0.35)
0.35) /

Experimental data (Depiak & Wierzba;4) =0.35)!


/
,
Calcu l
ate ddata (
present work;(j) =0.50)

/
- - . - - -

.9 60 - i Experimental data (present work; 4) =0.50)

40 -

20 -

a
,

200 400 600 800 1000


Inlet Temperature (K)
Fig. 6.6 Experimental and calculated methane conversion as a function of inlet

temperature for two equivalence ratios, C0304/Cr2O3 catalyst, u= 1.0 m/s.

100
/

I Co 3O4ICr2O3 Catalyst /
/
80
-, Experimental data (present work) /
- - - Calculated data (presentwork)
I
- l Experimental data (Depiak & Wietzba) /
Calculated data (present work) /
I

'A
40 00 u3.0m/s

u= 1.0 m/s
20

ii till I
400 600 800 1000 1200
Inlet Temperature (K)
Fig. 6.7 Experimental and calculated methane conversion as a function of inlet

temperature for two approach velocities, C0304/Cr2O3 catalyst, = 0.35.


125

6.3 Results of Model Prediction

In Chapter 3 the effects of the parameters that. influence methane conversion were

presented using kinetic data from the open literature. In this section the effect of some of
these parameters are presented using kinetic data obtained during the present

investigation for homogeneous and surface reactions.

6.3.1 Methane Oxidation within the Inert Bed

Effect of bed porosity. The effect of bed porosity on the methane conversion as afunction

of the inlet temperature is shown in Fig. 6.8. The decrease in the surface area by

increasing the bed porosity decreases the conversion of methane within the inert bed. The

effect of bed porosity is more pronounced between a= 0.2 and a= 0.4 than between a=

0.4 and a= 0.9,

Effect of bed length. The variation of methane conversion with inlet temperature, for

different bed lengths, may be seen in Fig. 6.9. Th6 curves of the conversion of methane

shifted to the left as aresult of increasing the residence time. The inlet temperatures
corresponding to the complete conversion of methane for bed lengths of 50,75, and 100

mm are 1088, 1050, and 1040 K, respectively.

Effect of solid thermal conductivity. Increasing the solid thermal conductivity enhanced

the conversion of methane and allowed complete conversion to take place at alower inlet

temperature (Fig. 6.10). This is due to increased heat conduction to the upstream region
of the bed.
126

100

Inert Bed
80

---"-6=0.4
8 - 0.
5

c40

20

I I
0 .

400 600 800 1000 1200


Inlet Temperature (K)
Fig. 6.8 Calculated methane conversion as afunction of inlet temperature for various bed

porosities within the inert bed, = 0.35, u= 1.0 m/s, k = 1.0 W/ m.K, L = 50 mm.

100

Inert Bed
80
- - - - L=50nun
L75nun
LI00mm

• 60

0
40

20

I I
0
400 600 800 1000 1200
Inlet Temperature (K)
Fig. 69 Calculated methane conversion as afunction of inlet temperature for different

bed lengths; s= 0.4, k= 0.35, u 1.0 m/s, k


= = 1.0 W/ m.K.
127

100

80
= Inert Bed
- - - - k = 0.5 W/(m.K)
k1.0W/(m.K)
- - - - - = 2.0 W/(m.K)

•9 60

40

20

I I
0
400 600 800 1000 1200
Inlet Temperature (IC)
Fig. 6.10 Calculated methane conversion as afunction of inlet temperature for various

solid thermal conductivities; 4= 0.35, u= 1.0 m/s, L= 50 mm, a= 0.4.

6.3.2 Methane Oxidation within the Heterogeneous Bed

Effect of bed porosity. The effect of bed porosities on methane conversion at the exit of

the reactor is shown in Fig. 6.11. The complete conversion of methane was obtained over

the entire range of bed porosities employed (a 0.2 to 0.9) for abed length of L = 50

mm. Increasing the bed porosities results in smaller catalytic surface area which yields

smaller conversion of methane. Thus, light-off temperature was increased significantly as

the bed porosity is increased. The light-off temperatures for bed porosities of a= 0.2, 0.4,

0.5, and 0.9 are 720, 870, 910, and 1000K, respectively.

Effect of bed length. The effect of bed length on the conversion of methane is illustrated

in Fig. 6.12 for abed porosity of a= 0.5. The complete conversion of methane is obtained

for the three bed lengths employed. Over the entire temperature range, the highest

conversion of methane is attained within the bed of 100mm long. This is due to the

increase of the catalytic surface area as well as the residence time for the longer bed.
128

Effect of solid thermal conductivity. The effect of solid thermal conductivity on the

conversion of methane for 1= 0.35, u= 1.0 rn/s within aheterogeneous bed is shown in

Fig. 6.13. As indicated earlier, the effect of thermal conductivity is negligible especially

for methane conversion of above 75%. The increase in the thermal conductivity increased

the conversion of methane in particular at the light-off regime. This trend is due to the

increased surface reactions and increased thermal dissipation at the solid.

6.4 Summary

The predicted performance of the catalytic reactor using the measured kinetic coefficients

of the present investigation revealed similar dependence on equivalence ratio, velocity,

bed porosity, bed length and solid thermal conductivity as when using the kinetic

coefficients obtained from the literature (Chapter 3). However, using the kinetic

coefficients obtained during the present investigation shifted methane conversion curves

toward lower inlet temperatures.

•100
- . • - - -•

Heterogeneous reaction only .

0 80 ------8=0.2 1
q
80.4 /
-----8=0.5 /
•- 60 ! I /
I.

40
- ! /
V.
I ,

20
I , /
-
I / /

I I I I
0 1(111!! i I I

400 600 800 1000 1200


Inlet Temperature (K)
Fig. 6.11 Calculated methane conversion as afunction of inlet temperature for various

bed porosities; 4) = 0.35, u= 1.0 m/s, k = 1.0 W/(m.K), L = 50 mm; Pt catalyst.


129

100

..- 80

•- 60

40

20

0
400 600 800 1000 1200
Inlet Temperature(K)
Fig. 6.12 Calculated methane conversion as afunction of inlet temperature for different

bed lengths; F, = 0. 5, k= 0.35, u= 1.0 m/s, k = 1.0 W/ m. K; Pt catalyst.

100

eaction only
Heterogeneous r

80 1ç0.5W/(m.K)
= 1.0 W/(m.K)

- - - - 1ç2.OW/(m.K) 1

60

I
40 k.=4.0 i m.K))/d

20

900 800 900 1000 1100 1200


Inlet Temperature (K)
Fig. 6.13 Calculated methane conversion as afunction of inlet temperature for various

solid thermal conductivities; = 0.35, u= 1.0 m/s, L = 50 mm, a= 0.4; Pt catalyst.


130

CHAPTER 7

OXIDATION OF OTHER GASEOUS FUELS IN A PACKED BED

REACTOR

7.1 Introduction

One of the important considerations in the selection of acatalyst is its ability to operate

on specific fuels. As mentioned earlier in the literature survey, several investigations have

shown variations in performance with fuel type for aparticular catalyst. For example,

higher molecular weight hydrocarbons such as propane can be burnt over awider range

of conditions than natural gas, and low calorific value gases may be even easier to

oxidize (Kesslering, 1986). The performance of acatalytic combustor is determined by a

large number of operational and physical conditions. Performance can be quantified by

measurements of combustion efficiency, emissions and fuel applicability.

In this chapter, the results of calculations of the oxidation of other common

gaseous fuels such as ethane, propane; carbon monoxide and hydrogen within apacked
bed reactor using the developed model are described. The results of the model predictions

are compared to the available experimental data (Depiak, 2003) obtained under the same

conditions for platinum and abinary mixture of C0304/Cr2O3 catalysts in order to obtain

the kinetic data for these fuels in the packed bed reactor.

The homogeneous (gas-phase) and heterogeneous (surface) reaction rates for the

fuel over the catalytic pellets were modeled as an irreversible one-step reaction rates as

discussed earlier in Chapter (3). They are listed below:


131

h = Ag c;Co" exp [ Eg
RTg1

C
o = AsCf C0. 5 expf --j (7.2)

The kinetic parameters for surface (As, E) and gas-phase (Ag,Ed reactions as well as

power coefficients in and nwere obtained from the literature for lean fuel-air mixtures
with platinum as acatalyst. They are listed in Tables 7.1 and 7.2. Other alternatives of

kinetic data could have been chosen, however these data were chosen because the range

of equivalence ratios and temperatures they correspond to are within the same range of

the present study. The kinetic data adopted from the literature are used in the parametric

study of the fuels oxidation in the packed bed reactor.


Table 7.1 Kinetic parameters for gas-phase reactions.

Mixture Conditions Ag Eg
Fuel m n Reference
(i.e., Flow Type, , Ti,,) (kmol/m3.$) (kJ/mol)

Premixed Present work (matching


CH4 0.2 1.30 6.6 x 10 10 180.0
0.15 <4 <0.50 experimental and predicted data)

Laminar flame model


C2H6 0.1 1.65 6.19 x10 9 125.0 Westbrook and Dryer, 1981
•O.5<<1.9

Laminar flame model; 0.5 < < 1.9


Westbrook and Dryer, 1981
C3H8 Shock tube; 0.12<<2.0 0.1 1.65 4.836x 10 9 125.0
Hautman et al., 1981
Ti
n

Sobolev, 1957
CO 820 <T1 < 2400 1.0 0.0 1.07 x i07 107.1
Aleksandrov and Azatyan, 1973

Premixed
H2 . 1.0 1.0 5.155 x 10 9 125.3 Mon et al., 1977
Tm<1000 •
Table 7.2 Kinetic parameters for surface reactions.

Temperature Range A E
Fuel Reactor Type Reference
(K) (kmollm2 .$) (kJ/mol)

CI- 4 723 <Tin < 853 platinum on alumina fibre 1.03 x i0 86.0 Trimm, and Lam, 1980

C2H6 435 <T1,


1< 770 platinum filament 4.237 x i0 68.85 Haim et al., 1968

435 <T m < 770 platinum filament; Haim et al., 1968;


C3H8 3.478x10 3 70.85
298 <T1 < 1223 platinum foil Song et al., 1991

CO 535 <Tm <760 Platinum catalytic converter 2.2 x10 3 23.56 Kuo et al., 1980

273 <Tm <450 platinum wire; 10.30 Boreskov, 1954;


H2 5.4x104
450 <Tm < 1073 platinum coated flat plate 16.05 Brown et al., 1983
134

7.2 Ethane Combustion

Natural gas composition varies significantly from one source to another. The methane

content of natural gas varies from 70% to 95%, the ethane content varies from 3% to 18%

with higher hydrocarbons, hydrogen, nitrogen and carbon dioxide making up the balance

(Griffin et al., 1989). Since methane is much less reactive than ethane the combustion

characteristics of natural gas depend on the amounts of minor constituents present.

Variations in the composition of natural gas can affect the performance of combustion

devices and may substantially increase -the emissions. It has also been shown by Pfefferle

et al. (1989a and 1989b), that the use of catalytic combustion reduced the dependence of

the system on the natural gas composition. Moreover, it has been shown that the addition

of asmall amount of ethane (less than 20% of the total fuel) to apredominantly methane

fuel enhanced significantly the OH radical production in the gas-phase allowing easier

ignition and flame propagation in their burner (Griffin et al., 1989). The catalytic

oxidation of premixed ethane in air within a packed bed reactor is discussed in this

section. The kinetic data used to obtain the predicted results are listed in Tables 7.1 and

7.2.

The conversion of ethane at different inlet temperatures is shown in Fig. 7.1. For

comparison, the conversion of methane is also shown. It can be seen that the noticeable

conversion within the inert bed starts at the inlet temperature of about 800 K for ethane

and 900 K for methane. The temperatures at which ignition of the fuel takes place are

855 and 1078 K for ethane and methane respectively, which is consistent with the fuels

reactivity. The ignition temperature of the fuel is defined as the temperature beyond

which a considerable amount of the fuel (10% and higher) is oxidized. As expected,

within the catalytic bed, the ignition and complete conversion take place at much lower

temperatures than within the inert bed, i.e. at 628 K for ethane and 745 K for methane.

- The conversion curves of ethane-air mixtures with different equivalence ratios'

and u1.0 m/s are shown in Fig. 7.2 for the inert bed. An increase in the equivalence ratio

towards stoichiometric value enhances the oxidation very significantly as would be

expected.
135

100

Catalytic Bekl CH
80
CH4
Fuel Conversion (%
C2
C2H6
60

40

Inert Bed
20

0 L

600 700 800 900 1000 1100


Inlet Temperature (K)
Fig. 7.1 Calculated fuel conversion as afunction of inlet temperature within an inert (*

symbol) and catalytic packed beds, 4= 0.35 and u= 1.0 m/s.

100

Inert Bed

80

0:15

0.24
.2' 60 0.35
0.50

,. c540
Q .

cJ
20

700 800 900 1000 . 1100


Inlet Temperature (K)

Fig. 7.2 Calculated ethane conversion as a function of inlet temperature within. an inert

bed for different equivalence ratios, u= 1.0 m/s, k5= 1.0 W/m.K, L = 50 mm.
136

For instance, the inlet temperature corresponding to 10% conversion of ethane decreases

from 995 to 677 K when the mixture equivalence ratio increases from 0.15 to 0.50.

Furthermore, aconsiderable decrease in the slope of the conversion curves i noticeable

when the mixture became very lean.

The equivalence ratio also has a similar effect on ethane oxidation within a

catalytic .bed (Fig. 7.3). However, the conversion of ethane starts at much lower inlet

temperatures than in case of the inert bed for the same conditions. For example, the

conversion of 100% ethane in mixtures of equivalence ratio of 0.50, 0.35, 0.24, and 0.15

is obtained at inlet temperatures of 470, 629, 760, and 845 K, respectively.

The gas and solid temperatures along the bed for these equivalence ratios at the

inlet temperature corresponding to the complete oxidation of ethane at the exit of the

reactor are plotted in Fig. 7.4. As the equivalence ratio of the mixtures decreases from

0.50 to 0.15 the exhaust temperature was reduced from 1711 to 1226 K respectively.

Moreover, the maximum gas temperatures corresponding to 84.4% ethane conversion

were decreased from 1816 to 1200 for = 0.50 and 0.15 respectively. It is also worth

emphasiing that the maximum value in the gas temperatures for mixtures of 4= 0.50,

0.35 and 0.24 are generated by the thermal energy released by ethane. This maximum

value in the gas temperatures diminishes as the mixture reaches the equilibrium state due

to heat exchange between the gas and solid phases by convection. Furthermore, the

maximum difference in the temperature between the gas and solid diminishes as

decreaes from 0.50 to 0.24. The thermal energy released for 4, = 0.15 was very small.

Thus, there was no peak value observed in the gas temperature under such conditions. It

appears that the thermal energy released by the mixture is reduced significantly with

decreasing the equivalence ratio in the mixture. This is rather expected. It is also noted

that the maximum temperature along the reactor is shifted from the left to the right as 4,

increases from 0.15 to 0.50 as shown in Fig. 7.4. This trend is expected because for the

mixture of 4, = 0.15 the inlet temperature is relatively high in comparison to the mixture

of 4, = 0.50; therefore, oxidation of ethane starts at an earlier location in the bed and

continues to increase gradually due to the low energy release from the fuel.
137

100

Catalytic Bed
80

. 60

0
Q40

QY
/ 4)
0.15
20- ' ! 0.24
1 0.35
- ! •.' -. 0.50
I ...... .
_______. . .

I I I

500 600 700 800 900 1000


Inlet Temperature (K)
Fig. 7.3 Calculated ethane conversion as a function of inlet temperature for different

equivalence ratios, u= 1.0 m/s, k = 1.0 W/m.K, L = 50 mm; Pt catalyst.

Fig. 7.4 Calculated temperature distribution for the gas (solid line) and the solid (broken

1800
Catalytic Bed 4=0.50
1600

1400

1200

1000

800

0.01 0.02 0.03 0.04 0.0


Distance along the bed (m)
line) along the bed for various equivalence ratios of ethane, u= 1.0 m/s; Pt catalyst.
138

The effect of the flow velocity is shown in Fig. 7.5. It can be seen that an increase in the

flow velocity reduces the residence time and that this requires an increase in the inlet
temperature for the complete conversion of the fuel within the bed. The catalytic bed

improves ethane conversion significantly and allows it to take place at much lower inlet

temperatures. For example, the complete conversion of ethane for u= 3.0 m/s is obtained

at the inlet temperature of 850 K within the catalytic bed while the corresponding
temperature in the case of the inert bed was - 1080 K.

100 T

- Catalytic Bed

80 - i---.— u'1.O&s *
- u2.Om/s I
! u =3 .
0 m 1s

- i—.—.— u1.Om/s I
o 60 u2.Om/s I
I
-

- ! 3.
0 m/s

• :'
1
I
I
0
L) 40 -.

- I I

ii

20-. /
Inert Bed I

o il
600 700 800 900 1000
'I
1100
Inlet Temperature (K)
Fig. 7.5 Calculated ethane conversion as a function of inlet temperature within the

catalytic (thick lines) and inert (*, thin lines) beds for different approach velocities,

=0.35.
139

7.3 Propane Combustion

Catalytic oxidation of homogeneous mixtures of propane-air within the catalytic and inert

packed bed reactors are discussed in this section. The kinetic data used to obttain the

results of predictions for the heterogeneous and gas-phase reactions are listed in Tables
7.1 and 7.2.

The results of the calculations of the oxidation of homogeneous propane-air mixture

within the inert bed are shown in Fig. 7.6 at u= 1.0 m/s for various equivalence ratios.

-The trends observed are similar to those for other fuels (i.e., methane and ethane). As the

equivalence ratio of these lean mixtures increases the conversion of the propane is

completed at lower temperatures. For example, the complete conversion of propane in the

mixture with equivalence ratio of 0.50 took place at the inlet temperature of 767 K while
the temperature of 1150 K was needed for the complete conversion of propane in the

mixture with the equivalence ratio of 0.15. The conversion of propane was very fast with

an increase in the inlet temperature for the mixture of 4= 0.50 as it reaches conditions
corresponding to its lean flammability limit in air.

In the catalytic bed, propane is oxidized completely at amuch lower inlet temperature.
-

than in case of the inert bed for the same conditions. Moreover, the effects of the inlet
temperature and mixture equivalence ratio are also significant for the mixtures passing

through the catalytic bed (Fig. 7.7). An-increase in the equivalence ratio lowers the inlet

temperature at which complete oxidation of C3H8 occurs. For example, the complete

- conversion of the propane in mixtures with equivalence ratios of 0.15 and 0.50 is

obtained at the inlet temperatures of 900 and 540 K respectively. A similar trend was

observed by Bruno et al. (1983) experimentally and numerically for propane in a

platinum monolithic reactor for arange of equivalence ratios from 0.19 to 0.32.
140

100

InertBed

80 4)
0.15
0.24
- - - - - 0.35

. 60

c40

20

I I , I I , I p

900 800 900 1000 1100 1200


Inlet Temperature (K)
Fig. 7.6 Calculated propane conversion as afunction of inlet temperature within the inert

bed at different equivalence ratios, u= 1m/s, k = 1.0 W/m.K, L = 50 mm.

100

Catalytic Bed i

4):
-------0.15 I
0.24
0.35
- ---0.50

20

400 500 600 700 800 900 1000


Inlet Temperature (K)
Fig. 7.7 Calculated propane conversion as a function of inlet temperature at different

equivalence ratios, u= 1m/s, k = 1.0 W/m.K, L = 50 mm; Pt catalyst.


141

The calculated gas and solid temperatures distribution curves along the packed bed for
the complete oxidation of propane-air mixtures of equivalence ratios ranging from 0.50 to

0.15 are shown in Fig. 7.8. The trends of these curves demonstrated similar behavior to

those of ethane. However, two differences from ethane are observed in these curves.

First, the difference between the gas temperatures and solid temperatures in the region

corresponding to the peak in the gas temperature is smaller for propane-air mixtures than

for ethane-air mixtures for 4) = 0.50, 0.35, and 0.24. Second, for propane-air mixture the

temperatures of the gas and solid are higher than for ethane-air mixtures at all the

equivalence ratios examined. These higher temperatures are attributable to the higher

inlet tempethtures for complete oxidation in the case of propane-air mixtures. Such

behavior is consistent with the fuel reactivity.

Figure 7.9 shows the results of the oxidation of lean propane-air mixtures in inert and

catalytic packed beds at different approach velocities and 4) =0.35. Propane conversion

exhibits adependence on the approach velocity similar to that of the ethane-air mixture

with either bed. An increase in the approach velocity increased the inlet temperature for

the complete conversion of propane within the catalytic and inert beds (Fig. 7.9).

Figure 7.10 represents the conversion of methane, ethane and propane within the catalytic

and inert beds as afunction of the inlet temperatures for the flow condition of 4) = 0.35

and u = 1.0 m/s. Methane's complete conversion takes place at much higher inlet

temperatures than those required for the complete conversion of ethane or propane within

either an .inert or catalytic bed. This is consistent with the fuel reactivity which is higher

for ethane and lower for methane.

For instance for a fuel-air mixture of 4) = 0.35 within the inert bed, the complete

conversion of ethane is obtained at about -860 K while it requires inlet temperatures of


about -940 and 1088 K to achieve the full conversion of propane" and methane
,

respectively. This is expected due to the lower reactivity of methane in air compared with

that of ethane or propane. It should be also noted, that when the mixture is passing
142

91400
L
I
I

Ll 1200

1000

800

0 0.01 0.02 0.03 0.04 0.0


Distance along the bed (m)
Fig. 7.8 Calculated temperature distribution for the gas (solid line) and the solid (broken
line) along the bed for various equivalence ratios of propane, u= 1.0 m/s; Pt catalyst.

100

80

. 60

C_) 40

20

700 800 900 1000 1100 1200


Inlet Temperature (K)
Fig. 7.9 Calculated propane conversion as a function of inlet temperature within the

catalytic (thick lines) and inert (*, thin lines) beds for different approach velocities,

=0.35.
143

through the inert bed, the effect of the mixture equivalence ratio is stronger for the

ethane-air and propane-air mixtures, than for the methane-air mixtures. Since only gas-

phase reactions are considered, the fuel r


eactivity mostly influences the conversion.

In the case of the catalytic bed, the corresponding temperatures for complete fuel

conversion are much lower as mentioned above (i.e. 628 Kfor ethane, 696 K for propane

and 745 K for methane). It can also be seen that in this case (catalytic bed) areverse trend

is observed, i.e. the effect of the mixture equivalence ratio is significantly stronger for the

methane-air mixtures than the ethane-air or propane-air mixtures. Such behavior is in

agreement with the experimental data of Griffin et al. (1992) for equivalence ratios

ranging from 0.03 to 0.5 and plate temperatures up to 1405 K. They found experimentally

that at atmospheric pressure, the platinum surface enhanced methane combustion to a

greater extent than that of ethane. It appears that the surface production of radicals

becomes more important to the ignition of methane.

100
CH4 *

C2H6 !
C3H8 1
80
CH4
I----- C2H.
C3H8
.9 60

40

20
)

600 700 800 900 1000 1100


Inlet Temperature (K)

Fig. 7.10 Calculated fuel conversion as afunction of inlet temperature within catalytic

(thick lines) and inert (*, thin lines) beds, = 0.35 and u= 1.0 m/s.
144

7.4 Carbon Monoxide Combustion

The oxidation of humid CO over platinum catalysts has been studied for many years, but

the conclusions regarding the mechanisms and rate equations are somewhat conflicting.

Tables 7.1 and 7.2 show the pre-exponential factor and the activation energy used to
obtain the predicted results in this section for catalytic and inert beds at atmospheric

pressure. The obtaining of asimple rate equation an aim of this investigation is of much

importance.

Figure 7.11 shows the results of the conversion within the inert bed for carbon
monoxide-air mixtures passing through the reactor as afunction of inlet temperatures

with u= 1.0 m/s and different equivalence ratios. An increase in the mixture equivalence

ratio significantly improved the firel conversion and allowed complete conversion at

much lower temperatures. For very lean mixtures of = 0.20 and 0.15 and at

temperatures below 1060 K, the effect of equivalence ratio appears negligible.

However, carbon monoxide conversion exhibited different behavior within the


catalytic bed than in the inert bed in its dependence on the equivalence ratio and inlet

temperatures. As illustrated in Fig. 7.12, the effect of the equivalence ratio is more

apparent at low temperatures. Whilst, the effect of the equivalence ratio diminishes as the

oxidation of the fuel reaches completion (i.e. exceeds 95%) carbon monoxide conversion

and for inlet temperatures of more than 800 K.

For the maximum conversion (- 97.5 %) of the carbon monoxide at the exit of the
50 mm long bed, the gas and solid temperature distributions for each of these equivalence

ratios is plotted in Fig. 7.13 along the catalytic bed. Similarly to the other fuels, the

increase in the equivalence, ratio increases the exhaust temperature and shifts the curves

toward the exit of the reactor. Since the energy release per unit volume of the carbon

monoxide mixtures is smaller than that for the hydrocarbon fuel, the peak in the gas

temperature distribution is not observed. The difference in the inlet temperature for all the

equivalence ratios tested is within 50 K. For instance, the inlet temperatures at = 0.35

and 0.15 are 800 and 850 K.


145

100

1
Inert Bed

80
-- - - - 0.15 II
- - -. - 0.20
0.35
60
.-

40 I ,
I,
C
U I,

20 '7 .

0
700 .800 900 1000 1100 1200 1300
Inlet Temperaturre (K)
Fig. 7.11 Calculated carbon monoxide oxidation as afunction of inlet temperature within
the inert bed for various equivalence ratios, u= 1.0 m/s.
CO Conversion (%)

400 500 . 600 700 800


Inlet Temperature (K)
Fig. 7.12 Calculated carbon monoxide oxidation as afunction of inlet temperattire for

various equivalence ratios, u= 1.0 m/s; Pt catalyst.


146

The effect of the approach velocity on the conversion of carbon monoxide within the

inert bed is similar to that of the other fuels as shown in Fig. 7.14 with stronger effects
between the velocities of u= 1.0 and 2.0 m/s.

A similar trend is observed for the dependence of carbon monoxide conversion on

approach velocities within the catalytic bed (see Fig. 7.15). The effect of the velocity is
stronger as the conversion reaches 80 % and above.

1800

1700

1600

1500

1400

1300

1200

1100

1000

900

800
0 0.01 0.02 0.03 0.04 0.05
Distance along the bed (m)
Fig. 7.13 Temperature distribution for the gas (solid line) and the solid (broken line)

along the bed for various equivalence ratios of carbon monoxide; u= 1.0 m/s; Pt catalyst
147

100

80

•o 60
.-

C
U
20

• 00 900 1000 1100. 1200 1300


Inlet Temperatune (K)
Fig. 7.14 Effect of approach velocity on calculated carbon monoxide conversion as a

function of inlet temperature within the inert bed, = 0.35.

100

80
CO Conveision (%)

60

40

20

600 800 1000 1200 1400


Inlet temperature (K)

Fig. 7.15 Effect of approach velocity on calculated carbon monoxide conversion as a

function of inlet temperature, 4 = 0.35; Pt catalyst.


148

7.5 Hydrogen Combustion

The effect of the equivalence ratio on the conversion of hydrogen is presented in Fig.

7.16 within the inert bed with u= 1.0 m/s. Unlike the other fuels, methane, ethane, and

propane, the slopes of the hydrogen conversion curves are less steep even for the

equivalence ratio of 0.35. As the mixture become leaner, the slope of these conversion

curves becomes smaller. However, all the fuels employed in this investigation exhibited

similar behavior in their dependence on the mixture equivalence ratio. That is, the higher

the value of the fuel concentration, the higher is the rate of conversion of these fuels as

expected based onEqs. (7.1) and (7.2).

Hydrogen oxidation in the catalytic packed bed is presented in Fig. 7.17 for

various equivalence ratios and u= 1.0 m/s. A considerable amount of hydrogen ('-.25 %)

was oxidized already at room temperature 293 K for the mixture of 4i 0.35 with the

complete oxidation of hydrogen obtained at the inlet temperature of 420 K. A comparison

between the performance of the catalytic and inert beds is presented in Fig. 7.18 for
hydrogen-air mixtures of different equivalence ratios and u 1.0 m/s. The following

conclusions can be drawn from Fig. 7.18:

• The complete conversion of hydrogen within the catalytic bed was obtained at a

much lower temperature than that within the inert bed.

• The effect of the equivalence ratio on the hydrogen conversion within the inert

bed becomes significant at inlet temperatures above 760 K. When ignition started

at temperatures 770 and 800 K for the mixtures of equivalence ratios of 0.35

and 0.05 respectively, the reaction rate is very fast for 4= 0.35 and much slower

for the very lean mixture of 4= 0.05.

• The effect of the equivalence ratio on hydrogen conversion within the catalytic

bed is much stronger than that within the inert bed. Pfefferle and Pfefferle (1987)

have experimentally observed similar behavior for hydrogen mixtures of

equivalence ratio less than 0.3. They have explained that such promotion is due to

the very efficient ability of platinum to dissociate molecular oxygen.


149

100
,
,
Inert Bed
/
80 l) /
0.05
0.10 /
- - - - 0.15
- - - - - 0.35
60

40
/

20

700 800 900 1000


Inlet Temperature (K)
Fig. 7.16 Calculated hydrogen conversion as afhnction of inlet temperature at different

equivalence ratios within the inert bed, u= 1.0 m/s.

100

/
80
/
-

/
/
60
I
I
/
40
I
/ Catalytic Bed
/
0.05
20
/ 0.10

I 0.15

I 0 i

300 400 500 600 700 800 900


Inlet Temperature (K)
Fig. 7.17 Calculated hydrogen conversion as afunction of inlet temperature at different

equivalence ratios, u= 1.0 m/s; Pt catalyst.


150

The temperature distribution along the catalytic bed for the complete conversion of

hydrogen in air is shown in Fig. 7.19 for various equivalence ratios and u= 1.0 m/s. The

temperature increases as the equivalence ratio increases which is rather expected.

Moreover, several differences were observed between hydrogen mixtures and the other

fuels considered in this investigation. A gradual increase in the temperature of the gas

and solid along the bed is observed. Furthermore, there is no peak in the gas temperature

curve after ignition observed for hydrogen. This can be attributed to the fact that the

energy release by hydrogen, on avolume basis of the mixture, is smaller. The influence

of ahot catalytic surface on the ignition of gas-phase combustion in fuel/air mixtures has

been shown to be a strong function of mixture stoichiometry (Pfefferle and Pfefferle,

1987). Besides, Schefer and Robben (1980) found that for ethane, propane and hydrogen

ignition the catalytic inhibition effect becomes lower as the mixtures were made leaner.

Figure 7.20 shows the temperature distribution of the solid and gas for all the fuels (i.e.

CH4,C2H6,C3H8,CO, and H2)employed in this investigation within the catalytic bed at 4


= 0.35 and u = 1.0 m/s. The complete conversion of the fuel is achieved at the lowest

inlet temperature of 420 K for the hydrogen-air mixture. Furthermore, the lowest exhaust

gas temperature is obtained for hydrogen fuel because it has the lowest inlet temperature

required for the complete conversion. This is consistent with its high reactivity. Also, it is

worth to mention that the maximum gas temperature for the fuels employed in the present

investigation is below 1700 K and hence thermal NO emission is insignificant.

The dependence of hydrogen conversion on the approach velocity within the inert bed is

similar to the other fuels. Figures 7.21 and 7.22 illustrate this dependence for two

different equivalence ratios of hydrogen-air mixture ( = 0.10 and 0.35). The lean

flammability limit for hydrogen in air is of = 0.15 at 400 °C for horizontal gas phase

propagation (Zabetakis, 1965). Thus, the mixture with an equivalence ratio of 0.35 is

within the flammable region while the mixture of k 0.10 is outside.


151

100
-
I.
-.—.,
,. II

80 - 'I Catalytic Bed / I


- I

I !
Inert Bed
40 0.05

:!
0.10
0.15
I/I 0.35
20 I
/1 0.05 •
0.10
I 0.15 •
/ 0.35

0
400 600 800 1000
Inlet Temperature (K)
Fig. 7.18 Comparison between calculated hydrogen conversion within the catalytic and

inert beds (*; thin lines) versus inlet temperature for various equivalence ratios, u=1.0

m/s.

1200

1100

800

700

600

500

I I i

400 o 0.01 0.02 0.03 0.04 0.0


Distance along the bed (m)
Fig. 7.19 Calculated temperature distribution for the gas (solid line) and the solid (broken

line) along the bed for various equivalence ratios of hydrogen; u= 1.0 m/s; Pt catalyst.
152

1800
Catalytic Bed

1600

g 1400

1200

Fi000

800

600

400 o
0.01 0.02 0.03 0.04 0.05
Distance along the bed (m)
Fig. 7.20 Calculated temperature distribution for the gas (solid line) and the solid (broken

line) for CH4,C2H6,C3H8, CO, and H2 along the bed, = 0.35, u= 1.0 m/s; Pt catalyst.

100
/
IneitBed
/
80 ------u1.Om/s I
u2.Om/s I
....... u=3.Om/s /

ç 40

20
/
/

700 800 900 1000


Inlet Temperature (K)
Fig. 7.21 Calculated hydrogen conversion as afunction of inlet temperature within the

inert bed for different approach velocities, = 0.10.


153

It is clear from Figs. 7.21 and 7.22 that for both mixtures the effect of the approach

velocity on hydrogen conversion is more significant at higher inlet temperatures.

Complete conversion of hydrogen at u = 1.0 m/s is achieved at the inlet temperatures of

820 K and 980 K for 4) = 0.35 and 0.10 respectively. Inlet temperatures of 875 K and

1050 K are needed to obtain the complete conversion of hydrogen at u= 2.0 m/s. For the

complete conversion of hydrogen with velocities of 1.0 and 2.0 m/s, an inlet temperature

difference of 55 K and 130 Kis obtained for the mixtures 4) 0.35 and 0. 10, respectively.

The effect of approach velocity on hydrogen conversion within the catalytic bed

(Fig. 7.23) is more pronounced than that within the inert, bed (Fig. 7.22). Figure 7.24

shows a comparison of hydrogen conversion between the catalytic and inert beds for

various approach vekocities at 4) = 0.35. The effect of the velocity within the catalytic bed

is significant especially for hydrogen conversion less than 40%. Moreover, it is

noticeable that at the velocity of 3.0 m/s within the catalytic bed a much higher inlet

temperature is required to achieve the complete conversion of hydrogen than when the

velocity is 1.0 rn/s.

Inert Bed
I

- - - - u1.OmJs
u2.Om/s
u3.Om/s

I
I
I
I
/
I
I

50 700 750 800 850 900


Inlet Temperature (K)
Fig. 7.22 Calculated hydrogen conversion as afunction of inlet temperature within the

inert bed for different approach velocities, 4) 0.35.


154

100

80

c 40

20

0
300 400 500 600 700 800
Inlet Temperature (K)
Fig. 7.23 Calculated hydrogen conversion as afunction of inlet temperature for different

approach velocities, 4) = 0.35; Pt catalyst.

100 ,

Ij
Ii
I I
80
i Catalytic Bed/
i I
Ii
I

60
I I
I I I I
I.

- I i/
1 I
I

I
(
I
40
I. InertBed
, I! - - - - u1.Om/s
uQ.Om/s
u3.Om/s
20 u1.0m1s
I u-2.0 m/s'
- / 1
• u=3.0 m/s *

I 1 !IIII I II I- 111 I I I I
0 300 400 500 600 700 800 900 1000 1100
Inlet Temperature (K)
Fig. 7.24 Comparison between calculated hydrogen conversion within the catalytic and

inert beds (*; thin lines) versus inlet temperature for various approach velocities, 4) =

0.35.
155

CHAPTER 8

COMPARISON OF MODEL PREDICTION WITH THE

EXPERIMENTAL DATA

8.1 Introduction

One of the aims of this investigation is to estimate more accurately the rate expression for

catalytic oxidation within apacked bed reactor. Based on experimental data obtained by

Depiak (2003) with two different catalysts, the kinetic data (i.e. the activation energy and

the pre-exponential factor) for the overall rate expression within the packed bed reactors

were determined using the same procedure discussed in section 6.2 of Chapter 6. The

gas-phase kinetic coefficients for C3H8, CO, and H2 are from the literature and are listed

in Table 7.1.

8.2 Determination of Kinetic Data for Surface Reactions of Common

Gaseous Fuels

Propane-air mixture: Fig. 8.1 shows the experimental and predicted results of propane

conversion within the platinum bed of 4= 0.35 and u= 1.0 m/s: For the platinum packed

bed, the corresponding value of the activation energy is 27.7 ± 1.1 kJ/mol and the pre-

exponential factor is 1.70 x10 -3 kmoll(rn2.


s).

However for the Co 3O4ICr2O3 packed bed these values, as can be seen in Fig. 8.2, are:

Activation energy (Es)of 25.10 ± 1.9 kJ/mol at inlet temperatures 700 K;

Activation energy (Es) of 29.87 ±0.9 k3/mol at inlet temperatures > 700 K;

with the pre-exponential factor (As)of 1.70 x i0 kmoll(m2.$).


156

100 • ..41- -

80
0
I
/
/
/

/
I
/
/

Pt Catalyst
20
- - - Calculated Data (present work)
• Experimental Data (Depiak, 2003)

400 500 600


I 700 800 900 1000
Inlet Temperature (K)
Fig. 8.1 Calculated and experimental propane conversion as a function of inlet

temperature within platinum bed, t1 = 0.35, u= 1.0 m/s.

100

80

° 60
$ 1

0
ç..) 40

C0 30./Cr2O3 catalyst
20 1'
- - - - Calculated Data (present work)
• Experimental Data (Depiak, 2003)

04 o8 I I I I I I
500 600 700 800
+I

900 1000
Inlet temperature (K)
Fig. 8.2 Calculated and experimental propane conversion as a function of inlet

temperature within C0304/Cr2O3 bed, 1= 0.35, u = 1.0 m/s.


157

Carbon monoxide-atmospheric air mixture: for the temperature range tested there seem

to be two activation energies for the surface reaction of carbon monoxide over aplatinum

catalyst. The following activation energies provided and an adequate representation of the

experimental data of Depiak (2003), as can be seen in Fig. 8.3.

Activation energy (Es) of 25.56 ± 1.7 kJ/rnol at inlet temperatures < 650 K;

Activation energy (E s)of 20.56 ±2.0 kJ/mol at inle


ttemperatures> 650 K.

Thd pre-exponential factor (As) for the above activation energies is 2.2 x 10-3 kmol/(m2.$).

Furthermore, the Co 3O4ICr2O3 catalyst yields an activation energy (Es) of 27.0 ±

1.0 kJ/mol at inlet temperatures 650 K and avalue for E of 23.0 ± 1.3 kJ/mol at inlet

temperatures> 650 K. The pre-exponential factor for the ço3O4JCr2O3 packed bed is 2.4

x i0 kmoll(m2.$) (Fig. 8.4).


158

100

80
CO Conversion (%)
I,

60
/
/
/
40
/
Pt Catalyst
/
/
20
- - - Calculated Data (Present work)
Experimental Data (Depiak, 2003)

0
400 500 600 700 800 900
Inlet Tem perature (K)
Fig. 8.3 Calculated and experimental carbon monoxide conversion as afunction of inlet

temperature within platinum bed, 4) = 0.35, u = 1.0 m/s.

100

80 /

Co 3O4ICiO3 Catalyst
20
- - - - Calculated Data (Present work)
Experimental Data (Depiak, 2003)

I I
500 600 700 800 900
Inlet Temperature (K)
Fig. 8.4 Calculated and experimental carbon monoxide conversion as afunction of inlet

temperature within C0304/Cr2O3 bed, 4) = 0.35, u= 1.0 m/s.


159

Hydrogen-air mixture: kinetic data for hydrogen within the platinum packed bed is also

obtained as shown in Fig. 8.5. At inlet temperature 350 K, the activation energy (E) is

10.0 ± 0.5 kJ/mol. An activation energy is 9.4 ± 0.6 kJ/mol at inlet temperature >350 K.

The pre-exponential factor for this mixture is 5.4 x 10-5 kmoL'(m2.$). The activation

energy determined in this investigation is close to the value reported by Boreskov (1954),

10 kJ/mol, for his high activity platinum wire in excess of oxygen. Moreover, Hanson

and Boudart (1978) reported avalue for the activation energy of 7.5 kJ/mol in their study

of hydrogen conversion to water (H20) over platinum supported on silica at temperatures

from 273 to 373 K at atmospheric pressure. Furthermore, avalue of 16.1 kJ/mol was

reported by Brown et al. (1983): Their reaction is in the temperature range from 450to

1074K and equivalence ratios from 0.05 to 0.20 for hydrogen on aplatinum catalyst

within the laminar heated boundary layer. Also, an earlier study of Gidaspow and

Ellington (1964) noted arather abrupt change in the mechanism at temperatures near 450

K. This change was expressed as a decrease in activation energies for temperatures

greater than 450 K.

Applying the same strategy to hydrogen oxidation over the C0304/Cr2O3 catalyst

will yield an activation energy of 10.3 ± 1.0 kJ/mol (Fig. 8.6).

The kinetic data for the common gaseous fuels in air mixtures employed in this

investigation over the platinum and C0304/Cr2O3 catalysts are summarized in Table 8.1.
160

100

80 -

40-

20-
r Pt catalyst

0 Experimental Data (Depiak, 2003)


- - - Calculated Data (present work)

9oo 300 400 , 500 600 700


Inlet Temperature (K)
Fig. 8.5 Calculated. and experimental hydrogen conversion as a function of inlet

temperature within platinum bed, 4) = 0.35, u= 1.0 m/s.

,100

a/

60

40

C0 304/Cr2O3 Catalyst
20
Experimental Data (Depiak, 2003)
- - - - Calculated Data (present work)

I I
00 300 400 500 600 700
Inlet Ternpemtun (K)
Fig. 8.6 Calculated and experimental hydrogen ,conversion as a function of inlet

temperature within C0304/Cr2O3 bed, 4) = 0.35, u= 1.0 m/s.


Table 8.1 Kinetic data for platinum and C0304/Cr2O3 catalysts within the packed bed reactor for all the fuels employed in this

study.

Platinum Catalyst C0 3 0 4 /Cr 2 O 3 catalyst Temperature Range


Fuel Type
(To
A (kmoLf(m2.K)) E (kJfmol) A (kmol/(m2.K)) E (kJ/mol)

42.0 42.8 775.0


CH4 1.03 x iO 3 2.2x1c13
46.0 46.3 >775.0

25.1 700.0
C3H8 1.70 x iO 3 27.7 1.7x i0
29.8 >700.0

25.56 27.0 650.0


CO 2.2x10 3 2.4x iO
20.56 23.0 > 650:0

10.0 350.0
H2 5.4x i0 5.4x iO 10.3
9.4 >350.0
162

8.3 Summary

A comparison between the. predicted results using the kinetic data obtained during the

present investigation with the predicted results using the kinetic data from the literature is

presented in Figs. 8.7 to 8.9 when platinum is the catalyst.

Propane-air mixture: the kinetic data used for the parametric study of propane reaction

mechanism in section 7.3 when compared to the kinetic data reported in the present study

as shown in Fig. 8.7 for 4= 0.35 and u= 1.0 m/s show significant differences. It appears
that the value of the activation energy of 27.7 ± 1.1 kJ/mol is far apart from that reported

by Song et al. of 70.85 kJ/mol for lean propane-air mixtures -over aplatinum flat plate.

Such large discrepancy in the values of the activation energy may be due to the different

reactor geometry which can have alarge influence on the values of the activation energy

as was indicated in the literature review. In addition, the effective activation energy is

dependent on the nature of the platinum surface (i.e., dispersed or particulate platinum)

which is not clearly stated in the various experimental results. The extent of fuel

conversion results is structure-sensitive to the platinum surfaces as shown by Otto (1989).

However, aan activation energy of 29.69 kJ/mol was


' reported by Kuo et al. (1980) for

fast oxidizing hydrocarbon group which includes prOpane over platinum catalytic

converter (packed spherical particles). This value of the 'activation energy, is only ,6.7 %

higher than the value obtained in the present investigation. In any case, it is ,to b

remembered that the concept of assuming that the multi-reactions on the catalytic,, surface

can be represented by asingle overall reaction step is highly approximate and c'an lead to

substantial differences in the values of the results.

Carbon monoxide-atmospheric air mixture: acomparison is presented (Fig. 8.8) of the

predicted carbon monoxide conversion using the kinetic data obtained here with the

predicted results from the literature (Table 7.2) over a platinum catalyst. Activation

energies of 25.56 ± 1.7 kl/mol and 20.56 ±2.0 kJ/mol for temperatures above and below

650 K, respectively are obtained which compare to that of 23.56 kJ/mol of Kuo et al.
163

(1980). The calculated carbon monoxide conversion when using the kinetic data of Kuo

et al. is over-predicted for the temperature range between 450 to 650 K, while

temperatures higher than 650 K, is under predicted. Such behavior is also mainly due to

the use of single activation energy over the entire range of temperature (i.e., 450 - 850 K)

while more than one activation energy is needed. Thus, atwo temperature regime (above

and below 650 K) for the activation energy is employed to better fit the experimental

data. The two regime approach is for example also consistent with the data of Song et al.

(1990). At low temperature regime, the difference in the activation energy between the

values of the present study and that of Kuo et al. is 13.0%, and at the higher temperature

regime is 8.0%.

Hydrogen-air mixture: the activation energy obtained during the present investigation is

10.0 ± 0.5 kJ/mol for temperatures 350 K and 9.4 ± 0.6 kl/mol for temperatures> 350

K. There are many low values of the activation energy which is consistent with the highly
reactive behavior of hydrogen. Calculated hydrogen conversion is compared to that of

Boreskov (1954) in Fig. 8.9 and show adeviation in the activation energy value from his

values of 2.9% for low temperature regime and 6.0% for the high temperature regime.

It can be concluded that using single activation energy for awide range of temperature

may lead to significant error in the fuel conversion. Therefore, it is recommended the use

of at least two or may be more than two activation energies for the entire range of

temperature required to achieve the complete conversion of the fuel.


164

100

Pt Catalyst
= - - - - Calculated Data (present work)
Calculated Data (Kinetic data of Haim et al.)
- Experimental Data (Depiak, 2003) I
I

.2 60 I

41
I

o
0 /
40 /

e. I

/
20

100

80

/
o 60
/

/
/
C I.
0 Pt Catalyst
/
20 Calculated Data (Kinetic data, present work)
Calculated Data (Kinetic data; Kuo et al.)
Experimental Data (Depiak, 2003)

I I I
400 500 600 700 800 90
Inlet Temperature (K)
Fig. 8.8 Comparison of calculated carbon monoxide conversion as afunction of inlet

temperature within platinum bed, 4= 0.35, u= 1.0 m/s.


165

60

540

Pt catalyst

- - - - - Calculated Data (KInetic data; present work)


Calculated Data (Kinetic data of Boreskov)
5 Experimental Data (Dcpiak, 2003)

II I I III I I II I i I
300 400 500 600 700
Inlet Temperature (K)
Fig. 8.9 Comparison of calculated hydrogen conversion as afunction of inlet temperature

within platinum bed, 4= 0.35, u = 1.0 m/s.


166

CHAPTER 9

CONCLUSIONS AND RECOMMENDATIONS

9.1 Summary and Conclusions

The oxidation of lean fuel-air mixtures within a packed bed .reactor was investigated

experimentally and numerically. An unsteady plug-flow mathematical model of fuel

conversion in lean premixed and preheated fuel-air mixtures in apacked bed reactor with

inert or platinum catalyst deposited on y-alumina pellets was developed. The governing

conservation equations of mass, chemical species, and energy for both the gas and the

solid and the corresponding boundary conditions were solved using afinite-difference

procedure. The reaction rates, for the gas-phase (homogeneous) and catalytic surface

(heterogeneous) of the Arrhenius type modeled as single-step reactions, were

incorporated to determine the conversion of fuel as afunction of inlet temperatures. The

model allows for considering the role of individual homogeneous and heterogeneous

reactions as well as coupled homogeneous-heterogeneous reactions in the fuel oxidation.

Even though, the model was based on aone-step reaction, it was adequate enough to
predict the features of the catalytic packed bed reactor. The predicted results obtained

were compared with experimental values. Accordingly, aparametric investigation was

performed. Also, effects of changes in the bed porosity, solid thermal conductivity, and

bed length on the conversion of lean methane-air mixtures were investigated.

Experiments were performed on two types of packed beds: inert and catalytic.

The fuel conversion was obtained for each catalytic bed tested namely platinum and a

binary mixture of cobalt and chromium oxides. The range of the operational parameters

covered in the experiments was between approach velocities of 0.75 to 3.0 m/s,

equivalence ratios of 0.2 to 0.6, and inlet temperatures of 293 to 1300 K. The kinetic data

for the oxidation of methane, propane, carbon monoxide, and hydrogen within the packed
167

bed reactor were obtained for platinum and the binary mixture of cobalt and chromium
oxide catalysts.

The major conclusions of this study for all the fuels employed within the inert and

catalytic packed beds are listed below:

, The complete conversion of the fuel, as expected, increased with increasing

inlet temperatures and equivalence ratios or decreasing approach velocities

regardless of fuel type.

• The conversion of the fuel is most sensitive to inlet temperature among the

three inlet parameters (Ti,,,f, and u). It was found that, aone percent change in

inlet temperature can cause asignificant change in overall conversion.

. Fuel conversion within catalytic beds takes place at much lower temperatures

than it does in inert beds for the same equivalence ratio and approach velocity

for all the fuels considered in this investigation.

• For only the heterogeneous reaction case, methane oxidation improved with the

decrease in the bed porosity and or increase in the bed length.

• The effect of solid thermal conductivity (k = 1.0 - 4.0 W/m.K) on methane

conversion was negligible for the only heterogeneous reaction case.

• For the coupled homogeneous-heterogeneous reactions case, an increase in the

bed porosity enhanced methane oxidation and allowed complete conversion to

take place at an earlier location in the reactor.

• A bed length of L = 50 mm was sufficient for completing the oxidation of


methane for a bed porosity of 0.4 or less when the coupled homogeneous-

heterogeneous reaction was considered for k 0.35


= and u= 1.0 m/s.

• An increase in the thermal conductivity of the solid improved the oxidation

process and allowed complete conversion to take place at an earlier location

within the reactor for the case of coupled homogeneous-heterogeneous reaction.

• It is important to consider the coupled homogeneous-heterogeneous reactions

when modeling the fuel oxidation within the packed bed reactor.
168

• The activation energies and pre-exponential factors were obtained for all the

fuel-air mixtures oxidized over the platinum and C0304/Cr2O3 catalysts in this
investigation. The values of these kinetic data are summarized in Table 7.3.

• Methane oxidation exhibited similar behavior with the metal oxides and

platinum catalysts. Increases in the equivalence ratio as well as in the inlet

temperatures or decreases in mixture approach velocity improved methane

conversion. The effect of equivalence ratio was more significant at intermediate


temperatures.

• For the platinum catalyst asubstantial decrease in the methane conversion was

associated with ageing. A fresh platinum catalyst exhibited higher reactivity

than aC0304/Cr2O3 catalyst in methane conversion. However, aC0304/Cr2O3

catalyst showed better performance in methane conversion than an aged

platinum one.

• No carbon monoxide was detected in the exhaust gas under any operating
conditions employed for the platinum catalyst. A noticeable increase in the

carbon monoxide concentration in the exhaust gas was observed with an

increase in the inlet temperatures and equivalence ratio for the C0304/Cr2O3
catalyst.

• The C0304/Cr2O3 catalyst exhibited higher thermal stability than the platinum

catalyst.

The developed mathematical model is sufficient for design purposes. The model was

used to provide a parametric investigation for fuel conversion within the packed bed

reactor such as inlet temperature, equivalence ratio, approach velocity, bed porosity, solid

thermal conductivity, and bed length. The developed model has proven successfully its

capability to provide the parametric study for the catalytic and the inert packed beds.
Accordingly, the costly and lengthy experimental verification can be reduced or

minimized.
169

9.2 Recommendations for Further Work

Several recommendations can be made for future work on catalytic oxidation of fuels

within packed bed reactors:

• To improve the modeling of the oxidation of fuel and air mixtures, it is

recommended that a detailed chemical kinetics for homogeneous reactions

should be included especially at low temperature to investigate the effect of

gas-phase radical formation on surface reactions.

• It is beneficial to include detailed surface chemistry in the modeling of catalytic

combustion in spite of the fact that it is very complicated and still at the early

stages of development. It can qualitatively account for the characteristic

behavior of the reactor and lead to further improvement in the combustion


processes.

• The model developed should be extended to consider the oxidation of lean

binary fuel mixtures in air.

• The model should be modified to include rich equivalence ratios for methane's

partial oxidation to produce hydrogen as apart of synthesis gas. This avenue, of

hydrogen production could be very beneficial.

• Obtaining experimental data on the effect of bed porosity for the oxidation of

lean fuel-air mixtures for the purposs of model validation. This could be done

by using different pellet diameters and/or different shapes to cover wider ranges

of porosity and topology of porous materials.

• A ceramic bed reactor capable of withstanding higher operating temperatures


than with quartz needs to be developed and employed.
170

REFERENCES

Adelman, B. J., Beutel, T., Lei, G.-D., and Sachtler, W. M. H., "Mechanistic Cause of

Hydrocarbon Specificity over Cu/ZSM-5 and Co/ZSM-5 Catalysts in the Selective

Catalytic Reduction of NO N", J. Catal., Vol. 158, pp. 327-335 (1996).

Agarwal, S. K., Jang, B. W.-L., Oukaci, R., Riley, A., and Marcelin, G., "NO Control by

Catalytic Combustion of Natural Gas", American Chemical Society chp. 18,.pp. 224-232
(1994).

Ahn, T., Pinczewski, W. V., and Trimm, D. L., "Transient Performance of Catalytic
Combustors for Gas Turbine Applications", Chem. Eng. Sci., Vol. 41(1), pp. 55-64

(1986).

Aleksandrov, E. N., and Azatyan, V. V., "Oxidation of Carbon Monoxide", Dokl. Akad.

Nauk. SSSR, Vol. 210, No. 6, pp. 1358-1361 (1973).

Alfa Aesar catalog, A Johnson Matthey Company, Ward Hill, Ma 01835-8099 'USA

(2001-02).

Anderson, S. J., Friedman, M. A., Krill, W. V., Kesselring, J. P., "Development of a

Small-Scale Catalytic Gas Turbine Combustor", Trans. ASME, Vol. 24, pp. 52-57
(1982).

Balakrishna, A., Schmidt, L. D., and Aris, R., "Pt-Catalyzed Combustion of CH4-C3H8

Mixtures", Chem. Eng. Sci., Vol. 49(1), pp. 11-18 (1994).'


171

Beebe, K, Cairns, K, and Pareek, V., "Development of Catalytic Combustion Technology

for Single-Digit Emission Industrial Gas Turbines", 4th International Workshop on

Catalytic Combustion, San Diego, 14-16 April (1999).

Bensalem, 0., and Ernst, W. R., "Mathematical Modeling of Homogeneous-

Heterogeneous Reactions in Monolithic Catalysts", Combust. Sci. Technol., Vol. 29, pp.

1-13 (1982).

Bharadwaj, S. S., and Schmidt, L. D., "Synthesis Gas Formation by Catalytic Oxidation

of Methane in Fluidized Bed Reactors", J. Catal., 146, pp. 11-21(1994).

Biloen, P., Dautzenberg, F. M., and Sachtler, W. M. H., "Catalytic Dehydrogenation of

Propane to Propene over Platinum and Platinum-Gold Alloys", J. ôatal., Vol. 50, pp. 77-

86(1977).

•Blazowski, W. S., and, Walsh, D. E., "Catalytic Combustion: an Important Consideration

for Future Applications", Combust. Sci. and Technol., Vol. 10, pp. 233-241 (1975).

Boehman, A. L., Niksa, S., and Moffat, R. 5., "Catalytié Oxidation of Carbon Monoxide

in aLarge Scale Planar Isothermal Passage", SAE 922332, pp. 119-128 (1992).

Bond, T. C., Noguchi, R. A., Chou, C.-P., Mongia, R. K., Chen, J.-Y., and Dibble, R. W.,

"Catalytic Oxidation of Natural Gas over Supported Platinum: Flow Reactor Experiments

and Detailed Numerical Modeling", Twenty-Sixth Symposium (International) on

Combustion/The combustion Institute, pp. 1771-1778 (1996).

Boreskov, G. K., J. Chim. Phys., Vol. 51, pp. 759 (1954).


172

Boudart, M., Collins, D. M., Hanson, F. V., and Spicer, W. E., "Reactions between H2

and 02 on Pt at Low and High Pressures: A Comparison", J. Vac. Sci. Technol., Vol.

14(1), pp. 441-443 (1977).

Bozo, C., Guilhaume, N., Garbowski, E., and Primet, M., "Combustion of Methane on

CeO 2-ZrO 2 Based Catalysts", 4th International Workshop on Catalytic Combustion, San

Diego, 14-16 April (1999).

Bradley, D., and Matthews, K. J., "Meauremënts .of High Gas Temperatures with Fine

Wire Thermocouples", J. Mech. Eng. Sci., Vol. 10(4), pp. 299-305 (1968).

Brown, N. J., Schefer, R. W., and Robben, F., "High-Temperature Oxidation of H2 on a

Platinum Catalyst", Combust. Flame, Vol. 51, pp. 263-277 (1983)

Bruno, C., Walsh, P.M., Santavicca, D.A., Sinha, N., Yaw, Y., and Bracco, F. V.,

"Catalytic combustion of propane/air mixtures on platinum", Combust. Sci. Technol.,

Vol. 31, pp. 43-73 (1983).

Bui, P.-A., Vaichos, D. G., and Westmoreland, P. R., "Homogeneous Ignition of

Hydrogen-Air Mixtures Over Platinum", Twenty-Sixth Symposium (International) on

Combustion/The Combustion Institute, pp. 1763-1770 (1996).

Butt, J. B. and Petersen, E. E., Activation, deactivation, and poisoning of catalysts,

Acadimic Press, Inc. (1988).

Cern, I., Saracco, G., and Specchia, V., "Methane Combustion over Low-Emission

Catalytic Foam Burners", Catal. Today, Vol. 60, pp. 21-32 (2000).
173

Chen, D. K. S., Bissett, E. J., Oh, S. H., and Ostrom, D. L. V., "A Three-dimensional

Model for The Analysis of Transient Thermal and Conversion Characteristics of

Monolithic Catalytic Converters", SAE paper 880282, pp. 1-13 (1988).

Cho, W., Back, Y.-S., Oh, Y., Lee, Y., Park, D., and Pang, H., "Development of the

Catalytic Burner of Natural Gas for a Textile Pad Dryer and aPrototype Boiler", 4th

International Workshop on Catalytic Combustion, San Diego, 14-16 April (1999).

Choudhary, V. R., Rajput, A. M., and Mamman, A. S., "NiO-Alkaline Earth Oxide

Catalysts for Oxidative Methane-to-Syngas Conversion: Influence of Alkaline Earth

Oxide on the Surface Properties and Temperature-Programmed Reduction/Reaction by

H2 and Methane", J. Catal., Vol. 178, pp. 576-585 (1998).

Choudhary, V. R., Rajput, A. M., and Prabhakar, B., "Nonequilibrium Oxidation of

Methane to CO and H2 with High Selectivity and Productivity over Ni/A1203 at Low

Temperatures", J. Catal:, Vol. 139, pp. 326-328 (1993).

Coffee, T. P., "On Simplified Reaction Mchanisms by Oxidation of Hydrocarbon Fuels

in Flames by C. K. Weestbrook and F. T. Dryer", Combust. Sci. Technol., Vol. 43, pp.

333-339 (1985).

Coltrin, M. E., Kee, R. J., and Rupley, F. M., "Surface Chemkin: A General Formalism

and Software for Analyzing Heterogeneous Chemical Kinetics at a Gas-Surface

Interface", Int. J. of Chem. Kinet., Vol. 3, pp. 1111-1128 (1991).

Corma, A., Palomares, A. E., and Fornes, V., "A Comparative Study on the Activity of

Metal Exchange MCM22 Zeolite in the Selective Catalytic Reduction of NOR", Res.

Chem. Intermed., Vol. 24, pp. 613-623 (1998).


174

Cullis, C. F., and Willatt, B. M., "Oxidation of Methane over Supported Precious Metal

Catalysts", J. Catal., Vol. 83 pp. 267-285 (1983).

Cybulski, A., and Moulijn, J. A., "Monoliths in Heterogeneous Catalysis", Catal. Rev.-

Sci. Eng., Vol. 35 (2), pp. 179-270 (1994).

Cybulski, A., and Moulijn, J., Structured Catalytsts and Reactors, Marcel Dekker (1998).

Dalla Betta, R. A., "Field Testing and Commercialisation of XONON Catalytic

Combustion Technology in Gas Turbines", 4tli International Workshop on Catalytic

Combustion, San Diego, 14-16 April (1999).

Davy, H., "Some New Experiments and Observations on the Combustion of Gaseous

Mixtures", in The Collected Works of Sir Humphery Davy (J. Davy, ed.), Vol. 6, Smith,

Elder and Co., Cornhill, London (1840).

Dekker, F. H. M., Dekker, M. C., Bliek, A., Kapteijn, F., and Moulijn, J. M., "A

Transient Kinetic Study of Carbon Monoxide Oxidation over Copper-Based Catalysts for

Automotive Pollution Control", Catal. Today, Vol. 20, pp. 409-422 (1994).

Depiak, A. M., "Catalytic Oxidation of Heated Lean Homogeneous Gaseous Fuel-Air

Streams" Ph D. Thesis, Department of Mechanical and Manufacturing Engineering, The

University of Calgary, Calgary, Alberta (2003).

Depiak, A. M., and Wierzba, I., "An Experimental Investigation of Oxidation of

Common Gaseous Fuels over Cobalt Oxide /Chromium Oxide Catalyst in aPacked-Bed

Reactor", ASME-Energy Sources Technology Conference & Exhibition, ETCE99-6709,

pp. 1-5 (1999).


175

Depiak, A. M., and Wierzba, I., "The Catalytic Oxidation of Heated Lean

Homogeneously Premixed Gaseous-Fuel Air Streams", J. Chem. Eng., Vol. 91, pp. 287-

295 (2003).

Desai, Amit .
J., Kovaichuk, Vladimir I., Lombardo, Eduardo A., and d'Itri, Julie L.,

"CoZSM-5: Why This Catalyst Selectively Reduces NO with Methane", J. Catal., Vol.

184, pp. 396-405 (1999).

Deutschmann, 0. D., and Schmidt, L. D., "Two-Dimensional Model of Partial Oxidation


of Methane on Rhodium in aShort Contact Time Reactor", Twenty-Seventh Symposium

(International) on Combustion/The Combustion Institute, pp. 2283-2291 (1998).

Deutschmann, 0., Schmidt, R., Behrendt, F., and Warnatz, J., "Numerical Modeling of

Catalytic Ignition", Twenty-Sixth Symposium (International) on Combustion/The

Combustion Institute, pp. 1747-1754 (1996).

Dupont, V., Moallemi, F., Williams, A. and Zhang, S.-H., "Combustion of Methane in

Catalytic Honeycomb Monolith Burner", mt. J. Engy Res., pp. 1-21 (2000).

Duterque, J., Borghi, R., and Tichtinsky, H., "Study of Quasi-Global Schemes for

Hydrocarbon Combustion", Cdmbust. Sci. Technol., Vol. 26, pp. 1-15 (1981).

Dwyer, F. G., "Catalysis for Control of Automotive Emissions", Catal-Rev., Vol 6, No.

2, pp. 261-291 (1972).

Gidaspow, D., and Ellington, R. T., A.I.Ch.E.J., Vol. 10, pp. 107 (1964).

Glassman, I., "The homogeneous oxidation kinetics of hydrocarbons: concisely and with

application", Dept. of mech. and Aerosp. Engin. Rep. 1446, Princeton Univ., Princeton,

N.J. (1980).
176

Goralski, C. T., and Schmidt, L. D., "Modeling Heterogeneous and Homogenous

reactions in the High-Temperature Catalytic Combustion of Methane", Chem. Eng. Sci.,

Vol. 54, pp. 579 1-5807 (1999).

Griffin, T. A., Calabrese, M., Pfefferle, L, D., Sappey, A., Copeland, R., and Crosley, D.

R., "The Influence of Catalytic Activity on the Ignition of Boundary Layer Flows. Part

III: Hydroxyl Radical Measurements in Low-Pressure Boundary Layer Flows", Combust.

Flame, Vol. 90, pp. 11-33 (1992)

Griffin, T. A., 'Pfefferle, L. D., Dyer, M. J., and Crosley, D. R., "The Ignition of

Methane/Ethane Boundary Layer Flows by Heated Catalytic Surfaces", Combust. Sci.

Technol., Vol. 65, pp. 19-37 (1989).

Groppi, G., Belloli, A., Tronconi, E. and Forzatti, P., "Comparison of Lumped and

Distri1uted models of Monolith Catalysts in Hybrid Combustors for Gas Turbines",

Chem.Eng. Sci., Vol. 50, pp. 2705-2715 (1995).

Groppi, G., and Tronconi, E., "Theoretical Analysis of Mass and Heat Transfer in

Monolith Catalysts with Triangular Channels", Chem. Eng. Sci., Vol. 52, No. 20, pp.
3521-3526 (1997).

Groppi, G., Tronconi, E. and Forzatti; P., "Mathematical Models of Catalytic

Combustors", Catal. Rev.-Sci. Eng., Vol 41(2), pp. 227-245 (1999)..

Hagen, J., Industrial Catalysis A Practical Approach, WILEY-VCH (1999).

Hanson, F. V., and Boudart, M., "The Reaction Between H2 and 02 over Supported

Platinum Catalysts", J. Catal, Vol. 53, pp. 56-67 (1978).


177

Harrison, B. K., and Ernst, "Catalytic Combustion in Cylindrical Channels: A

Homogeneous-Heterogeneous Model", Combust. Sci. Technol., Vol. 19, pp. 31-38

(1978).

Hautman, D. J., Dryer, F.L., Schug, K.P., and Glassman, I., "A multiple-step overall

kinetic mechanism for the oxidation of hydrocarbons", Combust. Sci. Technol., Vol. 25,

pp. 219-235 (1981).

Hayes, R. E., Kolaczkowski, S. T., and Thomas, W. J., "Finite-Element Model for a

Catalytic Monolith Reactor", Comp. Chem. Eng., Vol. 16, pp. 645-657 (1992).

Heywood, J. B., Internal Combustion Engine Fundamentals, McGraw-Hill, New York

(1988).

Hiam, L., Wise, H., and Chaikin, S., "Catalytic Oxidation of Hydrocarbons on Platinum",

J. Catal, Vol. 9, pp. 272-276 (1968).

Heck, R. M., and Farrauto, R. J., "Catalysis for Environmental Control", Special

Publication of the Royal Society of Chemistry, Chemically Modified Surfaces, Vol. 139,

pp. 12b-139 (1994).

Hicks, R. F., Qi, H., Young, M. L., and Lee, R. G., "Structure Sensitivity of Methane

Oxidation over Platinum and Palladium", J. Catal., Vol. 122, pp. 280-294 (1990).

Ikeda, H., Libby, P. A., and Williams, F. A., "Catalytic Combustion of Hydrogen-Air

Mixtur
es in Stagnation Flows", Combust. Flame, Vol. 93, pp. 138-148 (1993).

Ismagilov, Z. R., and Kerihentsev, M. A., "Catalytic Fuel Combustion - A Way of

Reducing Emission of Nitrogen Oxides", Rev.-Sci. Eng., Vol. 32(1 & 2), pp. 51-103

(1990).
178

Jones, R. L., "Surface and Coatings Effects in Catalytic Combustion in Internal

Combustion Engine", Surface and Coatings Technology 94-95, pp. 118-122 (1997).

Kang, S.-K., Jeong, N.-J., Rye, 1.-S., and Beak, Y.-S., "Combustion Control

Technologies of Catalytic Burner Applying to Reduction-Induced Furnace", 4th

International Workshop on Catalytic Combustion, San Diego, 14-16 April (1999).

Karim, G. A., and Klat, S. R., "The Measurement of the Mass Flow Rate of Different

Gases Using a Choked Nozzle", Journal of Laboratory Practice, Vol. 15, pp. 184-186

(1966).

Kendall, R. M., DesJardin, S. T., and Sullivan, J. D., Basic Research on Radiant Burners,

Gas Research Institute Report number 92/0181. Alzeta Corporation Report number 92-

7027-171 (1992).

Kesselring, J. P., Advanced Combustion Methods, Academic Press Inc. (London), (1986).

Kesselring, J. P., Krill, W. V., Atkins, H. L., Kendall, R. M., Kelley, and Kelley, J. T.,

"Design Criteria for Stationary Source Catalytic Combustion Systems", EPA-600/7-79-

181, August (1979).

Khalil, E. B. F., "Heat and Mass Transfer within Granular Beds" M.Sc. Thesis,

Department of Mechanical Engineering, The University of Calgary, Calgary, Alberta

(1993).

Khitrin, L. N., and Solovyeva, L. S., "Homogeneous-Heterogeneous Combustion of

Carbon' Monoxide in Narrow Tubes (Channels)", Seventh Symposium (International) on

Combustion, The Combustion Institute, Butterworth's Scientific Publication, London, pp.

532-538 (1959).
179

Koshland, C. P., "Impacts and Control of Air Toxics from Combustion", Twenty-Sixth

Symposium (International) on Combustion, The Combustion Institute, Pittsburgh; PA,

pp. 2049-2065 (1996).

Kuo, I. C. W., Morgan, C. R., and Lassen; H. G., "Mathematical Modeling of CO and

HC Catalytic Converter Systems", SAE Trans., Vol. 80, pp. 1098-1125 (1980).

Lee, H. H., and Ruckenstein, B., "Catalyst Sintering and Reactor Design", Catal. Rev.-

Sci. Eng., Vol. 25, pp. 475 (1983).

Li, S.-Y., and Li, B.-L., "Kinetic Characteristics of Catalytic Reactions of CO, HC and

NO on Different Three-Way Catalysts", React. Kinet. Catal. Lett., Vol. 56, pp. 283-289

(1995).

Li, S.-Y., and Li, B.-L., "Study on Non-Noble Metal Catalysts for Automotive Emissiion

Control" React. Kinet. Lett. Vol. 57, No. 1, pp. 183-190 (1996).

Magnus, E., and Jaras, S. G., "Catalytic Combustion of Low-Heating Value Fuel-Gases:

Ignition of CO, H2,and CH4,Symposium on Catalytic Combustion Presented Before the

Division of Petroleum Chemistry, Inc., 213 th National Meeting, American Chemical

Society, San Francisco, CA, April 13-17, pp. 146-150 (1997).

Markatou, P., Pfefferle, L. D., and Sinooke, M. D:, "A Computational Study of Methane-

Air Combustion over Heated. Catalytic and Non-Catalytic Surfaces", Combust. Flame,

Vol. 93, pp. 185-201 (1993).

Markatou, P., Pfefferle, L. D., and Smooke, M. D., "The Influence of Surface Chemistry

on the Development of Minor Species Profiles in the Premixed Boundary Layer

Combustion of an 112 /Air Mixture", Combust. Sci. Technol., Vol. 79, pp. 247-268

(1991).
180

Mehta, S. A., "An Examination of the Combustion and Transport Processes Within

Fractured Oil Sands Beds", Ph D. Thesis, Department of Mechanical Engineering, The

University of Calgary, Calgary, Alberta (1990).

Modest, M. F., Radiative Heat Transfer, McGraw-Hill, New York (1993).

Moallemi, F., Batley, G., Dupont, V., .Foster, T.J., Pourkashanian, M., and Williams, A.,

"Chemical Modelling and Measurements of the Catalytic Combustion of CH4Iair

Mixtures on Platinum and Palladium Catalysts", Catal. Today, Vol. 47, pp. 235-244

(1999).

Mori, Y., Miyauchi, T., and Hirano, M., "Catalytic Combustion of Premixed Hydrogen-

Air", Combust. Flame, .Vol. 30, pp. 193-205 (1977).

Nakajiina, F., "Air Pollution Control vith Catalysis", Catal. Today, Vol. 10, pp. 1-20

(1991).

Nàkhjava, A., Bjornbom, P., Zwinkels, M. F. M., and Jaras, S. G., "Numerical Analysis

of the Transient Performance of High-Temperature Monolith Catalytic Combustions:

Effect of Catalyst Porosity", Chem. Eng. Sci., Vol. 50(14), pp. 2255-2262 (1995).

Oh, S. H., and Cavendish, J. C., "Transients of Monolithic Catalytic Converters:

Response to a Step Change in Freestream Temperature as Related to Controlling

Automobile Eimission", I
nd. Eng. Chem., Prod. Res. Dev., Vol. 21, pp. 29 (1982).

Oh, S. H., Mitchell, P. J., and .Siewert, R. M., "Methane Oxidation over Alumina-

supported Noble Metal Catalysts with and without Cerium Additives", J. Catal., Vol. 132,

pp. 287-301 (1991).


181

Otto, K., "Methane Oxidation over Pt on 'y-Alumina: Kinetics and Structure Sensitivity",

Langmuir, Vol. 5, pp. 1364-1369 (1989).

Patankar, S. V., Numerical Heat Transfer and Fluid Flow, Hemisphere, Washington

(1980).

Pattas, K. N., Stamatelos, A. M., Pistikopoulos, P. K., Koltsakis, G. C., Konstandinidis,

P. A., Volpi, E., and Leveroni, E., "Transient Modeling of 3-Way Catalytic Converters",

SAE 940934, pp. 289302 (1994).

Pfefferle, L. D., and Pfefferle, W. C., "Catalysis in Combustion", Catal. Rev.- Sci. Eng.,

Vol. 29(2&3), pp. 219-267 (1987).

Pfefferle, L. D., Griffin, T. A, Dyer, M. J., and Crosley, D. R., "The Influence of

Catalytic Activity on the Gas Phase Ignition of Boundary Layer Flows Part II: Oxygen

Atom Measurements", Combust. Flame, Vol. 76, pp. 339-349 (1989).

Pfefferle, L. D., Griffin, T. A., Winter, M., Crosley, D. R., and Dyer "The Influence of

Catalytic Activity on the Ignition of Boundary Layer Flows Part I: Hydroxyl Radical

Measurements", Combust. Flame, Vol. 76, pp. 325-338 (1989).

Pfefferle, W. C., "Catalytically Supported Thermal Combustion", Belgian Pat. 814,752, 8

November (1974); and U.S. Pat. 3,928,961, 30 December (1975) (orig. filing 1971).

Pitchai, R., and Klier, K., "Partial Oxidation of Methane", Catal. Rev.-Sci. Eng., Vol. 28,

No. 1, pp. 13-88 (1986).

Prasad, R., Kennedy, L. A., and Ruckenstein, E., "Catalytic Combustion of Propane

Using Transitional Metal Oxides", Combust. Sci. and Technol., Vol. 22, pp. 271-280

(1980).
182

Prasad, R,, Kennedy, L. A., and Ruckenstein, E., "Oxidation of Fuel Bound Nitrogen in a

Transitional Metal Oxides Catalytic Combustor", Combust. Sci. and Technol., Vol. 27,

pp. 45-56 (1981).

Prasad, R., Tsia, H. L., Kennedy, L. A., and Ruckenstein, E., "Effect of Inlet Parameters

and Bed Length on the Operating Characteristics of aCatalytic Combustor", Combust.

Sci. and Technol., Vol. 25, pp. 71-84 (1981).

Prasad, R., Tsia, H. L., Kennedy, L. A., and Ruckenstein, E., "Occurance of Multiple

Steady States in the Catalytic Combustion of Propane", Combust. Sci. and Technol., Vol.

26, pp. 51-64 (1981).

Prasad, R., Kennedy, L. A., and Ruckenstein, E., "Catalytic Combustion", Catal. Rev.-

Sci. Eng., 26, pp. 1-58 (1984).

Prather, M. J., and Logan, J. A., "Combustion's Impact on the Global Atmosphere", -

Twenty-Fifth Symposium (International) on Combustion, The Combustion Institute,

Pittsburgh, PA, pp. 1513-1527 (1994).

Ragaini, V., Vitali, S., and Luzzato, D., "Development of a Catalytic Burner and

Evaluation of Its Performances in Comparison with Other Not Catalysed Burners", 4th

International Workshop on Catalytic Combustion, San Diego, 14-16 April (1999).

Razon, L. F., and Schmitz, R. A., "Intrinsically Unstable Behavior during the Oxidation

of Carbon Monoxide on Platinuth", Catal. Rev.-Sci. Eng., Vol. 28, No. 1, pp. 89-164

(1986).

Reid, R. C., Prausnitz, J. M., and Poling, B. E., The Properties of Gases and Liquids, 4th

Ed., McGraw-Hill, New York (1987).


183

Richardson, J. T., Principles of Catalyst Development, Plenum Press, New York, pp. 104-
106(1989).

Ruckenstein, E., and Dadyburjor, D. B., "Sintering and Redispersion of Supported Metal

Catalysts" Catal. Rev.-Sci. Eng., Vol. 1, pp. 251 (1983).

Rumminger, M. D., Hamlin, R. D., Dibble, R. W., "Numerical Analysis of aCatalytic


Radiant Burner: Effect of Catalyst on Radiant Efficiency and Operability", Catal. Today,

Vol. 47, pp. 253-262 (1999).

Schefer, R. W., "Catalyzed Combustion of H2/Air Mixtures in aFlat Plate Boundary

Layer: IL Numerical Model", Combust. Flame, Vol. 45, pp. 171-190 (1982).

Schefer, R. W., and Robben, and Cheng, R. K., "Catalyzed Combustion of H2/Air

Mixtures in aFlat-Plate Boundary Layer: I. Experimental Results", Combust. Flame. Vol.

38. pp. 51-63 (1980).

Schlegel, A., Benz, P. Griffin, T., Weisenstein, and Bockhorn, H., "Catalytic

Stabilization of Lean Premixed Combustion: Method for Improving NO x Emission",

Combust. Flame, Vol. 105, pp. 332-340 (1996).

Scholten, A., Van Yperen, R., and Emmerzaal, I. J., "A Zero NO Catalytic Ceramic

Natural Gas Cooker", 4th International Workshop on Catalytic Combustion, San Diego,
14-16 April (1999).

Seinfeld, J. H., Atmospheric Chemistry and Physics ofAir Pollution, John Wiely & Sons,

New York, 1986.


184

Seonhi, R., and Scholten, A., "Comparison of Catalytic and Catalytically Stabilised

Domestic Natural Gas Burners", 20th World Gas Conference Proceedings, Copenhagen,

(1997).

Smith, L., Karim, H, Castaldi, Etemad, S., and Pfefferle,W., "Catalytic Combustion f6r

Ultra Low Emission Ground Power Application", 4th International Workshop on

Catalytic Combustion, San Diego, 14-16 April (1999).

Sobolev, G. K., "High-Temperature Oxidation and Burning of Carbon Monoxide", Sixth

Symposium (International) on Combustion, Reinhold, New York, paper No. 57, pp. 386

(1957).

Sokolovskii, V. D., "Principles of Oxidative Catalysis on Solid Oxides", Catal. Rev.-Sci.

Eng., Vol. 32(1 & 2), pp. 1-49 (1990).

Song, X., W i
lliams, W. R., Schmidt, L. D., and Ails, R., "Bifurcation Behavior in

Homogeneous-Heterogeneous Combustion: II. Computations for Stagnation-Point Flow,

Combust. Flame, Vol. 84, pp. 292-311 (1991).

Song, X., Williams, W. R., Schmidt, L. D., and Aris, R., "Ignition and extinction of

homogeneous-heterogeneous combustion: CH4 and C3H8 oxidation on PT", Twenty-

Third Symposium (International) on Combustion, The Combustion Institute, pp. 1129-

1137(1990).

T'ien, J. S., "Transient Catalytic Combustor Model", Combust. Sci. and Technol.; Vol.

26, pp. 65-75 (1981).

Tochihara, Y., and Ozawa, Y., "Improvement of Catalyst Using Catalytic Combustor for

aGas Turbine", 4th International Workshop on Catalytic Combustion, San Diego, 14-16

April (1999.). .
185

Torniainen, P. M., Chu, X., and Schmidt, L. D., "Comparison of Monolith-Supported

Metals for the Direct Oxidation of Methane to Syngas", J catal., Vol. 146, pp. 1-10

(1994).

Treviflo, C., "Catalytic Flat Plate Boundary Layer Ignition", Combust Flame, Vol. 26, pp.
245-251 (1981).

Treviflo, C:,. and .Peters, N., "Gas Phase Boundary Layer Ignition on aCatalytic Flat Plate

with Heat Loss", Combust. Flame, Vol. 61, pp. 39-49 (1985).

Trimm, D. L., and Lam, C.-W., "The Combustion of Methane on Platinum-Alumina

Fiber Catalysts-I Kinetics and Mechanism", Chem. Eng. Sci., Vol. 35, pp. 1405-1413

(1980).

Twigg, M. V., Catalyst HandbooK, Manson Publishing Ltd., London (1996).

Twigg, M. V., and Wilkins, A. I 1, "Autocatalysts-Past, Present, and Future" in

Structured Catalysts and Reactors, Marcel Dekker, Inc., chapter 4, pp. 91-120 (1998).

Quinlan, M., Wise, H., and McCarty, J., "Basic Research on Natural Gas Combustion

Phenomena-Catalytic Combustion";, GRI Annual Report, pp. 1-59 (1986).

Vaichos, D. G., "The Interplay of Transport, Kinetics, and Thermal Interactions in the

Stability of Premixed Hydrogen IAir Flames Near Surfaces", Combust. Flame, Vol. 103,

pp. 59-75 (1995).

Valchos, D. G., Schmidt, L. D., and Aris, R., "Ignition and Extinction of Flames Near

Surfaces: Combustion of H2 in Air", Combust. Flame, Vol. 95, pp. 313-335 (1993).
186

Voltz, S. E., Morgan, C. R., Liedermañ, D., and Jacob, S. M., "Kinetic Study of Carbon

Monoxide and Propylene Oxidation on Platinum Catalysts", I


nd. Eng. Chem. Prod. Res.

Develop., Vol. 12(4), pp. 294-301 (1973).

Wakao, N., and Kaguei, S., Heat and Mass Transfer in Packed Beds, Gordon and Breach

Science Publishers (1982).

Wampler, F. B., Clark, D. W., and Gaines, F. A., "Catalytic Combustion of C3H8 on Pt

Coated Monolith", Combust, Sci. Technol.., Vol. 14, pp. 25-31 (1976).

Wanicer, R., Raupenstrauch, H., and Staudinger, G., "A fully Distributed Model for the

Simulation of aCatalytic Combustor", ,Chem. Eng. Sci., Vol. 55, pp. 4709-4718 (2000).

Wark, K. Jr., Thermodynamics, McGraw-Hill, New York (1988).

Westbook, C. K., "Simplified Reaction Mechanisms for the Oxidation of Hydrocarbon

Fuels in Flames", Combust. Sci. Technol., Vol. 27, pp. 31-43 (1981).

Will, N. S., and Bennett, C. J., "Flow Maldistributions in Automotive Converter

Canisters and their Effect on Emission Control", SAE 922339, pp. 18.3-210 (1992).

Younis, L. B., and Wierzba, "Numerical Analysis of Methane-Air Oxidation within

Catalytic and Non-Catalytic Packed Bed Reactors", The Combustion Institute Canadian

Section, Spring Technical Meeting, Windsor, Ontario, pp. 45.1-45.7 (2002).

Younis, L. B., and Wierzba, I., "A Numerical Investigation of Catalytic Oxidation of

Very Lean Methane-Air Mixtures within aPacked Bed Reactor", to be published in J.

Porous Media, Vol. 2(2004).


187

Zabetakis, M. G., "Flammability Characteristics of Combustible Gases and Vapors", U.S.

Bureau of Mines, Bulletin 627, Pittsburgh, Pennsylvania, (1965).

Zwinkels, M. F. M., Jaras, S. G., Govind Menon, P., and Griffin, T. A., "Catalytic

Materials for High-Temperature Combustion", Catal. Rev.-Sci. Eng., Vol. 35(3), pp.319-

358 (1993).
188

Appendix A

Thermophysical Properties

The specific heat at constant pressure of the gas mixture Cp g and the transport properties:

the mass diffusion coefficient Dk, and thermal conductivity of the gas mixture kg in the

calculation are listed in the following sections. Properties for each component are

calculated as afunction of the local temperature.

Ad Gas Specific Heats

The specific heat at constant pressure (heat capacity) for the pure components considered

in the modeling is determined from the following formula (Heywood, 1988):

Cp g (kJ/kmol.K) = 4.184 (ai + a2 0+ a3 92 + a4 0 3 + a50 -


2) (A. 1)

where 0 T (K) /1000 and ai, a2, a3, a4,and as are constants (curve fit coefficients)'

specific to the gas as listed in Table A. 1.

Table A.1 Gas specific heat correlation constants,

component a1 a2 a3 -
a
4 a5

CH4 -0.29149 26.327 -10.610 1.5656 0.16573

C3R8 -1.4867 74.339 -39.065 8.0543 0.01219

For air and other ideal gases the specific heat at constant pressure in (J/kmol.K) is

obtained from the following correlation (Wark, 1988):


189

Cpg
= a+bT+cT 2 +dT 3 +eT 4 (A.2)
R

The constants a, b, c, d, and ear listed in Table A.2 and T is in (K).

Table A.2 Gas specific heat correlation q. A.2) constant.

Component a bx10 3 cx 10 6 dx 10 9 fexlO'2 1


CO 3.710 -1.619 3.692 -2.032 0.240

H2 3.057 2.677 -5.810 5.521 -1.812

02 3.626 -1.878 7.055 -6.764 2.156

Air 3.653 -1.337 3.294 -1.913 0.2763

A.2 Diffusion Coefficient

The binary diffusion coefficients are calculated based on the Chapman-Enskog

theoretical description of the binary mixtures of gases at low to moderate pressures. In

this theory, the binary diffusion coefficient for the species pair A and B is

DAB 3 (42rkB T/MW)


2 fD (A.3)
16 (F/RU T)
-

AB 2D

where kB is the Boltzmann constant, T (K) is the absolute temperature, F (Pa) is the

pressure, Ru is the universal gas constant, and fD is atheoretical correction factor whose
190

value is sufficiently close to unity to be assumed the same. The remaining terms are

defined below: -

MWAB = 2{(1/M) + (1/MWB)f' (A.4)

where MJVA and MWB are the molecular weights of species A and B, respectively;

cYAB = (0 A + (A.5)

where qA7 and o are the hard-sphere collision diameters of species A and B,

respectively, values of species employed in the model prediction are shown in Table D.3.

The collision integral, 92 D, is a dimensionless quantity calculated using the following

expression:

A C E G
nD = * + * + * + * (A.6)
(2' ) exp(DT ) exp(FT) exp(HT )

where

A = 1.06036, B = 0.15610,

C=0.19300, D=0.47635,

E=1.03587, F=1.52996,

G= 1.76474, H=3.89411,

and where the dimensionless temperature f is defined by

= kB T/e = kB T/(SA SB )112 (A.7)

Values of the characteristic Lennard-Jones energy, Si, are also tabulated in Table A.3

(Reid et al., 1987).


191

Table A.3 Lennard-Jones parameters for species employed in the calculation.


o
Species a(A) e/1GB (K)

Air 3.711 78.6

CH4 3.758 148.6

CO 3.690 91.7

C2R 4.443 215.7

C3H8 5.118 237.1

112 2.827 59.7

02 3.467 106.7

Substituting numerical values for the constants in Eq D.3 results in

0.0266 T1/2
(A.8)
= FWWA112 OAB ≤ D

with the following associated units: D[=] m2/s, T[=] K, P [=] Pa, and o[=] A.

A.3 Gas Thermal Conductivity

The pure gas component thermal conductivity is determined from the following formula

(Oh and Cavendish, 1982):

kg = 2.269x10 Tg
°832 (A.9)

where Tg is in (K), and k


g is in (W/cm.K)
192

Appendix B

Choked Nozzles Calibration

The fuel (methane) and air were metered by the choked flow nozzle system connected to

the experimental set-up. These nozzles were calibrated with awet test meter using air as

the working fluid. These results were then converted to yield flow rates for methane

using the equation suggested by Karim and Klat (1966):

mair Cair IkIlVair pair


(B.1)
-

MWCH4 CH4
mcu 4

where m is the mass flow rate, C is the coefficient of discharge, MW is the molecular

weight.

Calibration of all choked nozzles was conducted while they were mounted on the panel
where they were eventually used. The calibration curves for air and methane with their

corresponding equations are shown in Fig. B.1, B.2, B.3, B.4 and B.5.

4,00E-02

3.50E-02

3.COEO2

2.50E-02

2.COE-02

1.50E-02

1.E-02

5.COE-03

O.00E+0O
0 5 10 15 20 25 30 35 40 4

pi-r° 5(kpaIK°5 )
Fig. B.1 Calibration curves for 0.09 mm diameter nozzle.
193

0.25

0.2 y= 0.0053x + 0.0061

E
0
>
0.05

0
0 5 10 15 20 25 30 35 40 45

pir ° (kPa/K ° )

Fig. B.2 Calibration curves for 0.20 mm diameter nozzle.

0.6

0.5

0.4

a
0
0.3
I..

0
E 0.2
0
>

0.1

0
0 5 10 15 20 25 30 35 40 45

PIT (kPa/K °)

Fig. B.3 Calibration curves for 0.34 mm diameter nozzle.


194

1.4

1.2

E
0.8
0

I-

o9
.6

0
E
0.4

0.2

0
5 15 20 25 30 35 40 45

PIT °5 (kPa/K °)

Fig. B.4 Calibration curves for 0.61 mm diameter nozzle.

6.00E+00

5.00E+00
y 0.2183x - 0.3662,' -

I19i

C.,

E 4.00E+00

0
CV
3.00+0O

2.00E+00

l.00E+00.

0.E+00
0 5 10 15 20 25 30
PIT °5 (kPa/K °5 )

Fig. B.5 Calibration curves for 1.49 mm diameter nozzle.


195

Appendix C

Thermocouple Correction

The error in the, thermocouple measurements due to radiation and convection was

calculated using the following formula (Bradley and Mathews, 1


,968):

0.
Tcorrec/
ed = Tuncorrecte d +
(C.1)
h

where 0. is Stefan Boltzmann constant, s is the emissivity, and his the convection heat

transfer coefficient.
The temperature distribution' at the inlet across the reactor was measured using a

thermocouple sheathed-with 600 Inconel as shown in Fig. C.1. This thermocouple was,

first calibrated with adigital calibration thermometer.

25

20

15
E
>-

10

0
0 5 10 1 15 20 25 30 35 40 45 50

Temperature (
°C)

Fig. C.1 Temperature distribution across the reactor, Ti,, = 41 °C.


196

Appendix D

Pressure Drop

The pressure drop within the packed bed was measured using differential pressure

transducer. Figure D. 1shows the pressure drop as afunction of the flow rates while Fig

D.2 illustrate the pressure drop across the bed as afunction of the approach velocities. 1

0.6

0.5 4

0.1

0
00 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 50

Flow rate (m 3lh)


Fig. Di. Pressure drop across the packed bed reactor as afunction
of the flow rate at Tm= 20 °c and P = 88.1 kPa.
197

0.6

0.5

0.4

a.

0.
0
0.3
U

U)
U)
U

0.2

0.1
*

0
00 0.6 1.0 1.5 2.0 25

Aproach Velocity (mis)

Fig. D.2 Pressure drop across the packed bed reactor as afunction

of the approach velocity at Tin = 20 °C and P= 88.1 kPa.


198

Appendix E

Gas Analysis

The concentration of major species such as CH4,CO, CO2,and 02 in the exhaust gas was

measured and recorded continuously using Beckman gas analyzers. The measured

concentrations of the species were based on adry volumetric percentage. Assuming that
there is no carbon monoxide (CO) formed which was confirmed experimentally. The

overall combustion equation can be written as:

xCH4 + (1-x) (0.2102 + 0. 79N2) - aCH4 - bCO2 + dH20 + e02

+ 0. 79('l-x) N2 (E.1)

where xrepresent the initial concentration of methane in the mixture while a, b, d, and e

represent the corrected concentrations of CH4,CO 2,H20 and 02, respectively in the

exhaust gas. The corrected concentrations of the mentioned species were calculated on

the basis of elemental mass balance as follows:

a (E.2)

(E.3)

= (O.79+O.2l)r m
[(1—rn)]
1
where k, 1and m represent the measured concentrations (dry basis) of CH4,CO 2 and 02,
respectively.

All the experimental data presented for species conversion were the corrected ones

obtained from equations (E.2 - E.4). The comparison of the "dry" and "wet" conversion
over the range of 0to 50% methane conversion, for methane-air mixture of equivalence

ratio 0.35, shows amaximum of 1.8%difference. In all cases with the platinum bed, CO

concentrations were below 1ppm, which is the detection limit of the CO gas analyzer.
Appendix F

Program Listing

1cpsl(4),porl(4),sigel(4),pdl(4),wol(4)
c ANONEQTJILIBRIUM MATHEMATICAL MODEL PROGRAM OF A dimension amda(11),eta(11),sige(11),pd(11),amu(11)
PACKED-BED REACTOR dimension condt(11),convt(I1),unst(11),radt(11)
c BYL.B.YOUNIS C
c THE MECHANICAL AND MANUFACTURING ENGINEERING, ENTRY GRID
UNIVERSITY C
OF CALGARY, CANADA. sumro0.
o SEP. 2001 iter=0.
CCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCC00000C00000CCCCCCCCCCCC nts=1
SUBROUTINE USER C Uniform grid
xu(2)=0.
logical lsolve,lstop dxxllfloat(11-2)
parameter(11 10 1) do 1i=3,11
common f(ll,6),rho(11),gam(11),con(11),aip(11),aim(11), 1xu(i)xu(i-1)+dx
1ap(11),x(11),xu(11),xdif(11),xcv(11),xcvs(L1),fo(11,6), return
2u(ll),xcvi(11),xcvip(11) C
common fr(ll),fxm(11),pt(11),qt(11) ENTRY START
commonhindx/nfnfinax,12,13,ist,iter,1astre1ax(5), C
ltime,dt,xl,lsolve(5),ntimes(5) ntsltmax/dt
common/cntlllstop,ntsl,nts nco0
commonIcoe'flow,diffacof isw=1
c do2i1,11
tf(i)tin
dimension hf(11),tf(11),ts(ll),ym(11),rhof(11),rhos(11),yf(11), ts(i)tin
1yper(ll) ym(i)fu11n
equivalence (f(1, 1),tf(1)),(f(1,2),ts(1)),(f(l,3),ym(1)), wfu(i)=0.
1(f(1,4),hf(1)) dqdx(i)=0.
real tkf(11),tks(ll),qrp(11),qrm(11),qrt(11),dqdx(11) 2 continue
real cpf(ll),cps(11),por(ll),hv(11),wfti(11) do 5i1,ll
real difc(ll),xo2(11),sr(11),wfuo(11),wo(ll),wsu(11),wsuo(11) do5n'l,6
dimension xly(4),xloc(4),nle(4),ñlb(4),rhosl(4),tksl(4), 5 fo(i,n)f(i,n)
C 22 continue
c print of input data--------- Cl 111111 Ill-H-Ill II
print ,' do i=2,12
print *,hl Multi-Layer xsp0.0
print fsp=0.
return if(x02(i).gt.0.0) xsp(rhof(i)*x02(i)/32.0)**0.5
C if(ym(i).gt.0.0) fsp=(rhof(i)*ym(i)/16.0)
ENTRY DENSE temps—ts(i)
C wsusprecat*xsp*fsp*exp(engsI(83 14.0*temps+1.0e20))
do 10 i1,11 wsu(i)1.0*((2.25/1.0)*wSus)
1) rho(i)=por(i)*rhof(i)*cpf(i)
if(nfeq. end do
if(nf.eq.2) rho(i)=(L_por(i))*rhos(i)*cps(i)
if(nfeq.3) rho(i)rhof(i) return

10 continue
return ENTRY OUTPUT
C
ENTRY BOUND if(iter.ne.0) return
C if(nts.ge.ntsl) go to 310
do 12 1=1,11 printcprintc4dt
u(i)un1et if(printc.lt.timepr) return
xo2(i)=xo2instmr*(fu1inym(i)) printc=0
12 continue 310 print *,hflme&,time

fmass=u(1)*rhof(1) print 316

heats=fmass*fulin*hfg 316 format(lx,'x', l0x,'ts', 10x,tf, 14x,'qrt', lOx,


heatsf=heats/flamt 1'qr+', lOx,'qr-', l0x,'Ym')
tf(1)=(u(1)*cpf(1)*rhof(1)*tgin+tkf(1)*tf(2)/xdif(2))/ do 315 i=1,11
1(u(1)*cpf(1)*rhof(1)itkf(1)/xdif(2)) print 317,x(i),ts(i),tf(i),qrt(i),qrp(i),qrrn(i),ym(i)
lf(11)=tf(12) 315 continue
ts(11)=ts(12) 317 format(lx,f6.4,2x,fl0.3,2x,flo.3,2x,f12.3,
ym(1)fulin 12x,fl2.3,2x,f12.3,2x,f12.3,2x,elO.3)
ym(11)ym(12) if(nts.lt.ntsl) return
hf(1)=hfin pel=u(1)*rhof(1)*cpf(1)*xdif(2)/tkf(1)
hf(11)=hf(12) aac=amaxl(0.0,(1.-0. 1*abs(pel))**5)
do 22 i=1,11 smkl=tkf(1)*aac*(tf(2)_tf(1))/xdif(2)
re=pd(i)*u(1)*rhof(1)/amu(i) enfl=u(1)*rhof(1)*cpf(1)*tgin
pr=(amu(i)*cpf(i)/tkf(i))**0.3333 enfo=u(1)*rhof(1)*cpf(11)*tf(11) -

hv(i)=6.0*tkf(i)*(2.0+1. 1*pr*re**0.6)/(pd(i)*pd(i)) ensi=tks(1)*(ts(2)_ts(1))/xdif(2)


enso=tks(11)*(ts(12)_ts(11))/xdif(11) 415 continue
qrad=(1.rfao)*qrm(1)+ endif
1(1._rfal)*qrp(11) do 412 i=2,12
1 emi*stf*por(1)*tgin**4emo*stf*por(11)*tamo**4 412 con(i)=con(i)+por(i)*wftlo(i)*hfg
heattO. go to 450
do 318 i=2,12 420 do 422 i=nfb,nfe
318 heatt'=heatf+w.fij(i)*hfg*xcv(i) 422 con(i)=con(i)+heatsf
if(time.gt.thin) heats=heatf 450 continue
balance=heats-I-enfl-enfo-enso-ensi-qrad rresi=rhof(i)*u(1)*cpf(i)*tkf(i)/(tkf(1)+rhof(i)*u(i)*cpf(1)
balance--balance/heats 1*xdjf(2))
lstop=.true. con(2)=con(2)+rresi*tginlxcv(2)
return ap(2)=ap(2)-rresilxcv(2)
C gam(11)=0.
ENTRY GAMSOR ,gam(i)=0.
C c---Solid matrix
c 1gas temperature 500 if(nfne.2) go to 600
c 2solid matrix temperature do 510 i=2,12
c 3mass fraction of the products ap(i)=_(1._por(i))*hv(i)
do i=2,12 con(i)(i._por(i))*hv(i)*tf(i)
xJflJ(j)=(ff*((ym(j)*rhof(j)/16O)**O2 c radiation effect
1*(x02(j)*rhof(j)/320)**]3)* radtt= 16. *slf*ts(i)*ts(i)*ts(i)/(3.*sige(i))
1exp(-eng/(83 14.0*tf(i)+1.0e_20))) gam(i)=gam(i)+(1._por(i))*radtt
end do 510 continue
do 401 i1,11 c---- set inlet boundary condition
gam(i)=por(i)*tkf(i) gam(i)=0.
if(nf.eq.2) gam(i)=(1._por(i))*tks(i) gam(li)=0.
if(nfeq.3) gam(i)=difc(i)*rhof(i) c for catalytic reaction
401 continue if(iter.eq. (last-i)) then
if(nfne.1) go to 500 do i2,12
do 410 1=2,12 wsuo(i)wsu(i)
con(i)(1._por(i))*hv(i)*ts(i) end do
ap(i)=_(1._por(i))*hv(i) end if
410 continue do i=2,12
if(time.lt.tnin) go to 420 con(i)con(i)+(1.0_por(i))*wsuo(i)*hfg
if(iter.eq. (last-i)) then end do
do 415 i=2,12 C
wfuo(i)=wfu(i) 600 if(nfne.3) go to 700
do 610 i=2,12 difc(i)1.00e-5
con(i)0.5*wfil(i) amu(i)=(1.40 le_6*t['(i)** 1.5)1(108.5 1+tf(i))
ap(i)_1.5*wfu(i)/(ym(i)+1.e_10) 800 continue
610 continue rhof(1)883251(rair*tin)
gam(11)=0. do 810 j1,nl
700 continue do 810 i=nlb(j),nle(j)
return por(i)=porl(j)
C rhos(i)=rhosl(j)
ENTRY PROP tks(i)=tksl(j)
C cps(i)=cpsl(j)
c----av =Surface area per unit volume pd(i)=0.0032/pdl(j)
c---- ii = Heat transfer coefficient sige(i)=sigel(j)
c---- hfg = Heat content of the fuel wo(i)=wol(j)
c 5xe7 J/Kg 810 continue
c---- p01= porosity C
if(nts.le. 1) then return
stf=5.67e-8
print *,'exafr',exafr C
fulin=16./(16.+(1.+exair)*274.56)
x02in64.*(1.+exair)/(16.+274.56*(1.+exair)) c******************************************************
c precat is pref'actor for catalyst and engs is activation energy for solid c=======Subroutine Layer to calculate the beginning and ending
c precat1.03e-5 node of each layer
engs=0.86e8 ENTRY LAYER
precatl.03e-3 sum0.
engs=0.46e8 do 900 j=1,nl
stmr4. sumsum+xly(j)
engl.8e8 xloc(j)sum
ff6.6e10 900 continue
runi=83 14.0 suni10.
hfg=5.00e7 do 915 j=1,nl
rair=run1128.97 do 910 i=2,12
endif sumlsuml+xcv(i)
do 800 i=1,11. if(suml.ge.xloc(j)) go to 920
c Density of air as function of P and T. 910 continue
rhof(i)=88325.0I(rair*tf(i)) 920 nle(j)=i
cpf(i)=1160. nlb(j+1)i+1
tkf(i)=(2.645 le-3 *tf(i)** 1.51(230.68+tf(i))) 915 continue
nle(nl)=11 read(3,*)unlet
nlb(1)=1 read(3,*)
C Calculated pre-set flame position read(3,*)exair
sunîfb=0. read(3,*)
sumfe=0. read(3,*)tin,tgin,tami,tamo
do 925 1=2,12 read(3,*)
sumfb=sumfb+xcv(2) read(3,*)hi,ho
if(sumfb.ge.floc) go to 930 read(3,*)
925 continue read(3,*)emi,emo
930 nfb=i c read(3,*)
do 935 i=2,12 c read(3,*) yJ
sumfe=sumfe+xcv(2) read(3,*)
if(sumfe.ge.(floc+fth)) go to 940 read(3,*) ni
935 continue read(3,*)
940 nfe=i read(3,*)(xly(i),11,nl)
flamt=0. sum0.0
do 941 intb,nfe dok=1,nl
941 flamt=flamt+xcv(i) sumsuni+xly(k)
print *,'Number of Layers=',nl end do
print *,'xloc. of each layer',(xloc(j)j=l,nl) xlsum
print *,'Beg inning node of each 1ayer'(nlb(j),j=1,nl) read(3,*)
print *,'End node of each 1ayer',(nle(j),jl,nl) read(3,*)(porl(i),i1,nl)
print *,'flame location',floc read(3,*)
print ,flame thickness',fth
* ' read(3,*)(rhosl(i),i1,nl)
print *,'Beginning node of flãme',nfb read(3,*)
print *,'End node of flame',nfe read(3,*)(tksl(i),i1,nl)
print *,f1ajne',f1amt read(3,*)
return read(3,*)(cpsl(i),i 1,nl)
C INPUT DATA read(3,*)
ENTRY INPUT read(3,*)(pdl(i),i1,nl)
read(3,*) read(3,*)
read(3,*)dt,tmax read(3,*)floc,fth
read(3,*) read(3,*)
read(3,*)timepr read(3,*)
read(3,*). read(3,*)(sigel(i),i=1,nl)
read(3,*)tnin return
read(3,*)
C OUTPUT
REPORT write(4,*)'
ENTRY REPORT
open(4,file='outp') write(4,*)I! SELECTED OUTPUT PARAMETERS II'
write(4,*)
write(4, * )' OUTPUT REPORT----------- write(4,*)'
write(4,*)'
write(4,*)hINPUT DATA' write(4,*)
write(4, *)'Flow rate of fuel and air (m3/s)
' balance=balance* 100.
write(4,*)volf,volair write(4,*)'Energy imbalance =',balance,' %'
write(4,*)'Cross sectional area of the duct (m2)' heatsheats*area
write(4,*)area heatb=heats*3 .413
write(4,*)hlnifial, gas inlet,and outlet ambient temperatures (K)' write(4,*)'Heat released by combustion=',heats,'W',heatb,'Btu/hr'
write(4,*)tin,,tgin,tarno exair=exair* 100
write(4,*)'Heat transfer coeff. at the inlet and outlet (W/m2 K)' write(4,*)'Excess air ',exair,' %'
write(4,*)hi,ho write(4,*)
write(4,*)'Emissivity of matrix at inlet and outlet surfaces' write(4,*)' AT THE OUTLET
write(4,*)emi,emo
write(4, *)'Total thiclaiess of the poous layers (rn)' C Plotting Files
write(4,*) xl open(17,file='xtgas')
write(4,*)'Number of layers' open(8,file='xtsol')
write(4,*) n1 open(9,f11e'xqrnet')
write(4,*)'Thickness of each layer (m)' open(10,fi1e'xfuel')
write(4,*)(xly(i),i=1,nl) open(11,fi1e'x02')
write(4,*)'Porosity of each layer' open(33,fi1e'xwfus')
rite(4,*)(porl(i),i1,nl) open(36,fi1e'xwsus')
write(4,*)'Density of solid matrix of each layer (kg/m3)' open(37,file='perfuel')
write(4,*)(rhosl(i),i=1,nl) do 1100 i1,ll
write(4, * )'Thermal conductivity of matrix of each layer (W/mK)' write(17,*)x(i),tf(i)
write(4,*)(tksl(i),i= 1,nl) write(8,*)x(i),ts(i)
wr.ite(4,*)'Specffic heat of each layer' write(9,*)x(i),qrt(i)
write(4,*)(cpsl(i),i=1,nl) write(10,*)x(i),ym(i)
write(4, *)'Number of pores per inch to each layer (PPI)' write(11,*)x(i),x02(i)
write(4,*)((i),i=1,nl) write(33,*)x(i),wfu(i)
write(4,*)'RADIATION PROPERTIES' write(36,*)x(i),wsu(i)
write(4,*)'Extinction coefficient of each layer (1/rn)' yper(i)((fulin_yln(i))ffulin)* 100.
write(4,*)(sigel(i),i=1,nl) write(37,*)x(i),yper(i)
1100 continue tkefw=2.*tks(i)*tks(i_1)/(tks(i)+tks(i_1))
return tkefe2.*tks(i)*tks(i+1)/(tks(i)+tks(i+1))
c condte=tkefe*(ts(i+1)ts(i))/xdif(i+1)
C Subroutine to calculate energy term contributions condtw=tkef'w*(ts(i)ts(i_1))/xdif(i)
entry TERM condt(i)(condte-condtw)/xcv(i)
if(iter.ne. last-i) return
advt=advt+dt convt(i)hv(i)*(tf(i)_ts(i))
if(advt.lt.50.) return unst(i)rhos(i)*cps(i)*(ts(i)_fo(i,2))/dt
advt=0. radt(i)dqdx(i)/xcv(i)
1200 continue
write(7,*) 'time=,time do 1220 i=2,12,2
write(7, 1202) 1220 write(7, 123 0)x(i),unst(i),condt(i),convt(i),radt(i)
1202 format(6x,'x',9x,'storge',9x,'conduction1,7x,'convection',6; 1230 forniat(fl0.5,4e16.5)
1'radiation') return
do 1200 i2,12 end

You might also like