Sie sind auf Seite 1von 149

i

SIMULATION OF CATALYTIC COMBUSTION OF AUTOMOTIVE EXHAUST


GAS IN MONOLITHIC REACTOR

TIANG HOCK YU

A project report submitted in partial fulfilment of the


requirements for the award of the degree of
Bachelor (Honours.) of Chemical Engineering

Faculty of Engineering and Science


Universiti Tunku Abdul Rahman

April 2019
ii

1 DECLARATION

I hereby declare that this project report is based on my original work except forcitations and
quotations which have been duly acknowledged. I also declare that it has not been previously
and concurrently submitted for any other degree or award at UTAR or other institutions.

Signature : _________________________

Name : Tiang Hock Yu

ID No : 15UEB05212

Date : 15/4/2019
iii

APPROVAL FOR SUBMISSION

I certify that this project report entitled SIMULATION OF CATALYTIC COMBUSTION


OF AUTOMOTIVE EXHAUST GAS IN MONOLITHIC REACTOR was prepared by
TIANG HOCK YU has met the required standard for submission in partial fulfilment of the
requirements for the award of Bachelor of Engineering (Honours.) Chemical Engineering at
Universiti Tunku Abdul Rahman.

Approved by,

Signature : _________________________

Supervisor : Dr Yap Yeow Hong

Date : 15/4/2019
iv

ACKNOWLEDGEMENTS

I would like to express my appreciation and gratitude to my FYP supervisor, Dr. Yap Yeow
Hong for his patience and wisdom in guiding me throughout the period. Without his meticulous
guidance, this project would have been a lot more difficult and time consuming. Also, I would
like to thank my dearest brother who provided invaluable comments and support when I am
doing my research. Lastly, to my friends who helped and provided encouragement throughout
the preparation of this report.
v

SIMULATION OF CATALYTIC COMBUSTION OF AUTOMOTIVE EXHAUST


GAS IN MONOLITHIC REACTOR

ABSTRACT

Catalytic converters are common fixture in almost all motor vehicle on the road for their ability
to treat the exhaust emission. Modern research and developments on catalytic converter employ
the use of computer models to help simulate various aspects of catalytic combustion that take
place in the converters, as this will provide valuable insight to develop an efficient product.
Typically, such simulation will involve modelling the ignition-extinction phenomena as well
as predicting emission for drive cycle. In this project, a simulation based on Island Model will
be developed and simulated and this will be compared with other established kinetic models.
Simulation will be done on a 5mm thin slice Pt/γ-Al2O3 Diesel Oxidation Catalyst (DOC) and
benchmarked with the results by Ye. et. al. (2012). The transient 1D-gas phase, 1D-solid phase
single channel model considers the mass and energy transport as well as the change of the
adsorbed species on the catalyst surface. The simulation results have shown that the Herz and
Marin basic model and Salomons’ compressed oxygen model gave the best similarity in terms
of both conversion and hysteresis characteristics while the Island Model provided a good
agreement in terms of conversion but not hysteresis characteristics. Fine-tuning of the effective
diffusivity of the species of the Island Model is done and its effect on the hysteresis curve is
shown to be significant. Effective diffusivity and along with other kinetic parameters should
be optimized to achieve a better agreement between experimental and simulated results.
vi

TABLE OF CONTENTS

DECLARATION ii

APPROVAL FOR SUBMISSION iii

ACKNOWLEDGEMENTS iv

ABSTRACT v

TABLE OF CONTENTS vi

LIST OF TABLES viii

LIST OF FIGURES ix

LIST OF SYMBOLS/ABBREVIATIONS xiv

CHAPTER 1 1

INTRODUCTION
1.1 Background 1
1.2 Problem Statement 2
1.3 Aims and Objectives 2
1.4 Scope of Study 3

CHAPTER 2 4

LITERATURE REVIEW
2.1 Diesel Oxidation Catalysts (DOCs) 4
2.2 Hysteresis 6
2.3 Global and Mechanistic Kinetics 21
2.4 Modifications to the Basic Model 27
2.5 Adsorption and Desorption Kinetics 30
vii

2.6 Hydrocarbon Oxidation 35


2.7 Programming Language Utilized 36
2.8 Numerical methods 36
2.9 Experimental data available to perform the simulation 39

CHAPTER 3 40

METHODOLOGY
3.1 Single Monolith Channel 40
3.2 Heat and Mass Transfer Equations 40
3.3 Adsorption and Desorption Kinetics 43
3.4 Reaction Models Investigated 45
3.5 Hydrocarbon Oxidation 49
3.6 Coding 50

CHAPTER 4 63

RESULTS AND DISCUSSION

4.1 Hysteresis (Inlet gas temperature) 63


4.2 Hysteresis (Solid temperature) 69
4.3 Surface Coverage 74
4.4 Temperature Profile 74

CHAPTER 5 95

CONCLUSION AND RECOMMENDATIONS


5.1 Conclusion 95
5.2 Recommendation 96

APPENDICES 97

REFERENCES 97
viii

LIST OF TABLES

TABLE TITLE PAGE


Table 2.2.1: Properties of the bimetallic Pt-Pd catalyst. (Daneshvar et al., 2017) 21

Table 3.6.1: Letter Designation for each Species 51

Table 4.1.1: Summary of Hysteresis Characteristics from the simulations done (Gas Inlet

Temperature) 68

Table 4.2.1: Summary of Hysteresis Characteristics on the simulation done (Solid

Temperature at 3mm) 73

Table 4.5.1: Summary of Hysteresis Characteristics after tuning kd 93


ix

LIST OF FIGURES

FIGURE TITLE PAGE


Figure 2.1.1: Diagram of a DOC (adapted from technopow.com) 5

Figure 2.1.2: Schematic Diagram of the Interior of the DOC (adapted from microscopy.cz) 5

Figure 2.2.1: Clockwise Hysteresis Observed When Inlet CO Concentration is Varied.

(Hlaváček and Votruba, 1979) 7

Figure 2.2.2: A Series of Ignition and Extinction Curves Obtained by Varying Gas Inlet

Temperature in Cu/Al2O3 and Pt/Al2O3 Tubular Packed Bed Reactors. (Hlaváček and

Votruba, 1979) 8

Figure 2.2.3: Extinction Curves Due to Increasing CO Concentration of the Inlet Gas.(Wicke,

Kummann and Schiefler, 1980) 8

Figure 2.2.4: Hysteresis Is Still Attained When Heat of Reaction Is Set to 0. (Raj, Harold and

Balakotaiah, 2015) 10

Figure 2.2.5: Simulated Curves Using the Global Kinetics by Voltz et al. (Salomons et al.,

2007) 11

Figure 2.2.6: A schematic diagram of cluster formation(Wicke, Kummann and Schiefler,

1980) 13

Figure 2.2.7: Development of Clusters Based on the Previous Scheme. (Wicke, Kummann

and Schiefler, 1980) 13

Figure 2.2.8: Ignition Curves Obtained by Varying Ramp Up Times (Ye et al., 2012) 15
x

Figure 2.2.9: Effect of ramp up rate on width of hysteresis. (Raj, Harold and Balakotaiah,

2015) 16

Figure 2.2.10: The effect of inlet CO concentration on the hysteresis phenomena and the rate

of oxidation of CO (Ye et al., 2012) 17

Figure 2.2.11: The effect of inlet CO concentration on the hysteresis phenomena and the

conversion of CO (Ye et al., 2012) 17

Figure 2.2.12: Experimental Evidence on the Effect of Inlet CO Concentration on the

Hysteresis Phenomena (Salomons et al., 2007) 18

Figure 2.2.13: Psedo Steady State Experiments on Ignition Curves Produced by Different CO

Concentrations (Ye et al., 2012) 19

Figure 2.2.14: Snapshots of the surface during reaction at different CO concentrations. (a)

Xeo = 0.39, near oxygen poisoning transition. (b) Xco = 0.50, in steady reaction region. (c)

Xco = 0.52, near CO poisoning. (black dots = O and white dots = CO) (Ehsasi et al., 1989) 19

Figure 2.2.15: Ignition Points as a Function of Pt:Pd Ratio. (Daneshvar et al., 2017) 20

Figure 2.3.1: 𝒓𝒄𝒐𝟐, and Surface Coverage of O and CO for a CO beam Impinging on an

Oxygen Saturated Surface on Pd (111) at T=374K.(Engel and Ertl, 1978) 22

Figure 2.5.1: Relative Sticking Coefficient as a Function of CO Coverage. (Ertl, Neumann

and Streit, 1977) 32

Figure 2.5.2: Initial Sticking Coefficient as a Function to the Substrate Temperature. (Engel,

1978) 34

Figure 2.8.1: Stencil for the Explicit Method (adapted from Wikipedia.com). 37

Figure 2.7.2: Stencil for the Implicit Scheme (adapted from Wikipedia.com). 38

Figure 2.7.3: Stencil for the Crank-Nicholson Method (adapted from Wikipedia.com). 38

Figure 3.6.1: Coding for Grid Discretisation 50

Figure 3.6.2: Coding for Monolith Data 50


xi

Figure 3.6.3: Coding for Molecular Properties and other constants 51

Figure 3.6.4: Coding for Variable Initialisation 52

Figure 3.6.5: Coding for Dynamic Parameters (Gas Phase Properties) 52

Figure 3.6.6: Coding for Dynamic Constants (Solid Phase Propeties) 53

Figure 3.6.7: Coding for Dynamic Constants (Heat and Mass Transfer Coefficient) 53

Figure 3.6.8: Coding for Dynamic Constants (Diffusivity Parameters) 54

Figure 3.6.9: Temperature against time plot from Ye et . al.(2012) 55

Figure 3.6.10: Coding for Setting up the Inlet Temperature Profile 55

Figure 3.6.11: Coding to start the model function 56

Figure 3.6.12: Coding for the declaration the differential variables 56

Figure 3.6.13: Coding for declaration of parameters of the reaction model. 57

Figure 3.6.14: Coding for the setting up of rate equations 57

Figure 3.6.15: Coding for Setting up the Heat and Mass Balances 58

Figure 3.6.16: Coding to close the model function and return the calculated differential values

59

Figure 3.6.17: Coding for the stage loop 60

Figure 3.6.18: Coding for the solution loop. 60

Figure 3.6.19: Coding for Data Storage 61

Figure 3.6.20: Coding for the compilation of data of all stages and its storage. 61

Figure 3.6.21: Coding for Graph Plotting 62

Figure 4.1.1: Conversion against inlet temperature plot from the experimental results of Ye

et. al.(Ye et al., 2012) 64

Figure 4.1.2: Conversion against inlet temperature plot for Voltz Model 64

Figure 4.1.3: Conversion against inlet temperature plot for Herz and Marin Basic Model 65
xii

Figure 4.1.4: Conversion against inlet temperature plot for Salomons' Compressed Oxygen

Model 65

Figure 4.1.5: Conversion against inlet temperature plot for CO Island Model 66

Figure 4.1.6: Conversion against inlet temperature plot for O Island Model 66

Figure 4.1.7: Conversion against inlet temperature plot for CO-O Island Model 67

Figure 4.1.8: Conversion against inlet temperature plot for Dumont Postulation 67

Figure 4.2.1: Conversion against solid temperature plot from the experimental results of Ye

et. al.(Ye et al., 2012) 69

Figure 4.2.2: Conversion against solid temperature plot for Voltz Model 69

Figure 4.2.3: Conversion against solid temperature plot for Herz and Marin Basic Model 70

Figure 4.2.4: Conversion against solid temperature plot for Salomons' Compressed Oxygen

Model 70

Figure 4.2.5: Conversion against solid temperature plot for CO Island Model 71

Figure 4.2.6: Conversion against solid temperature plot for O Island Model 71

Figure 4.2.7: Conversion against solid temperature plot for CO-O Island Model 72

Figure 4.2.8: Conversion against solid temperature plot for Dumont Postulation 72

Figure 4.3.1: Temperature plot for Voltz Model 74

Figure 4.3.2: Temperature plot for Herz and Marin Basic Model 75

Figure 4.3.3: Temperature plot for Salomons' Compressed Oxygen Model 75

Figure 4.3.4: Temperature plot for CO Island Model 76

Figure 4.3.5: Temperature plot for O Island Model 76

Figure 4.3.6: Temperature plot for CO and O Island Model 77

Figure 4.3.7: Temperature plot for Dumont Postulation 77

Figure 4.4.1: Fractional Coverage of CO against time for Herz and Marin Basic Model 78

Figure 4.4.2: Fractional Coverage of O against time for Herz and Marin Basic Model 79
xiii

Figure 4.4.3: Fractional Coverage of Vacant Sites against time for Herz and Marin Basic

Model 79

Figure 4.4.4: Fractional Coverage of CO against time for Salomons' Compressed Oxygen

Model 80

Figure 4.4.5: Fractional Coverage of O against time for Salomons' Compressed Oxygen

Model 80

Figure 4.4.6: Fractional Coverage of Compressed Oxygen against time for Salomons'

Compressed Oxygen Model 81

Figure 4.4.7: Fractional Coverage of Vacant Sites against time for Salomons' Compressed

Oxygen Model 81

Figure 4.4.8: Fractional Coverage of CO against time for CO Island Model 82

Figure 4.4.9: Fractional Coverage of O against time for CO Island Model 82

Figure 4.4.10: Fractional Coverage of Vacant Site against time for CO Island Model 83

Figure 4.4.11: Fractional Coverage of CO against time for O Island Model 84

Figure 4.4.12: Fractional Coverage of O against time for O Island Model 84

Figure 4.4.13: Fractional Coverage of Vacant Site against time for O Island Model 85

Figure 4.4.14: Fractional Coverage of CO against time for CO-O Island Model 86

Figure 4.4.15: Fractional Coverage of O against time for CO-O Island Model 86

Figure 4.4.16: Fractional Coverage of Vacant Sites against time for CO-O Island Model 87

Figure 4.4.17: Fractional Coverage of CO against time for Dumont Postulation 88

Figure 4.4.18: Fractional Coverage of O against time for Dumont Postulation 88

Figure 4.4.19: Fractional Coverage of Vacant Sites against time for Dumont Postulation 89
xiv

LIST OF SYMBOLS/ABBREVIATIONS

𝑧 Channel Length 𝑚𝑚
𝐷 Channel Width (without washcoat) 𝑚𝑚
𝑊 Washcoat Thickness 𝑚𝑚
𝐶 Average Washcoat Loading 𝑔
𝐷𝑂𝐶
𝜌 Density of Washcoat 𝑘𝑔
𝑚
𝐶 , Heat Capacity of Washcoat 𝐽
𝑘𝑔𝐾
𝑘 Heat transfer coefficient of Washcoat 𝑊
𝑚𝐾
𝑆 BET Surface Area 𝑚
𝑔
𝐷 Average Pore Diameter 𝑛𝑚
𝜀 Porosity

𝜏 Tortuosity

𝛿 Constrictivity

𝑣 Inlet air velocity 𝑚


𝑠
𝜌 Density of air 𝑘𝑔
𝑚
𝐶 Specific Mass Heat Capacity 𝐽
𝑘𝑔𝐾
𝑘 Heat Transfer Coefficient 𝑊
𝑚𝐾
xv

𝑅 Ideal Gas Constant 𝐽


𝑚𝑜𝑙 . 𝐾
𝑘 Boltzmann’s Constant 𝐽
𝐾
𝑀𝑟 Molar mass of species i 𝑔
𝑚𝑜𝑙
𝑇 Fluid temperature 𝐾

𝑇 Surface temperature 𝐾

𝑝 Partial Pressure of species i 𝑃𝑎

𝐷 , Diffusivity of species i 𝑐𝑚
𝑠
𝑟, Rate of adsorption of species i

𝑟, Rate of desorption of species i

𝑟 Rate of reaction of species i

𝑆 , Initial Sticking Coefficient of species i

Γ No of mol of metal atoms per surface area 𝑚𝑜𝑙


𝑚
𝐶 Gas phase concentration of species i 𝑚𝑜𝑙
𝑚
𝑌 Gas phase mol fraction of species i

𝜃 Surface concentration of vacant sites

𝜃 Surface concentration of species i

𝐴 Pre-exponential factor 1
𝑠
xvi

𝐸 Activation energy for adsorption 𝑘𝐽


𝑚𝑜𝑙
𝐸 Activation energy for desorption 𝑘𝐽
𝑚𝑜𝑙
𝐸 Activation energy for surface reaction 𝑘𝐽
𝑚𝑜𝑙
𝛽 Rate of change of activation energy per surface concentration of 𝑘𝐽
species i

𝑘 Rate constant for CO oxidation for the Voltz Model 1


𝑠
𝑘 Rate constant forC3H6 oxidation for the Voltz Model 1
𝑠
𝑘 Inhibition term constants j=1,2,3 𝑚
𝑚𝑜𝑙
𝑘 Rate constant for surface reaction 1
𝑠

𝑘 Rate constant for OO formation for the Compressed Oxygen Model 𝑚𝑜𝑙
𝑚 𝑠
𝑘 Rate constant for surface reaction of CO and OO for the Compressed 𝑚𝑜𝑙
Oxygen Model 𝑚 𝑠

𝑘 Rate constant for desorption of OO for the Compressed Oxygen 𝑚𝑜𝑙


Model 𝑚 𝑠

𝑖 CO, O2, CO2, OO, C3H6


1

CHAPTER 1

1 INTRODUCTION

1.1 Background
Diesel engines still drive most of the heavy-duty vehicles because it can produce more power
per unit of fuel used when compared with its gasoline counterpart. Moreover, diesel fuel
requires less processing cost and are generally cheaper. Lastly, the average lifetime of a diesel
engine is almost double of a gasoline engine.
However, diesel engines emits much more soot, unburnt hydrocarbons and carbon
monoxide. This is due to the heavy nature of the diesel fuels which makes incomplete
combustion a bigger possibility. Moreover, the fuel is injected just before the power stroke
causing incomplete combustion due to the lack of oxygen. These products of incomplete
combustion are harmful to both humans and the environment.
As a result, regulations were drafted and implemented by the government to control
emission of harmful products by fossil fuelled vehicles and engines. Therefore, the pressure is
building on the automobile manufacturers to reduce the emission from diesel vehicles. The
most popular and cost effective ways is to introduce a catalytic converter at the exhaust outlet.
These converters are able to convert products of incomplete combustion into less harmful
substances which are water and carbon dioxide. The improvement of catalytic converters or
diesel oxidation catalysts (DOCs) are of a huge interest over the past few decades.
2

1.2 Problem Statement


Simulation is important to the automotive industry as it can provide a low cost and effective
way to improve the performance of their catalyst instead of undertaking complicated
experiments, which can be very costly and time consuming due to the setup of the engine test
cell and its instrumentation. Simulation of catalytic combustion in monolithic reactors is
typically done by formulating mathematical models of heat and mass transfer and the reaction
kinetics together with the appropriate assumptions. The most studied reaction is the oxidation
of carbon monoxide since it is the dominant reaction in catalytic converters.
Generally, there are two approaches that we can employ to model the reaction kinetics
of the reaction. Most researchers follow the model first conceived by Voltz et. al. (1973), a
global reaction kinetics based on the Langmuir-Hinshelwood model of reaction. Another
approach for the reaction modelling is the use of mechanistic approach which utilises the
surface species concentration instead of bulk gas phase concentration. According to Salomons
et. al. (2008), this is a better way to model carbon monoxide oxidation since it takes into
account of the compressed oxygen effect unlike the global kinetics.
Numerous attempts have been made and published however researchers are yet to find
a common ground regarding the best model to simulate catalytic combustion. (Salomons et al.,
2007; Dadi, Luss and Balakotaiah, 2016; Daneshvar et al., 2017). However, most of them have
their own shortcomings when it comes to producing the important details of the catalytic
combustion such as the ignition-extinction hysteresis and the change in surface species during
the ignition-extinction. It is still rare that a simulation could give an output which fits the
experimentally observed results.

1.3 Aims and Objectives


This project aims to simulate the simulated curves of CO using the Scientific Python
Development Environment (Spyder) module. The following objectives are set:
a) To construct a 1D gas phase-1D solid phase mathematical model which simulates
catalytic oxidation of a 5mm thin slice diesel oxidation catalyst.
b) To compare the simulation result of different reaction models proposed.
c) To fine-tune the kinetic parameters and study its effect on the catalytic behaviour of
DOC.
3

1.4 Scope of Study

Chapter 2: Literature review


 Basic knowledge on the structure and principles behind DOC.
 Factors that result in the hysteresis phenomena.
 Reaction parameters that affect the hysteresis curve.
 Different kinds of models and kinetics that had been postulated.
 Fine tuning strategies
 Introduction to numerical methods.
 The programming language chosen.

Chapter 3: Methodology
 Equations for heat and mass transfer
 The models that are selected and their respective formulations and equations
 Kinetic parameters that are selected.
 Method used to solve the monolith system
 Application of Spyder in solving the monolith system
 Fine tuning strategies

Chapter 4: Results and Discussion


 Analysis of the simulated results.
 Discussing and justifying the discrepancy between simulated model and experimental
data.
 Determining and justifying the best model for simulation.

Chapter 5
 Conclusion of research
 Recommendation for the improvement of the simulation
4

CHAPTER 2

1 LITERATURE REVIEW

2.1 Diesel Oxidation Catalysts (DOCs)


2.1.1 Background

DOC (Diesel Oxidation Catalysts) is one of the many parts which makes up the complex
emission control system of a diesel powered vehicle. They usually come in the form of
monolithic catalytic reactors, which is a form of packed bed reactor which consists of many
long thin hollow channels. It is made up of cordierite/γ-Al3O2 which acts as supports the
catalysts and noble metals such as platinum and palladium. A plethora of improvements and
innovations were made to the catalyst composition, where 2 or 3 metals are used as the catalyst
(Nibbelke et al., 1997).
5

2.1.2 Structure of DOCs

Figure 2.1.1: Diagram of a DOC (adapted from technopow.com)

Most DOCs are based on ceramic “honeycomb” monoliths even though sometimes metallic
monoliths can be seen in the market. Ceramic monoliths are better in withstanding thermal
expansion. However, its low porosity makes it undesirable for catalyst dispersion. A remedy
to this is to introduce a layer of porous material on to the surface of the monolith channel which
is normally known as washcoat.

Figure 2.1.2: Schematic Diagram of the Interior of the DOC (adapted from
microscopy.cz)
6

Washcoat acts as a catalyst support, improving its surface area, and thermal stability.
Alumina is a very common washcoat material as it fulfills all the criteria for a good support.
Alumina is porous, providing more surface area of reaction and also being thermally stable
even at high temperatures. Other supports such as zeolites and zirconia are also used. Zeolites
generally improve the cold start performance of the DOC. It aids by trapping the unburnt
hydrocarbons at lower temperatures where oxidation is not possible. When the desired
temperature is reached, the hydrocarbons are then released and oxidized by the catalyst.
Zirconia also lowers the ignition temperature of the DOC.(Russell and Epling, 2011)

2.1.3 Modelling DOCs through simulation


The research and development for DOCs based on physical experiments can be quite costly.
An alternative way is to study the effects through computer simulations that can be achieved
by using the correct mathematical model. Derivations of such models are very complex and
rigorous. Numerous attempts and debates were made throughout the past half century on the
true kinetics and mechanisms of carbon monoxide oxidation. Langmuir (1952) first postulated
the rate of reaction. Later on, researchers like, Ertl and Engel (1979), and Herz and Marin
(1983), modified the Langmuir model in an attempt to improve the modelling capability of the
rate law. The papers by these researchers laid the foundation which then enabled many more
groundbreaking discoveries to be made. However, researchers still failed to provide a model
which can provide an accurate description of the DOCs during transient state operation
particularly when hysteresis of the ignition-extinction phase is concerned. (Salomons et al.,
2007)

2.2 Hysteresis

2.2.1 Origins and evidence


Hysteresis of CO oxidation has been a subject of study over the past few decades. One of the
very first comprehensive review on the subject is done by Hlavacek and Votruba (1979), and
Razon and Schmitz (1986). They compiled past experiments covering a wide range of
conditions, from UHV to near atmospheric, 50 to 450℃, packed bed reactors to Pt wires.
Almost all experiments reported oscillations and multiple steady states, for which the latter are
one of the reasons for hysteresis (Hlaváček and Votruba, 1979; Razon and Schmitz, 1986).
Therefore, it is imminent that we need to consider such effects on DOCs. Most reports
concluded that hysteresis is a result of heat and mass limitations within the reactors (either hot
spots or defects). However, there are reports that hysteresis can be observed even in thin sliced
7

DOC and they showed that hysteresis is a phenomena which is mostly related to the reaction
mechanism and is an intrinsic behavior rather than a result of heat and mass limitations. (Ye et
al., 2012). Hysteresis can be observed by either changing the gas inlet temperature or inlet CO
concentration.

Figure 2.2.1: Clockwise Hysteresis Observed When Inlet CO Concentration is Varied.


(Hlaváček and Votruba, 1979)
8

Figure 2.2.2: A Series of Ignition and Extinction Curves Obtained by Varying Gas Inlet
Temperature in Cu/Al2O3 and Pt/Al2O3 Tubular Packed Bed Reactors. (Hlaváček and
Votruba, 1979)

Figure 2.2.3: Extinction Curves Due to Increasing CO Concentration of the Inlet


Gas.(Wicke, Kummann and Schiefler, 1980)
9

2.2.2 Steady State and Dynamic Hysteresis

In general, there are two types of hysteresis: steady state and dynamic. (Dadi, Luss and
Balakotaiah, 2016)

Steady state hysteresis is well documented in literatures ever since the 1960s.(Razon
and Schmitz, 1986) Researchers noticed that CO oxidation can be at two stable steady states:
upper and lower, categorized based on the rate of reaction that they are having. Instead of using
hysteresis, they used the term steady state multiplicity. Most attributed this to the mathematical
nature of the heat and mass transfer solutions, which will be covered in detail on the next
section.

Dynamic hysteresis became a point of interest when researchers are investigating the
performance of catalytic converters. They noticed that during a cold-start, the conversion of
CO is low. However, when a certain inlet gas temperature is achieved, the ignition temperature,
conversion rises drastically, reaching to its maximum conversion shortly after. Then, when the
engine slows down, causing a drop in inlet gas temperature, the rate of reaction will not follow
the same path as before. Instead, the conversion will still remain until a certain temperature is
reached, the extinction temperature, conversion will often plummet or drop off at a fast rate.
The conversion will eventually plateau off at a low conversion state.

Notice that how both hysteresis are quite similar as two different states are achieved
(low and high conversion). The main difference is that steady state hysteresis is observed
during steady state experiments where the system is given sufficient time to reach steady state
while dynamic hysteresis is observed when the system is not in steady state or in a transient
state where inlet gas temperature or inlet CO concentration is varied with time and the readings
are taken instantaneously.

2.2.3 Mechanisms behind Hysteresis

a) Heat and Mass Transfer Limitations

One of the earliest reasons that are proposed by researchers is the heat and mass transfer
limitations within the reactor. They looked into the partial differential equations which governs
the heat and mass transfer of the reactor and found out that there are certain mathematical
criteria (usually involves the external parameters of the reaction) that is needed so that state
multiplicity or hysteresis to occur. (Hlaváček and Votruba, 1979)
10

Such analysis, stability analysis, dominated the literature on this subject during 1960s
and 1970s. Weisz and Hicks (1962) first mentioned the possibility of multiple steady states in
their classic paper regarding heat and mass transfer of packed bed reactors. Their works then
provided a good foundations for researchers later on. Aris, Schimtz and Ray then provided
papers that further refines the theory.(Hlaváček and Votruba, 1979)

Fast forward to this decade, in recent studies published by the researchers of the
University of Houston, they performed simulation and it provides computational evidence that
hysteresis occur due to the delay in temperatures between the feed inlet temperature and solid
(catalyst) temperature. (Raj, Harold and Balakotaiah, 2015; Dadi, Luss and Balakotaiah, 2016)
They established that the cause of the commonly observed extinction curve is due to the heat
transfer limitations within the reactor not the exothermic nature of the reaction itself. In their
simulation, they set the heat of reaction as zero and still observe the hysteresis phenomena.
(Raj, Harold and Balakotaiah, 2015; Dadi, Luss and Balakotaiah, 2016) Raj et al explained that
when decreasing the inlet gas feed, conversion is still maintained because of the heat transfer
limitation between the monolith and the inlet gas. The surface temperature is still high enough
for reactions to occur. Extinction will eventually occur when surface temperature is low enough.

Figure 2.2.4: Hysteresis Is Still Attained When Heat of Reaction Is Set to 0. (Raj,
Harold and Balakotaiah, 2015)

However, they did not try to fit their simulations with any available ignition-extinction
curve that is obtained from experiment. Salomons et al (2007) attempted that and stated that
11

using global kinetics by Voltz could not provide a good fit for the extinction curve. They
proposed the use of mechanistic models to model hysteresis instead. (Salomons et al., 2007)

Figure 2.2.5: Simulated Curves Using the Global Kinetics by Voltz et al. (Salomons et
al., 2007)

b) The asymmetrical behavior of CO and O2 adsorption.(Engel et al., 1979)

Hysteresis is still being reported on experiments that are performed with the elimination of
mass transfer limitations, and isothermal conditions or on Pt wires, where heat and mass
transfer limitations is nonexistent.(Eigenberger, 1978; Razon and Schmitz, 1986) Therefore, it
is important that we take into consideration the effect of each elementary step has.

It is long known that CO adsorption is not inhibited by adsorbed oxygen atoms but not
vice versa. Also, at low temperatures, CO is the dominant species that is adsorbed while O2
requires temperature at the excess of 200℃ to adsorb. Finally, O2 desorption is insignificant at
temperatures below 300℃. These behaviors could justify the hysteresis phenomena.

At low gas inlet temperatures, CO is the primary species that is adsorbed but O2
adsorption is not expected because CO poisons and inhibits O2 adsorption. As the temperature
increases, CO starts to desorb, creating vacancies for O2 to adsorb. This would lead to a LH
12

mechanism reaction and CO2 is released, creating more vacancies for O2 to adsorb. At high
temperatures, the adsorption of O2 is more favorable, causing O atoms to be the dominant
species on the surface.

However, adsorbed O species does not inhibit the adsorption of CO. CO is able to
squeeze in between the spaces of the adsorbed O atoms and hence compress the O atoms around
it. This effect of compression will lower the O-O bond strength, causing the rate of reaction to
be higher.(Engel et al., 1979) All of these factors would contribute to the acceleration of the
rate of reaction causing an ignition phenomena where at certain temperature (the ignition
temperature) a sudden increase in conversion can be seen.

When the gas inlet temperature is reduced, the conversion of CO will maintain at its
maximum until a temperature way lower than the light-off temperature. This is because of the
exothermic nature of the oxidation reaction and also the high temperature favored O2
adsorption and CO desorption. The exothermic nature of the reaction enabled the surface
temperature to be sufficiently high for both CO desorption and O2 adsorption to occur and
hence maintaining the rate of reaction. (Casapu et al., 2017)

Lastly, at the temperature where CO desorption is less likely to occur, adsorbed CO


species will start to accumulate and inhibit O2 adsorption, which reduces the rate of reaction.
Also, O2 adsorption is less favored due to lower surface temperature. Eventually, these factors
will cause a sharp decrease in rate of reaction before plateauing off to near zero rates at low
temperatures.
13

c) The island effect

The formation of CO and possibly O islands are reported by various literatures. Engel and Ertl
(1979) first provided possibility of it by introducing the concept of cooperative adsorption.

Figure 2.2.6: A schematic diagram of cluster formation(Wicke, Kummann and


Schiefler, 1980)

Figure 2.2.7: Development of Clusters Based on the Previous Scheme. (Wicke,


Kummann and Schiefler, 1980)

Wicke et. al. (1980) demonstrated through computer simulation that the island
formation is possible under the assumption that both CO and O2 have the same probability to
adsorb, and the surface reaction is a LH mechanism. The left figure of Figure 2.2.7 shows the
14

initial distribution of two species which is even and random. The right figure shows the
formation of clusters (or islands) after 104 reaction steps.

Haaland and Williams (1982) performed experiments using IR spectroscopy in an


attempt to explain such effects. Their work provided experimental proof for the existence of
such islands.

Wicke et al (1980) provided a detailed explanation on how islands of CO will cause the
light-off phenomena. As clusters of CO islands grow because of the low temperature favored
CO adsorption, the chemisorbed bonding between CO and Pt active sites will decrease. This is
a well-known effect that can be proven through calorimetric and equilibrium measurements,
and also infrared spectroscopy. This is due to the increased competition of the CO molecules
for the electrons of the Pt metal. As a result, it is easier for CO molecules that is attached to an
island to desorb than the isolated ones.

The increase in the island size and also reduced reaction rate of CO oxidation causes
the rate of CO desorption to increase.(Wicke, Kummann and Schiefler, 1980) It will then reach
a point where O2 adsorption becomes more favorable due to the increasing availability of two
vacant adjacent sites or the temperature is high enough such that O2 adsorption is high along
with rate of CO desorption. The latter can also happen in experiments where CO concentration
is varied due to the exothermic nature of the reaction. As rate of oxidation increases, the
temperature of the surface rises, which then leads to ignition as O2 adsorption and CO
desorption increases.

Haaland and Williams (1982) also mentioned that the increasing CO island size due to
increased CO adsorption would cause a drop in rate of reaction during extinction as the
available species for reaction at the domains are largely reduced and also the inhibition effect
of CO for O2 adsorption.

d) Oxidation of Platinum Nanoparticles

Inverse hysteresis is observed on Pt-nanoparticles that is smaller than 2nm in diameter.(Casapu


et al., 2017) The reduction of Platinum Oxides is more difficult for particles of that size than
the bigger one. Therefore, this would cause a drop in conversion and hence extinction
temperatures that are higher than the light-off temperature.
15

2.2.4 Factors that will affect the behavior of Hysteresis

The width of hysteresis, and its ignition and extinction point is affected by a number of factors
such as ramp up and cool down rates of inlet gas, composition and flow rate of inlet gas, size
of Pt nanoparticles, washcoat diffusivity and last but not least, the degradation or aging of
catalysts.

a) Ramp up and Cooling rates of Inlet Gas

Increasing the ramp up rate will result in higher ignition points as it is apparent from both
Figure 2.2.8 and Figure 2.2.9. (Ye et al., 2012; Raj, Harold and Balakotaiah, 2015)

Figure 2.2.8: Ignition Curves Obtained by Varying Ramp Up Times (Ye et al., 2012)
16

Figure 2.2.9: Effect of ramp up rate on width of hysteresis. (Raj, Harold and
Balakotaiah, 2015)

This can be explained through heat transfer limitations within the monoliths. As ramp
up rate increases, the rate of temperature change of the inlet gas increases too, However, this
effect does not translate into an equal rise in rate of change of the catalyst surface temperature.
As a result, a delay can be seen in reaching ignition point. For example, when the temperature
of inlet gas is at 200℃, the surface temperature might still be at 130℃. (Raj, Harold and
Balakotaiah, 2015)

From Figure 2.2.8, it can be seen that as ramp up rate is decreased, ignition point
decreases. Also, if ramp up rate is slow enough, we could attain steady state at every point of
the measurement. Low ramp up rates will actually allow time for heat transfer to take place and
hence the delay in between gas inlet temperature and surface temperature would be less
significant.

The cooling rate would also have some significant effect on the extinction point. As
cooling rate is increased, the extinction point decreases. Again this is due to the heat transfer
limitation of within the monolith itself. The gas inlet temperature would be a lot cooler than
the surface temperature.

The effect of ramp up and cooling rates have no effect on the real ignition and extinction
points where it is dependent on the surface temperature. The temperatures obtained due to heat
transfer limitations are actually the apparent ignition and extinction points. Therefore, to
improve the true ignition-extinction characteristic of the catalysts, this would be a non-factor.
17

b) Concentration of CO in the Inlet Gas

Figure 2.2.10: The effect of inlet CO concentration on the hysteresis phenomena and the
rate of oxidation of CO (Ye et al., 2012)

Figure 2.2.11: The effect of inlet CO concentration on the hysteresis phenomena and the
conversion of CO (Ye et al., 2012)
18

Figure 2.2.12: Experimental Evidence on the Effect of Inlet CO Concentration on the


Hysteresis Phenomena (Salomons et al., 2007)

Hysteresis is still observed at different CO concentrations although the ignition and extinction
points did not vary significantly. From Figure 2.2.11, the conversion of the DOC shows a slight
decrease when the inlet CO concentration is reeduced while it is vice versa for the rate of
oxidation of CO.

However, looking into Figure 2.2.12, increasing the CO concentration shifts the
hysteresis curve to the left where a significantly higher ignition-extinction points are obtained.
This is different from what Ye et al (2012) observed in their experiments (difference in
conversion and reaction rates). The main reason behind this difference is the ramp up rates used
by both researchers. Salomons et al used a lower ramp up rate when compared with Ye et al
(0.133K/s vs 1.556 K/s). Ye et al described that using low ramp up rates would result in the
experiment being at a pseudo steady state rather than transient state.
19

Figure 2.2.13: Psedo Steady State Experiments on Ignition Curves Produced by


Different CO Concentrations (Ye et al., 2012)

Indeed, looking at Ye et. al.’s pseudo steady state results from Figure 2.2.13, we can
confirm that the shift produced in Salomon’s experiment is performed at a pseudo steady state
condition since both results give identical conversion (both at 100%). Increasing CO
concentration during a steady state condtion will result in more CO surface coverage
demonstrated by both experiments and simulation which is evident in Figure 2.2.14. (Ehsasi et
al., 1989) As a result, a higher temperature is needed so that sufficient amount of CO is
desorbed for ignition to occur.

Figure 2.2.14: Snapshots of the surface during catalytic CO oxidation at different CO


concentrations. (a) Xeo = 0.39, near oxygen poisoning transition. (b) Xco = 0.50, in
steady reaction region. (c) Xco = 0.52, near CO poisoning. (black dots = O and white
dots = CO) (Ehsasi et al., 1989)
20

c) Monolith properties

The properties of the monolith, especially washcoat properties, are rather significant on the
hysteresis behavior of the DOC. Studies have been conducted on the effect of substrate material
on the hysteresis phenomena. They found out that the ignition point would be lower if a metallic
monolith is used. (Gundlapally and Balakotaiah, 2013)

The use of bimetallic catalyst would affect the ignition point. The actual mechanism
involved is very complex and it is beyond the scope of this project. Nevertheless, we will
include the results of the simulation conducted by Daneshvar et. al. (2017) and give a brief
explanation on it.

Figure 2.2.15: Ignition Points as a Function of Pt:Pd Ratio. (Daneshvar et al., 2017)

From Figure 2.2.15, we can observe a sharp decrease in ignition temperature as we


increase the ratio from 1:0 to 3:1. Daneshvar et al stated that this trend can be related to the
dispersion of the catalyst as we alter the Pt:Pt ratio, which is showed in Table 2.2.1. It can be
inferred that greater dispersion yield lower light-off temperatures(Daneshvar et al., 2017)
21

Table 2.2.1: Properties of the bimetallic Pt-Pd catalyst. (Daneshvar et al., 2017)
Pt/Pd Pt Pd Total Pt loading Pd loading Dispersion Particle size
Mol ratio Wt% Wt% Wt% g/ft3 g/ft3 % nm
Pt 1 0 1 27.75 0 5.8 19.4
3:1 0.75 0.13 0.88 20.81 3.78 15.6 7.2
1:1 0.5 0.27 0.77 13.87 7.57 21.7 5.2
1:3 0.25 0.41 0.66 6.93 11.35 23.5 4.8
Pd 0 0.54 0.54 0 15.13 26.4 4.3

2.3 Global and Mechanistic Kinetics

2.3.1 Global Kinetics

Global kinetics lumped every step into one single equation by making several
assumptions.(Salomons et al., 2007) It usually takes into account the gas phase concentration
species. Global kinetics do not take into account the surface concentration of the adsorbed
species. Two groups of researchers went on and suggested simplifications of the basic
elementary steps with the aid of experimental evidence. The two proposed mechanisms are as
below:

1. Langmuir-Hinshelwood
The mechanism proposed by Langmuir involves only adsorbed CO and O2:

𝐶𝑂 + 𝜃 → 𝐶𝑂. 𝜃 Reaction 1.1

𝑂 +𝜃 → 2𝑂. 𝜃 Reaction 1.2

𝐶𝑂. 𝜃 + 𝑂. 𝜃 → 𝐶𝑂 + 2𝜃 Reaction 1.3

CO is assumed to be at adsorption equilibrium. Dissociative adsorption of O2 is


also assumed. Finally, the rate limiting step of is also assumed to be the surface reaction.
These leads to the following global rate law:

𝐶 𝐶
𝑟 , =
1+𝐾 𝐶 +𝐾 𝐶

Equation 2.3.1
22

2. Eley-Rideal
This mechanism proposed that adsorbed oxygen atoms reacts with an incoming gas
phase CO.
𝐶𝑂 + 𝜃 → 𝐶𝑂. 𝜃 Reaction 1.1

𝑂 +𝜃 → 2𝑂. 𝜃 Reaction 1.2

𝐶𝑂 + 𝑂. 𝜃 → 𝐶𝑂 + 𝜃 Reaction 1.4

A multitude of experimental evidence suggest that ER mechanism is physically


unreasonable.(Engel et al., 1979; Campbell et al., 1980). First of all, a molecular beam
experiment by Engel and Ertl showed that rate of reaction is not maximum at maximum
coverage of O, shown in Figure 2.3.1. If ER mechanism is to be followed, we would expect it
to be at maximum at maximum O coverage.

Figure 2.3.1: 𝒓𝒄𝒐𝟐 , and Surface Coverage of O and CO for a CO beam Impinging on an
Oxygen Saturated Surface on Pd (111) at T=374K.(Engel and Ertl, 1978)

Moreover, calculations of by Campbell et al. (1980) showed that the ER mechanism is


impossible.

The limitations of this simplified global model is due to two of its assumptions. First,
it assumed CO adsorption to be at equilibrium. Second, the surface reaction is the rate limiting
23

step. The first assumption is wrong when we are dealing dynamic hysteresis modelling, where
there is no equilibrium state when a system is transient. Also, there is no experimental evidence
which supports the assumption of equilibrium CO adsorption. The second assumption is also
erroneous as it is proven that surface reaction between the adsorbed species is not the rate
limiting step.(Herz and Marin, 1980).

However, it is still prevalent throughout recent literatures.(Raj, Harold and Balakotaiah,


2015; Dadi, Luss and Balakotaiah, 2016; Daneshvar et al., 2017) Most researchers used the
model postulated by Voltz et al, which takes into account the CO and hydrocarbon inhibition
effects by introducing an inhibition term.

2.3.2 Mechanistic Kinetics

Mechanistic kinetics, however, takes into account every elementary steps and hence the surface
concentration of species adsorbed. Also, assumptions made when formulating such kinetics are
kept as minimal as possible. The model by Herz and Marin (1980) is one of the earliest works
which uses this approach. In their paper, they argued that the LHWM model is not a correct
representation of this catalytic reaction. The reasons are covered in the previous section. Their
model produced a much better fit than the LH rate law. (Herz and Marin, 1980)

The biggest downfall of this method is that we need to obtain a lot of kinetic parameters
and the calculation could be resource intensive if not very tedious to formulate and key into the
program. Mechanistic calculations are often simplified using appropriate assumptions.
Salomons et. al’s simulation is one of the many simulations done in this way and often provide
satisfactory agreement with the experimental results.(Salomons et al., 2007)

With the introduction of better software capable in handling he tedious and complex
mathematics such as FOTRAN and DETCHEM. Researchers such as Chatterjee et al (2001)
are able to perform more rigorous methods to model CO oxidation. They included the rate laws
of all possible steps when simulating the reaction. A total of 61 pathways are included with
every parameter for the reaction obtained from numerous literatures.

They are able to obtain very good agreements with the experimental data at lean,
stoichiometric and rich conditions but propane oxidation at rich conditions does not fit very
well. They attributed that to the existence of greater variety of species that reside on the surface
24

of the catalyst which can reduce the oxygen coverage. This can be improved by taking into
account the products of partial oxidation of propane. (Chatterjee, Deutschmann and Warnatz,
2001)

Another new approach is to use the Monte Carlo simulation method to replace the
adsorption-desorption model, which takes into account of the average concentration of the
adsorbed species, with a two dimensional surface which is able to take into account of non-
uniform nature of the adsorbed species on the surface. (Ziff, Gulari and Barshad, 1986)
However, this method requires even more computing power when compared with the usual
method. So, we will not attempt such simulation in this project.

Salomons et al (2007) discussed on the importance of using the mechanistic approach


to model the transient light off and extinction curves. They mentioned that the commonly used
global models (mostly modified LH by Voltz et al) are not capable of modelling a converter
performance. Success is only attained in light-off curves over a narrow range of operating
conditions.

The exact mechanistic mechanism and rate law of CO oxidation is highly debated over the past
few decades. The basic elementary steps are:

𝐶𝑂 + 𝜃 → 𝐶𝑂. 𝜃 Reaction
12.1
𝑂 + 2𝜃 ⎯ 𝑂 . 2𝜃 Reaction 2.2

𝑂 +𝜃 ⎯ → 2𝑂. 𝜃 Reaction 2.3

𝐶𝑂. 𝜃 + 𝑂. 𝜃 → 𝐶𝑂 + 2𝜃 Reaction 2.4

Langmuir’s rate law assumed that CO is in adsorption equilibrium (rate of adsorption


equals to the rate of desorption). According to Herz and Marin (1980) this is not supported by
surface science. They attempted to formulate a new model by keeping the assumptions minimal
and reasonable.

In their classical paper, Herz and Marin introduced a series of assumptions which would
later be known as the Oxygen Exclusion Model throughout numerous literatures. (Salomons et
al., 2007)
25

(1) An adsorbed CO excludes other adsorbed CO from an area on the surface equivalent in size
to the area occupied by one surface Pt atom.
(2) An adsorbed O excludes other adsorbed CO from an area on the surface equivalent in size
to the area occupied by two surface Pt atom.
(3) An adsorbed CO excludes other adsorbed O from an area on the surface equivalent in size
to the area occupied by one surface Pt atom.
(4) An adsorbed O excludes other adsorbed O from an area on the surface equivalent in size to
the area occupied by one surface Pt atom.

The first two assumptions are made based on the limiting values of surface concentrations
of CO and O being 1.0 and 0.5 respectively. The latter assumptions agree qualitatively with the
absence of inhibition by an O adlayer on CO adsorption. As a result, the amount of vacant sites
available for CO adsorption will be:

𝜃 , =1−𝜃 −𝜃

Equation 2.3.2

For O2 adsorption it will be:

𝜃 , = 1 − 𝑓𝜃 − 2𝜃

Equation 2.3.3

1 − 2𝜃
𝑓=
1−𝜃

Equation 2.3.4

Most literatures before this one derived rate laws for adsorption based on the surface
concentration of either CO or O but not both. This could be the reason why they find it difficult
to fit the LH model with the experimental data.

Herz and Marin then started their postulation with the basic elementary steps that is presented
earlier:

𝐶𝑂 + 𝜃 → 𝐶𝑂. 𝜃 Reaction
12.1
26

𝑂 + 2𝜃 ⎯ 𝑂 . 2𝜃 Reaction 2.5

𝑂 +𝜃 ⎯ → 2𝑂. 𝜃 Reaction 2.6

𝐶𝑂. 𝜃 + 𝑂. 𝜃 → 𝐶𝑂 + 2𝜃 Reaction 2.7

They are interested in reactions that occur under 573K. At this temperature, oxidation
and reduction of metal oxides by oxygen can be excluded for platinum and palladium catalyst.
Moreover, association of adsorbed oxygen is very slow, which results:

𝑂 . 2𝜃 ⎯ 2𝑂 . 𝜃 → 𝑂 . 2𝜃 2𝑂 . 𝜃

They also assumed that oxygen is in adsorption equilibrium since sticking coefficient is low
on platinum. This also results in another assumption that the rate of dissociation of O2 is very
slow when compared with the rate of desorption. Moreover, surface concentration of O2* is
assumed to be negligible since it is weakly adsorbed. As a result, oxygen adsorption and
dissociation can be assumed as 1 elementary step as O* does not need to be determined
explicitly. Then in the end,

𝐶𝑂 + 𝜃 → 𝐶𝑂. 𝜃 Reaction
12.1
𝑂 + 2𝜃 → 2𝑂. 𝜃 Reaction 2.5

𝐶𝑂. 𝜃 + 𝑂. 𝜃 → 𝐶𝑂 + 2𝜃 Reaction 2.8

The corresponding rate law is as below:

𝑟 , =𝑘 𝜃 𝜃

Equation 2.3.5

The surface coverage of CO and O needed to be calculated from the following system of
ordinary differential equations:

𝜃̇ =𝑘 𝐶 𝜃 , −𝑘 𝜃 −𝑘 𝜃 𝜃

Equation 2.3.6
27

𝜃̇ = 𝑘 𝐶 𝜃 , −𝑘 𝜃 𝜃

Equation 2.3.7

𝜃 , =1−𝜃 −𝜃

Equation 2.3.8

𝜃 , = 1 − 𝑓𝜃 − 2𝜃

Equation 2.3.9

1=𝜃 , +𝜃 , +𝜃 +𝜃

Equation 2.3.10

This set of reactions and rate laws will then provide the basis for researchers later on who try
to explain on the other two phenomena which are known ever since CO oxidation experiments
are conducted: state multiplicity/hysteresis and periodic oscillations. Razon and Schmitz (1986)
described the model by Herz and Marin as the “basic model” which would then be tinkered
around to explain the aforementioned phenomena.

2.4 Modifications to the Basic Model

2.4.1 Island Model

Herz and Marin (1980) dismissed the possibility of island formation due to the small size of
their Pt crystals (in the magnitude of nanometers). They do consider the potential effects of
boundaries of islands where islands of CO and O2 is present and reaction only occurs on the
boundaries. However it is not included since in their experiment the analyzed area is small
enough so that the effect is diminished.

Mukesh et al (1984) formulated a set of mathematical models which takes into account of the
effect of CO islands. They pointed out that the presence of CO islands (and potentially O
islands) will limit the rate of reaction. It is believed that reactions only occur on the island
boundaries. To take into CO island boundaries, they used the following rate law:

𝑟 , ∝𝜃 𝜃

Equation 2.4.1
28

The square root term of 𝜃 is a mathematical treatment to the island boundaries. They also
assumed that there will be no O within the CO islands and vice versa. Already we can see that
this assumption is flawed as there is evidence that CO are able to adsorb into the O adlayer or
in this case O island. (Engel et al., 1979) In the end, they concluded that CO islands affects the
rate more significantly.(Mukesh et al., 1984)

However, the presence of CO islands will affect the rate of adsorption and desorption as
well. Graham and Lynch (1987) provided new rate equations that takes into account of such
effects. Then, they came up with a two models. The first model is a simple island model while
the second model takes into account the effects of increased local concentration due to island
formation. Then, it is concluded that the second model is the better because it fits the
experimental data better when compared with the other two models.

Graham and Lynch (1987) modified the rate law for adsorption and desorption of CO as:

𝑟 , ∝ 𝑎 [𝐶𝑂](1 − 𝜃 −𝜃 )

Equation 2.4.2

𝑟 , ∝𝑎 𝜃

Equation 2.4.3

Then, the net rate of CO adsorption will be:

𝑚𝜃 𝑚𝜃
𝑟 , =𝑘 , 𝐶 (1 − 𝜃 −𝜃 )−𝑘 , 𝜃
𝑉𝑌 𝑉𝑌

Equation 2.4.4

The modified rate law for oxygen adsorption and desorption would be:

𝑎 [𝑂 ](1 − 𝜃 − 𝜃 )
𝑟 , ∝
(1 − 𝜃 )

Equation 2.4.5

𝑎 𝜃
𝑟 , ∝
(1 − 𝜃 )

Equation 2.4.6
29

Then, the net rate of O2 adsorption will be:

𝑚𝜃 (1 − 𝜃 − 𝜃 ) 𝑚𝜃 𝜃
𝑟 , =𝑘 , 𝐶 −𝑘 ,
𝑉𝑌 (1 − 𝜃 ) 𝑉𝑌 (1 − 𝜃 )

Equation 2.4.7

Finally, the surface reaction should be modified as:

𝑚𝜃 𝜃
𝑟 , , =𝑘 , 𝜃
𝑉𝑌 (1 − 𝜃 )

Equation 2.4.8

The above surface reaction assumes only CO islands are formed.

If only O islands are present:

𝑚𝜃 (1 − 𝜃 − 𝜃 ) 𝑚𝜃
𝑟 , =𝑘 , 𝐶 −𝑘 , 𝜃
𝑉𝑌 (1 − 𝜃 ) 𝑉𝑌

Equation 2.4.9

𝑚𝜃 𝜃
𝑟 , , =𝑘 , 𝜃
𝑉𝑌 (1 − 𝜃 )

Equation 2.4.10

If both CO and O form surface islands:

𝑚𝜃 𝑚𝜃
𝑟 , =𝑘 , 𝐶 (1 − 𝜃 −𝜃 ) −𝑘 , 𝜃
𝑉𝑌 𝑉𝑌

Equation 2.4.11

𝑚𝜃
𝑟 , , =𝑘 , 𝜃 𝜃
𝑉𝑌

Equation 2.4.12
30

The above equation is investigated and refined by Dumont et. al. and the following is
formulated as the best fit equation to the experiment (Dumont, Poriaux and Dagonnier, 1986):

𝑚𝜃
𝑟 , =𝑘 , 𝐶 (1 − 𝜃 −𝜃 )
𝑉𝑌

Equation 2.4.13

where 𝛾 is a constant in between 1.5 to 2.

𝑚𝜃
𝑟 , , =𝑘 , 𝜃 (1 − 𝜃 )𝜃 (1 − 𝜃 )
𝑉𝑌

Equation 2.4.14
The dimensionless kinetic parameters reported in this paper is temperature independent.
Therefore, it would not be suitable for our simulation since we are dealing with changing inlet
temperature rather than changing the inlet CO concentration as done by the researchers. So, we
need to import kinetic parameters from other models which varies with temperature.

2.4.2 Salomons’ Compressed Oxygen Model


Salomons et al (2007) concluded in their paper that the compressed oxygen model best fits
their experimental data when they are trying to model a hysteresis curve. In this model, the
aforementioned basic model is utilized with the inclusion of a few steps which describes the
compression effect that happens while CO is adsorbed onto an oxygen adlayer. This
compression effect can affect the ignition temperature, usually the effect being lowering the
said property. Subsequent reason of why it is such is covered in section 2.2.3 part b).

𝐶𝑂 + 2𝑂. 𝜃 → 𝐶𝑂. 𝜃 + 𝑂𝑂 . 𝜃 Reaction 2.9


𝐶𝑂 . 𝜃 + 𝑂𝑂. 𝜃 → 𝐶𝑂 + 𝑂 . 𝜃 + 𝜃 Reaction 2.10
𝑂𝑂. 𝜃 → 2𝑂. 𝜃 Reaction 2.11

2.5 Adsorption and Desorption Kinetics

Another key component in simulating an accurate hysteresis phenomena is the adsorption and
desorption kinetics of the model. It is key that we use the most appropriate and accurate kinetics.
31

a) CO Adsorption Kinetics

The rate of adsorption of a species onto a solid surface that is formulated by Langmuir is given
as (Langmuir, 1922):

𝑑𝑛 𝑝
𝑟, = = ∙ 𝑠(𝜃)
𝑑𝑡 2𝜋𝑀 𝑘𝑇

Equation 2.5.1
where 𝑛 is the density of the adsorption species per cm2, 𝑝 is the gas pressure of the species,
𝑚 the absorbate mass, k the Boltzmann constant, T the absolute temperature, and 𝑠(𝜃)the
probability of adsorption or simply the sticking coefficient, which is the function of surface
concentration.

The first term represents the gas flux on the surface of the solid while the second term
describes the kinetics of the adsorption. The above rate equation is based on the assumption
that desorption rate is negligible which is valid if only the surface temperature is low enough.

The initial sticking coefficient, s0, can be determined experimentally using molecular beam
experiments which can yield accurate results if s0 is close to 1, which is true for Pt surfaces.
Sticking coefficient is independent on surface temperature, between 300K and 650K, and also
the angle of incidence of the gas flux (Engel et al., 1979). Also, it is shown experimentally that
s0 is also independent on the atomic oxygen surface coverage (Gland et al., 1983) . However,
McCabe and Schmidt (1977) showed experimentally that s0 would vary between 0.24 and 0.95
when they varied the angle of incidence. Therefore, generally:

𝑠 = 𝑓(𝜃 )

The mechanism of the adsorption dynamics according to Langmuir is:

𝑠(𝜃) = 𝑠 (1 − 𝜃 )

Equation 2.5.2
𝑠(𝜃) = 𝑠 (1 − 𝜃 )

Equation 2.5.3
32

where Equation 2.5.2 is used if only one free site is needed and Equation 2.5.3 is used if two
adjacent free sites are required.

However, sticking coefficient of CO are constant and decreases at a lesser rate at lower
coverage, evident from Figure 2.5.1.

Figure 2.5.1: Relative Sticking Coefficient as a Function of CO Coverage. (Ertl,


Neumann and Streit, 1977)

The precursor state model is used to explain this behavior. This model tells us that a
molecule may diffuse into a layer below the chemisorbed species and it will stay there until a
vacant adjacent site is present so that it can chemisorb. The following is the sticking coefficient
according to the precursor state model (Kisliuk, 1957):

(1 − 𝜃)
𝑠(𝜃) = 𝑠
(1 − 𝑐𝜃)

Equation 2.5.4
,
𝑓 +𝑓 −𝑓
𝑐=
𝑓 +𝑓

Equation 2.5.5
33

where,

𝑓 is the probability for adsorption of the precursor molecule on an empty site

𝑓 is the probability for desorption of the precursor from a bare site

𝑓 , is the probability of desorption from an occupied site.

For c = 0, the Langmuir model is obtained. Experiments were carried out to obtain the
parameter, c, and satisfactory results were obtained where c = 0.85 for Pt (111). Yet, the model
does not agree quantitatively well with the experiments as mentioned by Engel et al (1979)

b) CO Desorption Kinetics

According to Langmuir (1922), the rate of desorption is as below:

𝑑𝑛 −𝐸 (𝑛 )
𝑟, =− = 𝑣 (𝑛 ) exp ∙𝑛
𝑑𝑡 𝑘𝑇

Equation 2.5.6

There is nothing extraordinary reported by researchers regarding CO desorption. The above


equation should suffice in modelling rate of desorption as it has both temperature and surface
coverage dependence terms. As temperature of the surface increases, the rate of desorption will
increase too. Also, due to the repulsive effects of neighboring adsorbed species, 𝐸 will
decrease with increasing surface coverage. It is usually expressed as:

𝐸 (𝑛 ) = 𝐸 − 𝛽𝜃

Equation 2.5.7

c) O2 Adsorption Kinetics

Rate of adsorption of O2 on Pt (111) is similar as the one for CO:

𝑑𝑛 𝑝
𝑟 = = ∙ 𝑠(𝜃)
𝑑𝑡 √2𝜋𝑚𝑘𝑇
Equation 2.5.1
34

However, the sticking coefficient does not follow the precursor state model. It decreases
exponentially with O2 coverage. Moreover, various experiments suggested that O2 adsorption
increases with temperature unlike the adsorption of CO. Also from Figure 2.5.2, initial sticking
coefficient decreases linearly with temperature.(Engel, 1978)

Figure 2.5.2: Initial Sticking Coefficient as a Function to the Substrate Temperature.


(Engel, 1978)

Bonzel and Ku (1973) developed the following model for the sticking coefficient:

−(𝐸 + 𝛽𝜃)
𝑠 = 𝛿𝑓(𝜃)𝑒𝑥𝑝
𝑅𝑇

Equation 2.5.8
where,

𝛿 is the accommodation coefficient,

𝑓(𝜃) is the aforementioned Equation 2.5.2 or Equation 2.5.3 ,

𝐸 is the initial adsorption energy which is 60kcal/mol


35

𝛽 is a constant which is determined to be 23 kcal/mol.(Bonzel and Ku, 1973)

Some years later, Hopster, Ibach and Comsa (1977) developed the following model:

. )
−(3.2𝑒𝑉𝜃
𝑠 = 0.018𝑒𝑥𝑝
𝑘𝑇

Equation 2.5.9
where k is the Boltzmann constant.

This model is developed by taking into account the effects of electrostatic potential on
activation energy which then predicts the sign of effect, the exponential dependence, and
approximately, the exponent.(Hopster, Ibach and Comsa, 1977).

d) O2 desorption

O2 desorption is generally not observed at temperatures that a DOC will experience. (Herz
and Marin, 1980) Therefore, it will not be reviewed in this project.

2.6 Hydrocarbon Oxidation


Apart from CO oxidation, unburnt hydrocarbons are also catalytically combusted in the DOC.
There is a need to include the effects of such reactions in the simulation to provide a better
approximation. Propylene is a good molecule to use as a representative for all the unburnt
hydrocarbon in the DOC. (Raj, Harold and Balakotaiah, 2015; Dadi, Luss and Balakotaiah,
2016)

𝑪𝟑 𝑯𝟔 + 𝟒𝑶𝟐 → 𝟑𝑪𝑶𝟐 + 𝟐𝑯𝟐 𝑶 Reaction 2.12

2.7 Fine Tuning Strategies


A common practice to fit the calculations from a model with real experimental data is to fine-
tune certain parameters. The usual strategy adopted is to fine tune the kinetic parameters of the
elementary reactions such as pre-exponential factors and activation energies. However, this
strategy often leads to the parameters being specific to the catalyst that is used. Sjoblom, Ab
and Creaser (2014) suggested to include parameters that influence the mass transport properties
of the washcoat such as effective diffusivity to avoid such problem. Indeed with this new
strategy, researchers such as Lundberg et. al. (2015) are able to obtain a better fit when both
effective diffusivity and kinetic parameters are tuned simultaneously.
36

However, simultaneous tuning of parameters to obtain the optimal fit is beyond the
scope of this project. Therefore, tuning the effective diffusivity through diffusivity constants
would be the strategy adopted.

2.8 Programming Language Utilized

The Python Programming Language is chosen for this task. The following are the advantages
of using Python:

1) Contains existing common mathematical methods such as plotting a graph or solving


an ODE. We do not need to translate the method into codes again as Python already has
it prebuilt inside the language.
2) It is easy to learn and master.
3) Does not contain complicated symbols or syntax so that anyone reading the code would
understand what the coder is trying to do.
4) It is developed with efficiency as its first priority. We are able to quickly develop fast
codes with Python.
5) It can be applied to and solve many problems in different areas

For this project, we will use the Scientific Python Development Environment to compile
our codes. It is an integrated development environment where we can compile, debug and run
our program in it. Moreover, it has the necessary libraries such as NumPy and MatPlotLib that
we can use to perform our simulation.

2.9 Numerical methods

The solutions to the partial differential equations of heat and mass transfer cannot be obtained
analytically. Therefore, we need to employ numerical methods to solve for our PDEs and ODEs.
The method we will be using is the method of lines. Method of lines generally refers to the re-
expression of PDEs as a system of ODEs. The spatial derivative of the PDE can be replaced
with algebraic expressions. As a result, we would obtain a system of ODEs with time as its
only derivative. (Hamdi et al, 2007) Derivatives can be approximated into algebraic
expressions using the finite difference method. This method can be categorized into three
methods: explicit, implicit, and Crank-Nicholson method. If the boundary conditions are
37

specified the solution of every recurrence equation can be obtained and through numerical
iterations we can obtain the whole complete solution albeit an approximated one.

2.9.1 Explicit method

Figure 2.9.1: Stencil for the Explicit Method (adapted from Wikipedia.com).

This method uses the forward difference for the time derivative at time tn and a central
difference for the space derivative at xi (forward-time-central space scheme or just simply
FTCS). A recurrence equation can be derived for a 1D heat equation

𝑈 = 𝛼𝑈

Equation 2.9.1

𝑢 −𝑢
𝑈 =
𝑘

Equation 2.9.2

𝑈 =

Equation 2.9.3

𝑢 −𝑢 𝑢 − 2𝑢 + 𝑢
=𝛼
𝑘 ℎ

Equation 2.9.4

There exist a stability criteria for explicit method to work.(Chapra and Canale, 2010) The
criteria is given as:

𝛼𝑘 1

ℎ 2

Equation 2.9.5
38

2.9.2 Implicit method

Figure 2.9.2: Stencil for the Implicit Scheme (adapted from Wikipedia.com).

This method utilizes the BTCS scheme where the solution to the point j,n+1 is obtained
implicitly with the following recurrence equation:

𝑢 −𝑢 𝑢 − 2𝑢 +𝑢
=
𝑘 ℎ

Equation 2.9.6

There will be no stability condition for this method unlike the explicit method. However,
accuracy is sometimes greatly sacrificed.(Chapra and Canale, 2010)

2.9.3 Crank-Nicholson method

Figure 2.9.3: Stencil for the Crank-Nicholson Method (adapted from Wikipedia.com).

This method employs a central difference for the time derivative therefore, a CTCS scheme.
The recurrence equation is:

𝑢 −𝑢 1 𝑢 − 2𝑢 +𝑢 𝑢 − 2𝑢 + 𝑢
= +
𝑘 2 ℎ ℎ

Equation 2.9.7
39

This method is unconditionally stable with an even higher accuracy when compared with the
implicit method.(Chapra and Canale, 2010)

2.9.4 Selecting the appropriate method

Explicit method is chosen due to its ease of translation into Python coding. The ODEINT
function uses an explicit method by default. The stability criteria will be addressed and its
details will be covered in the methodology section.

2.10 Experimental data available to perform the simulation

The data from Ye et. al.’s experiment on thin slice DOCs with an inlet CO concentration of
3000ppm is chosen. One of the reason being that it requires a lot less computing power and
time to perform numerical integration of the transient partial differential equation. Thin slice
DOCs do exhibit hysteresis behavior which is the key of this project. (Ye et al., 2012) Only
one data is chosen since up this time of report writing, no other researchers have performed
experiments on thin slice DOCs.
40

CHAPTER 3

METHODOLOGY

3.1 Single Monolith Channel

A single channel can represent the whole channel by assuming uniform inlet gas velocity and
catalyst distribution, and no heat loss along the radius of the channel. The boundaries of the
simulation would be from the centerline of the channel up to half of the substrate thickness.

The use of the 1D-2D axi-symmetric model is seen for most literatures, with gas phase
heat and mass transfer being 1 dimensional and heat and mass transfer through washcoat 2
dimensional. Experimental evidence, which justifies the use of 1D modelling for gas phase
diffusion, shows that radial gradient at gas phase is non-existent. On the other hand, most would
agree on the significance of concentration gradient that exist on the washcoat layers so that a
2D model is required. (Mukadi and Hayes, 2002) However, due to the difficulty in translating
the 2D model into a working code, simplifications are made and a 1D model is used for a
washcoat instead.

3.2 Heat and Mass Transfer Equations

In the fluid phase, the one dimensional mole balance is:

𝝏𝑪𝒊,𝒇 𝝏 𝝏𝑪𝒊,𝒇 𝝏 𝒗𝒛 𝑪𝒊,𝒇 𝟒


= 𝑫𝒆𝒇𝒇,𝒊,𝒇 − − 𝒌 𝑪 − 𝑪𝒊,𝒔
𝝏𝒕 𝝏𝒛 𝝏𝒛 𝝏𝒛 𝑫𝑯 𝒎,𝒊 𝒊,𝒇

Figure 3.2.1
41

The effective axial dispersion coefficient, 𝐷 , , is:

𝑣 𝑅
𝐷 , =𝐷 +
48𝐷

Equation 3.2.1

The boundary conditions are:

𝜕 𝑣 𝐶, 𝜕 𝑣 𝐶,
= 𝑣 𝐶, 𝑎𝑛𝑑 =0
𝜕𝑧 𝜕𝑧

The one dimensional energy balance is:

𝜕𝑇 𝜕 𝜕𝑇 𝜕𝑇 4
𝜌 𝐶 = 𝐾 , −𝜌 𝐶 𝑣 − 𝑘 𝑇, − 𝑇
𝜕𝑡 𝜕𝑧 𝜕𝑧 𝜕𝑧 𝐷

Equation 3.2.2

The effective axial thermal diffusivity,𝐾 , , is:

𝑣 𝑅
𝐾 , =𝛼 +
48𝛼

Equation 3.2.3
The boundary conditions are:

𝜕𝑇 𝜕𝑇
= 0 𝑎𝑛𝑑 =0
𝜕𝑧 𝜕𝑧

In the washcoat, the one dimensional mole balance is:

𝜕𝐶 , 𝑊 𝜕 𝜕𝑠
𝜀 = 𝑘 , 𝐶, −𝐶, + 𝐷 ,, + 𝜉𝑅
𝜕𝑡 𝐴 𝜕𝑧 𝜕𝑧

Equation 3.2.4
42

The boundary conditions are

𝜕 𝐶, 𝜕 𝐶,
= 0 𝑎𝑛𝑑 =0
𝜕𝑧 𝜕𝑧

The one dimensional energy balance within the washcoat is:

𝜕𝑇 𝑊 𝜕 𝜕𝑇
𝜌 𝐶 = 𝑘 𝑇, − 𝑇 + 𝐾 , + (∆𝐻𝑅)
𝜕𝑡 𝐴 𝜕𝑧 𝜕𝑧

Equation 3.2.5

The boundary conditions are:

𝜕 𝑇 𝜕(𝑇 )
= 0 𝑎𝑛𝑑 =0
𝜕𝑧 𝜕𝑧

The Nusselt and Sherwood numbers are utilized to obtain the heat and mass transfer
coefficients. Their values are set to equal 4.0 since fully developed flows are assumed with the
conditions of constant wall temperature and wall flux.

𝑘 𝐷
𝑁𝑢 = = 4.0
𝑘
𝑘 𝐷
𝑆ℎ = = 4.0
𝐷

Knudsen diffusion mechanics are the dominant diffusion mechanism in the reactor washcoat.

.
𝜀 𝑇
𝐷 , = 97𝑎
𝜏 𝑀

Equation 3.2.6
43

3.3 Adsorption and Desorption Kinetics


For this project, all the models other than the Island models will adopt the adsorption and
desorption kinetics used by Herz and Marin (1980) in their model.

a) CO Adsorption Kinetics

1000𝑅𝑇 𝑃𝑌 ,
𝑟 , = 𝜎 𝑆 (1 − 𝜃 −𝜃 )
2𝜋𝑀 𝑅𝑇

Equation 3.3.1

where,

𝑅 is the ideal gas constant, J/molK

𝑇 is the temperature of the gas, K

𝑀 is the molar mass of CO, g/mol

𝑃 is the pressure of the gas, Pa

𝑌 , is the mol fraction of CO in the washcoat

𝑆 is the sticking coefficient off CO on the surface

𝜃 is the fractional coverage of CO on the surface

𝜃 is the fractional coverage of O on the surface


44

b) CO Desorption Kinetics

1000(𝐸 − 𝛽𝜃 )
𝑟 , = 𝑘 𝑒𝑥𝑝 − 𝜃
𝑅𝑇

Equation 3.3.2

where,

𝑘 is the pre exponential factor for CO desorption, s-1

𝐸 is the activation energy for CO desorption, kJ/mol

𝛽 is the bond-weakening capacity of CO on the surface, kJ/mol

𝜃 is the fractional coverage of CO on the surface

𝑅 is the ideal gas constant, J/molK

𝑇 is the washcoat temperature, K

c) O2 Adsorption Kinetics

1000𝑅𝑇 𝑃𝑌 , 1 − 2𝜃
𝑟 , =2 𝜎 𝑆 1− 𝜃 − 2𝜃
2𝜋𝑀 𝑅𝑇 1−𝜃

Equation 3.3.3

where,

𝑅 is the ideal gas constant, J/molK

𝑇 is the temperature of the gas, K

𝑀 is the molar mass of O, g/mol

𝜎 is the surface are covered by 1 mol of surface metal atoms, m2/mol


45

𝑃 is the pressure of the gas, Pa

𝑌 , is the mol fraction of CO in the washcoat

𝑆 is the sticking coefficient of oxygen on the surface

𝜃 is the fractional coverage of CO on the surface

𝜃 is the fractional coverage of O on the surface

3.4 Reaction Models Investigated

In this project, a number of global and mechanistic model were simulated and investigated.

3.4.1 Voltz Global Kinetic Model

Voltz et al (1973) developed the global kinetic rate equation for catalytic CO oxidation.

𝑌 𝑌
𝑟 , =𝑘
𝑅(𝜃)

Equation 3.4.1

𝑅(𝜃) is a term which takes into account of the inhibition effects of CO, C3H6, and NO on the
reaction rates.

.
𝑅(𝜃) = 1 + 𝑘 𝑌 +𝑘 𝑌 × 1+𝑘 𝑌 𝑌 × [1 + 𝑘 𝑌 ]

Equation 3.4.2

The first term takes into account of the inhibition effect of CO and C 3H6 while the second term
is needed to fit the experimental data at higher concentrations of CO and C3H6.(Voltz et al.,
1973) The last term can be ignored since it takes into account of the effect of NO which is not
considered in this project.

3.4.2 Herz and Marin Basic Model

The elementary steps involved are:


46

𝐶𝑂 + 𝜃 → 𝐶𝑂. 𝜃 Reaction
12.1
𝑂 + 2𝜃 → 2𝑂. 𝜃 Reaction 2.5

𝐶𝑂. 𝜃 + 𝑂. 𝜃 → 𝐶𝑂 + 2𝜃 Reaction
2.13

The adsorption and desorption kinetics are mentioned in section 3.3. The rate of surface
reaction is given by:

1000(𝐸 − 𝛽𝜃 )
𝑟 , , = 𝑘 exp − 𝜃 𝜃
𝑅𝑇

Equation 3.4.3

A small alteration to the activation energy is made such that it will decrease with increasing
surface coverage of CO. The involved parameters, 𝛽, will be taken from Salomons’ work.

3.4.3 Island Model

In this project, a modification of the Islands models was made. The rate constants, k, capacity
factor, Z, and gas phase concentration will be replaced with the rate constants, collision
frequency and sticking coefficient from Herz and Marin (1980). Also, oxygen desorption is
assumed to be negligible. The following equations are proposed:

CO islands

1000𝑅𝑇 𝑃𝑌 ,
𝑟 , = 𝜎 𝑆 (1 − 𝜃 −𝜃 )
2𝜋𝑀 𝑅𝑇

Equation 3.4.4

1000(𝐸 − 𝛽𝜃 )
𝑟 , = 𝑘 𝑒𝑥𝑝 − 𝜃
𝑅𝑇

Equation 3.4.5

1000𝑅𝑇 𝑃𝑌 , (1 − 𝜃 − 𝜃 )
𝑟 , , = 𝜎 𝑆
2𝜋𝑀 𝑅𝑇 (1 − 𝜃 )

Equation 3.4.6
47

1000(𝐸 − 𝛽𝜃 ) 𝜃
𝑟 , = 𝑘 exp − 𝜃
𝑅𝑇 (1 − 𝜃 )

Equation 3.4.7

O islands

,
𝑟 , = 𝜎 𝑆 (1 − 𝜃 −𝜃 )

Equation 3.4.4

1000(𝐸 − 𝛽𝜃 )
𝑟 , = 𝑘 𝑒𝑥𝑝 − 𝜃
𝑅𝑇

Equation 3.4.8

1000𝑅𝑇 𝑃𝑌 , (1 − 𝜃 − 𝜃 )
𝑟 , , = 𝜎 𝑆
2𝜋𝑀 𝑅𝑇 (1 − 𝜃 )

Equation 3.4.9

1000(𝐸 − 𝛽𝜃 ) 𝜃
𝑟 , = 𝑘 exp − 𝜃
𝑅𝑇 (1 − 𝜃 )

Equation 3.4.10

CO and O Islands

1000𝑅𝑇 𝑃𝑌 ,
𝑟 , = 𝜎 𝑆 (1 − 𝜃 −𝜃 )
2𝜋𝑀 𝑅𝑇

Equation 3.4.4
48

1000(𝐸 − 𝛽𝜃 )
𝑟 , = 𝑘 𝑒𝑥𝑝 − 𝜃
𝑅𝑇

Equation 3.4.11

1000𝑅𝑇 𝑃𝑌 ,
𝑟 , , = 𝜎 𝑆 (1 − 𝜃 −𝜃 )
2𝜋𝑀 𝑅𝑇

Equation 3.4.12

1000(𝐸 − 𝛽𝜃 )
𝑟 , = 𝑘 exp − 𝜃 𝜃
𝑅𝑇

Equation 3.4.13

Dumont Postulation

1000𝑅𝑇 𝑃𝑌 ,
𝑟 , = 𝜎 𝑆 (1 − 𝜃 −𝜃 )
2𝜋𝑀 𝑅𝑇

Equation 3.4.4

1000(𝐸 − 𝛽𝜃 )
𝑟 , = 𝑘 𝑒𝑥𝑝 − 𝜃
𝑅𝑇

Equation 3.4.14

1000𝑅𝑇 𝑃𝑌 ,
𝑟 , , = 𝜎 𝑆 (1 − 𝜃 −𝜃 )
2𝜋𝑀 𝑅𝑇

Equation 3.4.15

1000(𝐸 − 𝛽𝜃 )
𝑟 , = 𝑘 exp − 𝜃 (1 − 𝜃 )𝜃 (1 − 𝜃 )
𝑅𝑇

Equation 3.4.16

𝛾 is chosen to be 2 for this project.


49

3.4.4 Salomon’s Compressed Oxygen Model

The same elementary steps are used from the Herz and Marin basic model with an addition of
elementary steps accounting for the compressed oxygen effect:

𝐶𝑂 + 2𝑂. 𝜃 → 𝐶𝑂. 𝜃 + 𝑂𝑂 . 𝜃 Reaction 2.14


𝐶𝑂 . 𝜃 + 𝑂𝑂. 𝜃 → 𝐶𝑂 + 𝑂 . 𝜃 + 𝜃 Reaction 2.15
𝑂𝑂. 𝜃 → 2𝑂. 𝜃 Reaction 2.16
The respective rate law of the above reactions are:
𝑟 =𝑘 𝐶 𝜃
Equation 3.4.17

𝑟 =𝑘 𝜃 𝜃
Equation 3.4.18
𝑟 =𝑘 𝜃 𝜃 ,

Equation 3.4.19

The rate constants for Equation 3.4.17, Equation 3.4.18, and Equation 3.4.19 follows the
Arrhenius Equation(Salomons, 2008):

𝐸
𝑘 = 𝐴 exp
𝑅 𝑇

Equation 3.4.20

3.5 Hydrocarbon Oxidation


To simplify the model, a global kinetic approach is taken when dealing with hydrocarbon
oxidation.

𝑌 𝑌
𝑟 =𝑘
𝑅(𝜃)

Equation 3.5.1

𝑅(𝜃) is a term which takes into account of the inhibition effects of CO, C3H6, and NO on the
reaction rates.
50

.
𝑅(𝜃) = 1 + 𝑘 𝑌 +𝑘 𝑌 × 1+𝑘 𝑌 𝑌 × [1 + 𝑘 𝑌 ]

Equation 3.4.2

3.6 Coding
Step 1: Discretisation of Monolith Channel

The monolith channel is divided sections of 1mm each. Therefore, a 5mm slice would be
divided into 6 sections: 0mm, 1mm, 2mm, 3mm, 4mm, and 5mm.

Figure 3.6.1: Coding for Grid Discretisation

Step 2: Defining Monolith Properties and Characteristics and also other Constants

The monolith properties are defined from the data provided by Ye et. al (2012).

Figure 3.6.2: Coding for Monolith Data


51

Constants such as ideal gas constant and molecular weight are also defined at this stage.

Figure 3.6.3: Coding for Molecular Properties and other constants

Step 3: Initialisation of the Variables

Variables are stated. 4 types of variables are needed: gas phase concentration denoted by Y,
gas phase concentration within the solid washcoat denoted by Ys, fractional coverage denoted
by fc and temperature denoted by T. Also, a letter is used as the name of each species to simplify
the code.

Table 3.6.1: Letter Designation for each Species

Letter Species

A Carbon Monoxide

B Carbon Dioxide

C Propane

The initial values of the variables are also declared.


52

Figure 3.6.4: Coding for Variable Initialisation

Step 4: Setting up the dynamic parameters

Kinetic parameters that are dependent on physical properties of the system are defined.

Figure 3.6.5: Coding for Dynamic Parameters (Gas Phase Properties)


53

Figure 3.6.6: Coding for Dynamic Constants (Solid Phase Propeties)

Figure 3.6.7: Coding for Dynamic Constants (Heat and Mass Transfer Coefficient)
54

Figure 3.6.8: Coding for Dynamic Constants (Diffusivity Parameters)


55

Step 5: Setting up the inlet gas temperature profile

The inlet gas temperature is obtained by examining the graph below:

Figure 3.6.9: Temperature against time plot from Ye et . al.(2012)

Python is able to plot the temperature profile by just identifying the end points of each stage.
There will be six different ramp up and ramp down rates in this experiment. Six end points of
the stages are identified: 417K, 505K, 545K, 545K, 485K, 445K, and 417K.

Figure 3.6.10: Coding for Setting up the Inlet Temperature Profile


56

Step 6: Setting up the models

A model is created which returns the value of the differentials. The details on the mathematics
behind it is in appendix.

Variables from the input, y0, is ‘dissected’ into its constituents

Figure 3.6.11: Coding to start the model function

The differential variable is declared

Figure 3.6.12: Coding for the declaration the differential variables

The differential variable for the fractional coverage of each species will be declared later on.
57

Then, the parameters of the reaction model is declared

Figure 3.6.13: Coding for declaration of parameters of the reaction model.

The elementary equations:

Figure 3.6.14: Coding for the setting up of rate equations


58

Then, the heat and mass balances:

Figure 3.6.15: Coding for Setting up the Heat and Mass Balances
59

Finally, the differential value calculated will be returned as a single array, D. This is when the
boundary condition is defined.

Figure 3.6.16: Coding to close the model function and return the calculated differential
values

Step 7: Setting up the ODE solver and storing the calculated values

An ODE solver is set up by using the defined models and initial values of the variables are
stated. The heat and mass balance equations are solved in this step.

Figure 3.6.17: Coding to setting up the storage array

A two dimensional array is created, first axis having the size of the maximum length of
the stage and the second axis having the size of the returned array from the odeint function. Its
function is to store all the computed values for a stage so that it can be called back for graph
plotting and to be exported to a spreadsheet. In this case, it will be sol_grid.

Another array will be created for the same purpose but this array will store all the
computed values for all the six stages. It will have another dimension with the size as the
number of stages. In this case, it will be sol_grid_2.
60

A loop is then created so that each iteration will solve a particular stage

Figure 3.6.18: Coding for the stage loop

The loop starts off by calculating the temperature slope of the stage. This slope will be
named as mp. The purpose of this slope is to be the boundary condition of our inlet gas and
solid temperature as indicated at the previous step.

Next, an array called ‘time’ is created based on the time length of each stage and each
with the number of intervals calculated as:

𝑠𝑡𝑎𝑔𝑒 𝑙𝑒𝑛𝑔𝑡ℎ
𝑑𝑒𝑠𝑖𝑔𝑛𝑎𝑡𝑒𝑑 𝑡𝑖𝑚𝑒 𝑖𝑛𝑡𝑒𝑟𝑣𝑎𝑙

In this project, the designated time interval would be 1s.

Figure 3.6.19: Coding for the solution loop.

Then, another loop is then created so that the odeint will solve the model from t = 0s to
t = 1s, instead of the conventional method whereby most would solve from t = 0s to t = 200s.
This method will allow us to avoid the stability issues that would be eminent in our case as the
differential in length is small. Having a small differential length means we need to have a
61

smaller differential in time to fulfill the stability criteria. So, instead of solving for a big
timeframe i.e. from 0 to 200s, we need to solve it at a smaller timeframe i.e. from 0 to 1s. The
disadvantage of this method is that we need more time to compute the solution as we need to
solve for 1800 time steps for our case (total time involved is 1800s).

The initial condition will be declared as y0. Then, the solution will be stored in sol_grid
array. The solution will also be stored in the y0 array which will then be used as the initial point
of the next solution. After that, sol_grid will be resized such that the excess parts of the array
will be removed.

Figure 3.6.20: Coding for Data Storage

Finally, with the pd.DataFrame function, we can export the array into a spreadsheet to
save the computed data externally. Also, the sol_grid will be stored into sol_grid_2. Then, the
sol_grid array needs to be reset at the end of this loop.

When the loop has finished every calculation, sol_grid_2 will be converted into a two
dimensional array and exported as a spreadsheet and graph plotting will be initiated.

Figure 3.6.21: Coding for the compilation of data of all stages and its storage.
62

Step 8: Graph plotting

The required graphs are plotted by using the plotting function and are exported as a PNG image
file. An array ‘timespace’ is created to become the x axis of certain plots.

Figure 3.6.22: Coding for Graph Plotting


63

CHAPTER 4

RESULTS AND DISCUSSION

4.1 Hysteresis (with respect to inlet gas temperature)


A conversion against inlet gas temperature is plotted to check how close it can resemble to
the experimental results.
64

Figure 4.1.1: CO Conversion against inlet temperature plot from the experimental
results of Ye et. al.(Ye et al., 2012)

Figure 4.1.2: CO Conversion against inlet temperature plot for Voltz Model
V
65

Figure 4.1.3: CO Conversion against inlet temperature plot for Herz and Marin Basic
Model

Figure 4.1.4: CO Conversion against inlet temperature plot for Salomons' Compressed
Oxygen Model
66

Figure 4.1.5: CO Conversion against inlet temperature plot for CO Island Model

Figure 4.1.6: CO Conversion against inlet temperature plot for O Island Model
67

Figure 4.1.7: CO Conversion against inlet temperature plot for CO-O Island Model

Figure 4.1.8: CO Conversion against inlet temperature plot for Dumont Postulation
68

Generally, for all models other than the Dumont Postulation, agreement with the experimental
results on the conversion of CO is within the range of ±5%. However, only the Herz and Marin
Basic Model, and Salomons’ Compressed Oxygen Model showed good fit to the overall
hysteresis curve. Meanwhile, both O Island Model and CO-O Island Model showed hysteresis
albeit the width of hysteresis is very small (around 10℃ or less). Finally, Voltz Model, CO
Island Model and Dumont Postulation failed to show hysteresis.

Table 4.1.1: Summary of Hysteresis Characteristics from the simulations done (Gas
Inlet Temperature)

Ignition Extinction Width Conversion


Temperature Temperature Of
Hysteresis
℃ ℃ ℃ %
Experimental Result 240 200 40 50
Voltz Model - - - 45
Herz and Marin Basic Model 250 190 60 55
Salomons’ Compressed Oxygen Model 255 195 60 45
CO Island Model - - - 55
O Island Model 225 220 5 55
CO-O Island Model 225 220 5 52
Dumont Postulation - - - 0.085
69

4.2 Hysteresis (with respect to solid temperature)

Figure 4.2.1: CO Conversion against solid temperature plot from the experimental
results of Ye et. al.(2012)

Figure 4.2.2: CO Conversion against solid temperature plot for Voltz Model
70

Figure 4.2.3: CO Conversion against solid temperature plot for Herz and Marin Basic
Model

Figure 4.2.4: CO Conversion against solid temperature plot for Salomons' Compressed
Oxygen Model
71

Figure 4.2.5: CO Conversion against solid temperature plot for CO Island Model

Figure 4.2.6: CO Conversion against solid temperature plot for O Island Model
72

Figure 4.2.7: CO Conversion against solid temperature plot for CO-O Island Model

Figure 4.2.8: CO Conversion against solid temperature plot for Dumont Postulation
73

Generally, the same observations and conclusions can be made with this plot. However, it can
be seen that the simulated result gives a solid ignition temperature 20 to 40℃ lower than the
experimental value. This could mean that heat transfer between the gas and solid washcoat is
not simulated accurately in our models. More evidence can be found when we look at the
temperature plot in the next section.

Table 4.2.1: Summary of Hysteresis Characteristics on the simulation done (Solid


Temperature at 3mm)

Ignition Extinction Width Conversion


Temperature Temperature Of
Hysteresis
℃ ℃ ℃ %
Experimental Result 270 210 60 50
Voltz Model - - - 45
Herz and Marin Basic Model 250 195 55 55
Salomons’ Compressed Oxygen Model 260 200 60 45
CO Island Model - - - 55
O Island Model 230 220 10 55
CO-O Island Model 230 225 5 52
Dumont Postulation - - - 0.085
74

4.3 Temperature Profile

The change in temperature profile of the inlet, outlet, and 1mm, 2mm, 3mm of the solid is
plotted against time and will provide us key information about the accuracy of the simulation.

Figure 4.3.1: Temperature against time plot from Ye et . al.(2012)

Figure 4.3.2: Temperature plot for Voltz Model


75

Figure 4.3.3: Temperature plot for Herz and Marin Basic Model

Figure 4.3.4: Temperature plot for Salomons' Compressed Oxygen Model


76

Figure 4.3.5: Temperature plot for CO Island Model

Figure 4.3.6: Temperature plot for O Island Model


77

Figure 4.3.7: Temperature plot for CO and O Island Model

Figure 4.3.8: Temperature plot for Dumont Postulation


78

Generally, the outlet temperature is lower or almost the same as the inlet temperature for all
simulated models. This is not true as catalytic combustion is an exothermic reaction, so the gas
coming out of the monolith should be a lot hotter than the inlet. This is evident from Figure
3.6.9. Hence, the temperature plots exposed a flaw in this simulation which is poorly simulated
heat and mass transfer characteristics. Better correlations needed to be chosen instead.

Moreover, the fact that the solid wall temperature does not agree well with the
experimental value (20 to 40℃ lower), implies that maybe the heat of reaction of the surface
reaction should be higher than the one used in the simulation (-282.55 kJ/mol) . Further
research needs to be done in selecting better parameters.

4.4 Surface Coverage


Another key result from the simulation is how the surface coverage of each model varies over
time. Surface coverage of each species is plotted against time.

Figure 4.4.1: Fractional Coverage of CO against time for Herz and Marin Basic Model
79

Figure 4.4.2: Fractional Coverage of O against time for Herz and Marin Basic Model

Figure 4.4.3: Fractional Coverage of Vacant Sites against time for Herz and Marin Basic
Model
80

Figure 4.4.4: Fractional Coverage of CO against time for Salomons' Compressed


Oxygen Model

Figure 4.4.5: Fractional Coverage of O against time for Salomons' Compressed Oxygen
Model
81

Figure 4.4.6: Fractional Coverage of Compressed Oxygen against time for Salomons'
Compressed Oxygen Model

Figure 4.4.7: Fractional Coverage of Vacant Sites against time for Salomons'
Compressed Oxygen Model
82

Figure 4.4.8: Fractional Coverage of CO against time for CO Island Model

Figure 4.4.9: Fractional Coverage of O against time for CO Island Model


83

Figure 4.4.10: Fractional Coverage of Vacant Site against time for CO Island Model
84

Figure 4.4.11: Fractional Coverage of CO against time for O Island Model

Figure 4.4.12: Fractional Coverage of O against time for O Island Model


85

Figure 4.4.13: Fractional Coverage of Vacant Site against time for O Island Model
86

Figure 4.4.14: Fractional Coverage of CO against time for CO-O Island Model

Figure 4.4.15: Fractional Coverage of O against time for CO-O Island Model
87

Figure 4.4.16: Fractional Coverage of Vacant Sites against time for CO-O Island Model
88

Figure 4.4.17: Fractional Coverage of CO against time for Dumont Postulation

Figure 4.4.18: Fractional Coverage of O against time for Dumont Postulation


89

Figure 4.4.19: Fractional Coverage of Vacant Sites against time for Dumont Postulation
90

The reason why Dumont Postulation failed to give the expected conversion values are clear
when we look at Figure 4.4.17 where CO coverage is 1.0 at throughout the simulation. This
would let us draw the following possibilities:

a) The surface reaction is not happening at a fast enough rate.


b) Desorption of CO is not happening (CO Poisoning).

The kinetic parameters can be fine-tuned to provide a better fit with the hysteresis curve.

The fractional coverage behaviour in the O Island Model and CO-O Island Model tells us
that ignition is not happening in the same manner observed in the fractional coverage of the
CO and O in both Herz and Marin Basic Model and Salomons’ Compressed Oxygen Model.

First of all, the time at which ignition occur is different for the models. For the lesser models
(O Island Model and CO-O Island Model), it is apparent from Figure 4.4.11 and Figure 4.4.14
that ignition is happening at around t = 200s while for the greater models (Herz and Marin
Basic Model and Salomons’ Compressed Oxygen Model), it is happening at around t = 400s
based on Figure 4.4.1 and Figure 4.4.4. However, one similarity between the models is the
minimum CO coverage attained in the process. For the lesser models, 0.7 is the minimum
coverage while for the greater models, 0.65. A strong hypothesis that we can draw from this
observation is that coverage of CO and O atoms can tell us the information about the state of
conversion of the models since all models reported similar maximum conversion and this is
correlated with similar maximum/minimum CO and O coverage. A possible suggestion to
improve the hysteresis behaviour of the lesser models is to try to delay the ignition by tuning
the kinetic parameters of the elementary steps such as pre-exponential factors, activation
energies and effective diffusivity of the molecules(Lundberg et al., 2016).

Also, looking into the fractional coverage of compressed oxygen in Figure 4.4.6, we
realised that only 0.008% is covered by compressed oxygen. This indicates that only a small
amount of CO will adsorb on sites that is preoccupied with an O atom. However, considerable
amounts of CO should be able to adsorb onto the O ad layer.(Engel et al., 1979; Mongeot et
al., 1998) Improvements can still be made in Salomons’ Compressed Oxygen Model by tuning
the parameters for Equation 3.4.17, Equation 3.4.18, and Equation 3.4.19.
91

4.5 Fine Tuning


Compared to Herz and Marin Basic Model and Salomons’ Compressed Oxygen Model, the
hysteresis was less significant for the CO-O Island Model. It is thought that the kinetic
parameters for CO-O Island Model can be tuned such that the hysteresis closely resembles the
experimental hysteresis. From several trials of kinetic parameters, it has been identified that
the bulk diffusivity coefficient of the species can be adjusted to modify the hysteresis.

Figure 4.5.1: Conversion Plot of CO-O Island Model after increasing kd by 10 times
92

Figure 4.5.2: Conversion Plot of CO-O Island Model after increasing kd by 2 times

Figure 4.5.3: Conversion Plot of CO-O Island Model after reducing kd by 10 times
93

Figure 4.5.4: Conversion Plot of CO-O Island Model after reducing kd by half

Table 4.5.1: Summary of Hysteresis Characteristics after tuning kd

Ignition Extinction Width Conversion


Temperature Temperature Of
Hysteresis
℃ ℃ ℃ %
Experimental Result 240 200 40 50
10𝐷 250 250 0 100
2𝐷 230 230 0 75
0.5𝐷 215 209 6 30
0.1𝐷 178 167 11 4

As observed in Figure 4.5.1 and Figure 4.5.2, increasing the bulk diffusion coefficient constant
of the species leads to the model behaving more like the Voltz Model. In other words, the
model is showing the characteristic curves of a global kinetic model. Similarities between
Figure 4.5.1 and Figure 4.5.2, and Figure 4.5.5 can be observed.
94

Figure 4.5.5: Simualtion done by Daneshvar et. al. using Global Kinetic Model
(Daneshvar et al., 2017)
On the other hand, reducing the diffusivity constant of the species would result in a
more pronounced hysteresis character as evident in Figure 4.5.3 and Figure 4.5.4. These
observations suggest that hysteresis of the conversion is not only dependent on the mechanics
of the reaction, it is also strongly influenced by the washcoat properties. Further research needs
to be done on this subject.
95

CHAPTER 5

CONCLUSION AND RECOMMENDATIONS

5.1 Conclusion
In this project, simulation of CO oxidation in a thin slice (5mm) DOC has been performed.
Voltz Model, Herz and Marin Basic Model, Salomons, Compressed Oxygen Model, CO
Island Model, O Island Model, CO-O Island Model, and Dumont Postulation are investigated.
The following conclusions are drawn:

 Herz and Marin Basic Model and Salomons’ Compressed Oxygen Model gives the
best results with the latter being marginally better in terms of ignition and extinction
temperature.
 The Island Models (with an exception of Dumont Postulation) gives good fit for
conversion but cannot simulate a good hysteresis curve.
 The temperature profile of the models does not match with experimental data. This
might be due to poor heat and mass transfer modelling or incorrect value of heat of
reaction for catalytic combustion chosen.
 Increasing the diffusivity constant of the species will lead the CO-O Island Model to
behave like a global model i.e. no hysteresis is present while reducing it will result in
hysteresis. This leads us to hypothesise that hysteresis might be due to the mass
transfer limitation of the washcoat.
96

5.2 Recommendation
Below are several recommendations that can be implemented in future research.

 To use other programs such as ANSYS and DETCHEM where the CFD is used and a
better approximation could be achieved with models of higher dimensions.
 To fine tune effective diffusivity of the species.(Lundberg et al., 2015)
 To implement the use of mechanistic kinetics for propylene oxidation.
 Robustness and stability of the better models (Herz and Marin Basic Model, and
Salomons’ Compressed Oxygen Model) can be tested using experimental data from
other researchers.
97

REFERENCES

Bonzel, H. P. and Ku, R. (1973) ‘On the kinetics of oxygen adsorption on a Pt(111) surface’,
Surface Science, 40(1), pp. 85–101. doi: 10.1016/0039-6028(73)90053-8.

Campbell, C. T. et al. (1980) ‘A molecular beam study of the catalytic oxidation of CO on a


Pt(111) surface’, The Journal of Chemical Physics, 73(11), pp. 5862–5873. doi:
10.1063/1.440029.

Casapu, M. et al. (2017) ‘Origin of the Normal and Inverse Hysteresis Behavior during CO
Oxidation over Pt/Al2O3’, ACS Catalysis, 7(1), pp. 343–355. doi: 10.1021/acscatal.6b02709.

Chapra, S. C. and Canale, R. P. (2010) Numerical Methods for Engineers. 7th edn. McGraw-
Hill Education.

Chatterjee, D., Deutschmann, O. and Warnatz, J. (2001) ‘Detailed Surface Reaction


Mechanism in a Three-Way Catalyst’, Faraday Discussions, (119), pp. 371–384. doi:
10.1039/b101968f.

Dadi, R. K., Luss, D. and Balakotaiah, V. (2016) ‘Dynamic hysteresis in monolith reactors
and hysteresis effects during co-oxidation of CO and C2H6’, Chemical Engineering Journal.
Elsevier B.V., 297, pp. 325–340. doi: 10.1016/j.cej.2016.03.139.

Daneshvar, K. et al. (2017) ‘Experimental and modeling study of CO and hydrocarbons light-
off on various Pt-Pd/Γ-Al2O3diesel oxidation catalysts’, Chemical Engineering Journal.
Elsevier B.V., 323, pp. 347–360. doi: 10.1016/j.cej.2017.04.078.

Dumont, M., Poriaux, M. and Dagonnier, R. (1986) ‘ON SURFACE REACTION KINETICS
IN THE PRESENCE OF ISLANDS’, Surface Science, 169(2–3), pp. 307–310. doi:
10.1016/0039-6028(86)90598-4.

Ehsasi, M. et al. (1989) ‘Steady and nonsteady rates of reaction in a heterogeneously


catalyzed reaction : Oxidation of CO on platinum , experiments and simulations Steady and
nonsteady rates of reaction in a heterogeneously catalyzed reaction : Oxidation of CO on
platinum , experiments and simulations’, 4949. doi: 10.1063/1.456736.
98

Eigenberger, G. (1978) ‘Kinetic instabilities in heterogeneously catalyzed reactions-II.


Oscillatory instabilities with langmuir-type kinetics’, Chemical Engineering Science, 33(9),
pp. 1263–1268. doi: 10.1016/0009-2509(78)85092-1.

Engel, T. (1978) ‘A molecular beam investigation of He, CO, and O2 scattering from
Pd(111)’, The Journal of Chemical Physics, 69(1), p. 373. doi: 10.1063/1.436363.

Engel, T. et al. (1979) ‘Elementary Steps in the Catalytic Oxidation of Carbon Monoxide on
Platinum Metals’, Advances in Catalysis, Volume 28, pp. 1–78. Available at:
http://www.sciencedirect.com/science/article/B7CS3-4S875SR-
6/2/84e0400997c8fec81806548661ee3d5b.

Ertl, G., Neumann, M. and Streit, K. M. (1977) ‘Chemisorption of CO on the Pt(111)


surface’, Surface Science, 64(2), pp. 393–410. doi: 10.1016/0039-6028(77)90052-8.

Gland, J. L. et al. (1983) ‘Carbon monoxide oxidation on the Pt ( 111 ) surface : Temperature
programmed reaction of coadsorbed atomic oxygen and carbon monoxide Carbon monoxide
oxidation on the Pt ( 111 ) surface : Temperature programmed reaction of coadsorbed atomic
oxygen and car’, 963(111). doi: 10.1063/1.444801.

Graham, W. R. C. and Lynch, D. T. (1987) ‘ISLAND MODELS AND THE CATALYTIC


OXIDATION OF CARBON MONOXIDE’, Surface Science, 187, pp. 633–638.

Gundlapally, S. R. and Balakotaiah, V. (2013) ‘Analysis of the effect of substrate material on


the steady-state and transient performance of monolith reactors’, Chemical Engineering
Science. Elsevier, 92, pp. 198–210. doi: 10.1016/j.ces.2013.01.051.

Haaland, D. M. and Williams, F. L. (1982) ‘Simultaneous Measurement of CO Oxidation


Rate and Surface Coverage on Pt/Al203 Using Infrared Spectroscopy : Rate Hysteresis and
CO Island Formation’, Journal of Catalysis, 76, pp. 450–465. doi: 10.1016/0021-
9517(82)90274-3.

Herz, K. and Marin, S. P. (1980) ‘Chemistry Models Supported of Carbon Monoxide


Platinum Catalysts Oxidation on’, 296, pp. 281–296.

Hlaváček, V. and Votruba, J. (1979) Hysteresis and Periodic Activity Behavior in Catalytic
Chemical Reaction Systems, Advance in Catalysis. doi: 10.1016/S0360-0564(08)60054-1.
99

Hopster, H., Ibach, H. and Comsa, G. (1977) ‘Catalytic oxidation of carbon monoxide on
stepped platinum(111) surfaces’, Journal of Catalysis, 46(1), pp. 37–48. doi: 10.1016/0021-
9517(77)90133-6.

Kisliuk, P. J. (1957) ‘The sticking probability of gases chemisorbed on the surfaces of solids’,
J.Phys.Chem.Solids, 3, pp. 95–101.

Langmuir, I. (1922) ‘The mechanism of the catalytic action of platinum in the reactions 2Co
+ O2= 2Co2 and 2H2+ O2= 2H2O’, Transactions of the Faraday Society, 17, p. 621. doi:
10.1039/tf9221700621.

Lundberg, B. et al. (2015) ‘Model-based experimental screening for DOC parameter


estimation’, Computers and Chemical Engineering. Elsevier Ltd, 74, pp. 144–157. doi:
10.1016/j.compchemeng.2015.01.004.

Lundberg, B. et al. (2016) ‘DOC modeling combining kinetics and mass transfer using inert
washcoat layers’, Applied Catalysis B: Environmental. Elsevier B.V., 191, pp. 116–129. doi:
10.1016/j.apcatb.2016.03.024.

McCabe, R. W. and Schmidt, L. D. (1977) ‘binding states of CO and H2 on clean and


oxidized (111)Pt’, Surface Science, 65(1), pp. 189–209. doi: 10.1016/0039-6028(77)90301-6.

Mongeot, F. B. De et al. (1998) ‘CO adsorption and oxidation on bimetallic Pt / Ru ( 0001 )


surfaces – a combined STM and TPD / TPR study’, 411, pp. 249–262.

Mukadi, L. S. and Hayes, R. E. (2002) ‘Modelling the three-way catalytic converter with
mechanistic kinetics using the Newton – Krylov method on a parallel computer’, 26, pp. 439–
455.

Mukesh, D. et al. (1984) ‘Island models and the catalytic oxidation of carbon monoxide and
carbon monoxide-olefin mixtures d. mukesh’, 138, pp. 237–257.

Nibbelke, R. H. et al. (1997) ‘Kinetic Study of the CO Oxidation over Pt / γ -Al 2 O 3 and Pt
/ Rh / CeO 2 / γ -Al 2 O 3 in the Presence of H 2 O and CO 2’, 373, pp. 358–373.

Raj, R., Harold, M. P. and Balakotaiah, V. (2015) ‘Steady-state and dynamic hysteresis
effects during lean co-oxidation of CO and C 3 H 6 over Pt / Al 2 O 3 monolithic catalyst’,
CHEMICAL ENGINEERING JOURNAL. Elsevier B.V., 281, pp. 322–333. doi:
100

10.1016/j.cej.2015.06.057.

Razon, L. F. and Schmitz, R. A. (1986) ‘Intrinsically Unstable Behavior during the Oxidation
of Carbon Monoxide on Platinum’, Catalysis Reviews, 28(1), pp. 89–164. doi:
10.1080/03602458608068086.

Russell, A. and Epling, W. S. (2011) ‘Diesel oxidation catalysts’, Catalysis Reviews - Science
and Engineering, 53(4), pp. 337–423. doi: 10.1080/01614940.2011.596429.

Salomons, S. et al. (2007) ‘On the use of mechanistic CO oxidation models with a platinum
monolith catalyst’, Applied Catalysis B: Environmental, 70(1–4), pp. 305–313. doi:
10.1016/j.apcatb.2006.01.022.

Salomons, S. (2008) ‘Kinetic Models for a Diesel Oxidation Catalyst’, Dissertation


University of Alberta, pp. 1–280. doi: 12(4):294-301,1973.

Sjoblom, J., Ab, S. and Creaser, D. (2014) ‘Parameter Estimation of a DOC from Engine Rig
Experiments with a Discretized Catalyst Washcoat Model Björn Westerberg’. doi:
10.4271/2014-01-9049.

Voltz, S. E. et al. (1973) ‘Kinetic Study of Carbon Monoxide and Propylene Oxidation on
Platinum Catalysts’, Industrial & Engineering Chemistry Product Research and
Development, 12(4), pp. 294–301. doi: 10.1021/i360048a006.

Wicke, E., Kummann, P. and Schiefler, J. (1980) ‘Unstable and Oscillatory Behaviour in
Heterogeneous Catalysis’, 797(1962), pp. 315–323.

Ye, S. et al. (2012) ‘Catalyst “light-off” experiments on a diesel oxidation catalyst connected
to a diesel engine-Methodology and techniques’, Chemical Engineering Research and
Design. Institution of Chemical Engineers, 90(6), pp. 834–845. doi:
10.1016/j.cherd.2011.10.003.

Ziff, R. M., Gulari, E. and Barshad, Y. (1986) ‘Kinetic phase transitions in an irreversible
surface-reaction model’, Physical Review Letters, 56(24), pp. 2553–2556. doi:
10.1103/PhysRevLett.56.2553.
101

APPENDICES

5 APPENDIX A: Translation of Monolith Model Using Method of Lines

A.1 Mass Balance

A.1.1 Gas Phase

For CO:

𝜕𝐶 , 𝜕 𝜕𝐶 , 𝜕 𝑣𝐶 , 4
= 𝐷 , , − − 𝑘 , 𝐶 , −𝐶 ,
𝜕𝑡 𝜕𝑧 𝜕𝑧 𝜕𝑧 𝐷

Equation A.1
The following assumptions are made:

𝐷 , , and 𝑣 are assumed to be constant throughout the channel length. This assumption
is valid since we are considering a very small channel length where the decrease in both
velocity and dispersion coefficient is negligible.

The mass balance then becomes:

𝜕𝐶 , 𝜕 𝐶 , 𝜕 𝐶 , 4
=𝐷 , , −𝑣 − 𝑘 , 𝐶 , −𝐶 ,
𝜕𝑡 𝜕𝑧 𝜕𝑧 𝐷

Equation A.2

Using the explicit method the following is obtained:

𝜕𝐶 , 𝐶 ,, −𝐶 ,,
=
𝜕𝑡 𝑑𝑡

Equation A.3
102

𝜕 𝐶 , 𝐶 , , − 2𝐶 , , + 𝐶 ,,
=
𝜕𝑧 𝑑𝑧

Equation A.4
𝜕𝐶 , 𝐶 , ,−𝐶 ,,
=
𝜕𝑧 𝑑𝑧

Equation A.5
where,

𝑑𝑡 is the differential time interval

𝑑𝑧 is the differential channel length

𝑗 denotes the time position

𝑖 denotes the spatial position

Then substituting these into Equation A.2, the PDE is then transformed into:

,, ,, , , ,, ,,
=𝐷 , ,

, , ,,
−𝑣

− 𝑘 , 𝐶 , −𝐶 ,

Equation A.6

Since the concentration of CO is given as ppm in the experiment, it is necessary to change the
units of concentration in the equation as well. Equation A.6 is divided by the total concentration
of the gas phase.

,, ,, , , ,, ,,
=𝐷 , ,

, , ,,
−𝑣

− 𝑘 , 𝐶 , −𝐶 ,
103

This would lead to the mass balance equation to be in terms of mol fraction, the unit of ppm.

,, ,, , , ,, ,,
=𝐷 , ,

, , ,,
−𝑣

− 𝑘 , 𝑌 , −𝑌 ,

Equation A.7

Similar operations would be performed on the other components leading to:

O2 balance

,, ,, , , ,, ,,
=𝐷 , ,

, , ,,
−𝑣

− 𝑘 , 𝑌 , −𝑌 ,

Equation A.8

CO2 balance

,, ,, , , ,, ,,
=𝐷 , ,

, , ,,
−𝑣

− 𝑘 , 𝑌 , −𝑌 ,

Equation A.9
104

A.1.2 Solid Phase

𝜕𝐶 , 𝑊 𝜕 𝜕𝐶 ,
𝜀 = 𝑘 , 𝐶 , −𝐶 , + 𝐷 , + 𝜉𝑅
𝜕𝑡 𝐴 𝜕𝑧 𝜕𝑧

Equation A.10
The following assumptions are made:

𝐷 , , is assumed to be constant throughout the channel length. This assumption is valid


since we are considering a very small channel length where the decrease in dispersion
coefficient is negligible.

𝜕𝐶 , 𝑊 𝜕 𝐶 ,
𝜀 = 𝑘 , 𝐶 , −𝐶 , +𝐷 , , + 𝜉𝑅
𝜕𝑡 𝐴 𝜕𝑧

Equation A.11
The reaction term of CO is the net change in CO due to adsorption and desorption.

𝜉𝑅 = −𝑟 , +𝑟 ,

Equation A.12
Generally, the reaction term is expressed as:

−𝑟 , +𝑟 ,

1000𝑅𝑇 1000(𝐸 − 𝛽𝜃 )
=− 𝜎𝑆 𝐶 , 𝑓 (𝜃 , 𝜃 ) + 𝑘 𝑒𝑥𝑝 − 𝜃
2𝜋𝑀 𝑅𝑇

Equation A.13

where,

𝑓(𝜃 ) is a function that is dependent on the reaction model used. For all of our models, they
will be in the form of:

𝑓 (𝜃 , 𝜃 ) = 1 − 𝜃 −𝜃

Equation A.14
105

Similar to the one dimensional gas phase mass balance, we need to change the concentration
term into mol fractions

With the ideal gas assumption, the total concentration of gases in the washcoat is given by:

𝑃
𝐶 =
𝑅𝑇

Equation A.15

Also,

𝐶 ×𝑌 =𝐶

Equation A.16

Therefore,

𝑃
𝐶 =𝑌
𝑅𝑇

Equation A.17

Then substituting it into Equation A.13:

−𝑟 , +𝑟 ,

1000𝑅𝑇 𝑃 𝑃 1000(𝐸 − 𝛽𝜃 )
=− 𝜎𝑆 𝑌 𝑓(𝜃 , 𝜃 ) + 𝑘 𝑒𝑥𝑝 − 𝜃
2𝜋𝑀 𝑅𝑇 𝑅𝑇𝑠 𝑅𝑇

Equation A.18

Then, into Equation A.11:

𝜕𝐶 , 𝑊 𝜕 𝐶
𝜀 = 𝑘 , 𝐶 , −𝐶 , +𝐷 ,
𝜕𝑡 𝐴 𝜕𝑧

1000𝑅𝑇 𝑃
− 𝜎𝑆𝐶𝑂 𝑌𝐶𝑂,𝑠 𝑓 (𝜃 , 𝜃 )
2𝜋𝑀𝐶𝑂 𝑅𝑇𝑠 1 𝐶𝑂 𝑂

𝑃 1000(𝐸2 − 𝛽𝜃𝐶𝑂 )
+ 𝑘2 𝑒𝑥𝑝 − 𝜃𝐶𝑂
𝑅𝑇 𝑅𝑇

Equation A.19
106

Using the explicit method the following is obtained:

𝜕𝐶 , 𝐶 ,, −𝐶 ,,
=
𝜕𝑡 𝑑𝑡

Equation A.20
𝜕 𝐶 , 𝐶 , , − 2𝐶 ,, +𝐶 ,,
=
𝜕𝑧 𝑑𝑧

Equation A.21

Then substituting it into Equation A.19:

𝐶 ,, −𝐶 ,,
𝜀
𝑑𝑡
𝑊
= 𝑘 , 𝐶 , ,, −𝐶 , ,,
𝐴
𝐶 , , − 2𝐶 ,, +𝐶 ,,
+𝐷 , ,
𝑑𝑧

1000𝑅𝑇 𝑃
− 𝜎𝑆 𝑌 , (1 − 𝜃 −𝜃 )
2𝜋𝑀 𝑅𝑇

1000(𝐸 − 𝛽𝜃 )
+ 𝑘 𝑒𝑥𝑝 − 𝜃
𝑅𝑇

Equation A.22

Then Equation A.22 is divided by CT to transform concentration into mol fraction


𝑌 ,, −𝑌 ,,
𝑑𝑡

1 𝑊
= 𝑘 , 𝑌 , ,, −𝑌 , ,,
𝜀 𝐴

𝑌 , , − 2𝑌 ,, +𝑌 ,, 1000𝑅𝑇
+𝐷 , , − 𝜎𝑆 𝑌 , (1
𝑑𝑧 2𝜋𝑀

1000(𝐸 − 𝛽𝜃 )
−𝜃 − 𝜃 ) + 𝑘 𝑒𝑥𝑝 − 𝜃
𝑅𝑇

Equation A.23
107

Applying the same operation on the O2 solid phase mass balance:

𝑌 ,, −𝑌 ,,
𝑑𝑡

1 𝑊 𝑌 , , − 2𝑌 ,, +𝑌 ,,
= 𝑘 , 𝑌 , ,, −𝑌 , ,, +𝐷 , ,
𝜀 𝐴 𝑑𝑧

1000𝑅𝑇
−2 𝜎𝑆 𝑌 𝑓 (𝜃 , 𝜃 )
2𝜋𝑀

Equation A.24

𝑓 (𝜃 , 𝜃 ) is dependent on the model used

Table A.1: 𝒇𝟐 (𝜽𝑪𝑶 , 𝜽𝑶 ) for each model

Reaction Model 𝒇𝟐 (𝜽𝑪𝑶 , 𝜽𝑶 )

Basic Model 1 − 2𝜃
1− 𝜃 − 2𝜃
1−𝜃

Salomons’ Compressed Oxygen 1 − 2𝜃


1− 𝜃 − 2𝜃
1−𝜃

CO Island Model 1 − 2𝜃
1 − 1 − 𝜃 𝜃 − 2𝜃
(1 − 𝜃 )

O Island Model 1 − 2𝜃
1 − 1 − 𝜃 𝜃 − 2𝜃
(1 − 𝜃 )

CO-O Island Model 1 − 2𝜃


1− 𝜃 − 2𝜃
1−𝜃

Dumont Postulation Island Model 1 − 2𝜃


1− 𝜃 − 2𝜃
1−𝜃
108

Applying the same operation on the CO2 solid phase mass balance:

𝑌 ,, −𝑌 ,,
𝑑𝑡
1 𝑊
= 𝑘 , 𝑌 , ,, −𝑌 , ,,
𝜀 𝐴
𝑌 , , − 2𝑌 ,, +𝑌 ,,
+𝐷 , ,
𝑑𝑧
1000𝐸 𝑌 𝑌
+ 𝑘 exp − 𝑓 (𝜃 , 𝜃 ) + 𝑘 ,
𝑅𝑇 𝑅(𝜃)

Equation A.25

𝑓 (𝜃 , 𝜃 ) is dependent on the model used

Table A.2: 𝒇𝟑 (𝜽𝑪𝑶 , 𝜽𝑶 ) for each model

Reaction Model 𝒇𝟑 (𝜽𝑪𝑶 , 𝜽𝑶 )

Basic Model 𝜃 𝜃

Salomons’ Compressed Oxygen 𝜃 𝜃

CO Island Model 𝜃
𝜃
(1 − 𝜃 )

O Island Model 𝜃
𝜃
(1 − 𝜃 )

CO-O Island Model 𝜃 𝜃

Dumont Postulation Island Model 𝜃 (1 − 𝜃 )𝜃 (1 − 𝜃 )


109

A.2 Energy Balance

A.2.1 Gas Phase

𝜕𝑇 𝜕 𝜕𝑇 𝜕𝑇 4
𝜌 𝐶 = 𝐾 , −𝜌 𝐶 𝑣 − 𝑘 𝑇, − 𝑇
𝜕𝑡 𝜕𝑧 𝜕𝑧 𝜕𝑧 𝐷

Equation A.26
The following assumptions are made:

𝐾 , is assumed to be constant throughout the channel length. This assumption is valid


since we are considering a very small channel length where the decrease in dispersion
coefficient is negligible.

Then,

𝜕𝑇 𝜕 𝑇 𝜕𝑇 4
𝜌 𝐶 =𝐾 , −𝜌 𝐶 𝑣 − 𝑘 𝑇, − 𝑇
𝜕𝑡 𝜕𝑧 𝜕𝑧 𝐷

Equation A.27

Rearranging yields:

𝜕𝑇 𝐾 , 𝜕 𝑇 𝜕𝑇 4
= −𝑣 − 𝑘 𝑇, − 𝑇
𝜕𝑡 𝜌 𝐶 𝜕𝑧 𝜕𝑧 𝐷 𝜌 𝐶

Equation A.28
Using the explicit method the following is obtained:

𝜕𝑇 𝑇,, −𝑇,,
=
𝜕𝑡 𝑑𝑡

Equation A.29
𝜕 𝑇 𝑇, , − 2𝑇 , , + 𝑇 , ,
=
𝜕𝑧 𝑑𝑧

Equation A.30
𝜕𝑇 𝑇, ,−𝑇,,
=
𝜕𝑧 𝑑𝑧

Equation A.31
110

Then Equation A.28 is transformed into:

𝑇,, −𝑇,, 𝐾 , 𝑇, , − 2𝑇 , , + 𝑇 , , 𝑇, ,−𝑇,, 4


= −𝑣 − 𝑘 𝑇, , −𝑇, ,
𝑑𝑡 𝜌 𝐶 𝑑𝑧 𝑑𝑧 𝐷 𝜌 𝐶

Equation A.32

A.2.2 Solid Phase

𝜕𝑇 𝑊 𝜕 𝜕𝑇
𝜌 𝐶 = 𝑘 𝑇 −𝑇 + 𝐾 , + (∆𝐻𝑅)
𝜕𝑡 𝐴 𝜕𝑧 𝜕𝑧

Equation A.33
The following assumptions are made:

𝐾 , is assumed to be constant throughout the channel length. This assumption is valid


since we are considering a very small channel length where the decrease in dispersion
coefficient is negligible.

Then,

𝜕𝑇 𝑊 𝜕 𝑇
𝜌𝐶 = 𝑘 𝑇 −𝑇 +𝐾 , + (∆𝐻𝑅)
𝜕𝑡 𝐴 𝜕𝑧

Equation A.34

Rearranging yields:

𝜕𝑇 𝑊 𝑘 𝐾 , 𝜕 𝑇 1
= 𝑇 −𝑇 + + (∆𝐻𝑅)
𝜕𝑡 𝐴 𝜌𝐶 𝜌𝐶 𝜕𝑧 𝜌𝐶

Equation A.35
Using the explicit method the following is obtained:

𝜕𝑇 𝑇,, −𝑇,,
=
𝜕𝑡 𝑑𝑡

Equation A.36
111

𝜕 𝑇 𝑇, , − 2𝑇 , , + 𝑇 , ,
=
𝜕𝑧 𝑑𝑧

Equation A.37

Then the PDE is transformed into:

𝑇,, −𝑇,, 𝑊 𝑘 𝐾 , 𝑇, , − 2𝑇 , , + 𝑇 , , 1
= 𝑇, , −𝑇, , + + (∆𝐻𝑅)
𝑑𝑡 𝐴 𝜌𝐶 𝜌𝐶 𝑑𝑧 𝜌𝐶

Equation A.38

The reaction term is expressed as

(∆𝐻𝑅) = 𝑟 × −∆𝐻 , + −𝑟 × −∆𝐻 , ×𝐻×𝐴

Equation A.39

where

𝑟 is the rate of reaction of the surface reaction with CO

−∆𝐻 , is the heat of reaction of the surface reaction with CO, J/mol

𝑟 is the rate of reaction of the surface reaction with C3H6 given by:

𝑌 𝑌
𝑟 =𝑘 ,
𝑅(𝜃)

Equation A.40
𝑹(𝜽) is a term which takes into account of the inhibition effects of CO, C3H6, and NO on the
reaction rates.

.
𝑅(𝜃) = 1 + 𝑘 𝑌 +𝑘 𝑌 × 1+𝑘 𝑌 𝑌 × [1 + 𝑘 𝑌 ]

Equation A.41

−∆𝐻 , is the heat of reaction of the surface reaction with C3H6 , J/mol

𝐻 is the mol of surface metal atoms per surface area, mol/m2

𝐴 is the specific surface area per volume of catalyst, m2/m3


112

Then,

𝑇,, −𝑇,, 𝑊 𝑘 𝐾 , 𝑇, , − 2𝑇 , , + 𝑇 , ,
= 𝑇,, −𝑇,, +
𝑑𝑡 𝐴 𝜌𝐶 𝜌𝐶 𝑑𝑧
𝐻𝐴 1000𝐸 𝑌 𝑌
+ 𝑘 exp − 𝑓 (𝜃 , 𝜃 ) × −∆𝐻 , +𝑘 ,
𝜌𝐶 𝑅𝑇 𝑅(𝜃)

× −∆𝐻 ,

Equation A.42
𝑓 (𝜃 , 𝜃 ) is dependent on the reaction model and is available in Table A.2

5
113

5 APPENDIX B: List of Reaction Parameters

Table B.1: Monolith Properties

Monolith Properties
Parameter Symbol Value Unit Ref.
Channel Length 𝐿 5 𝑚𝑚 Ye, 2012
Channel Radius 𝐷 410 𝜇𝑚 Ye, 2012
Washcoat Thickness 𝑊 100 𝜇𝑚 Ye, 2012
Catalyst Loading 𝐿 2.0 × 10 𝑚𝑜𝑙 Ye, 2012
𝑔𝐷𝑂𝐶
Density of Washcoat 𝜌 1300 𝑘𝑔 Ye, 2012
𝑚
BET Surface Area 𝑆 160 𝑚 Ye, 2012
𝑔
Average Pore Diameter 𝐷 12.25 𝑛𝑚 Ye, 2012
Porosity 𝜀 0.55 Mukadi, 2002

Table B.2: Fluid Properties

Fluid Properties
Parameter Symbol Value Unit Ref.
Ideal Gas Constant 𝑅 8.314 𝐽 Raj 2015
𝑚𝑜𝑙𝐾
114

Table B.3: Kinetic Parameters of the Voltz Model

Temperature Dependent Parameters for Voltz Model


Parameter Symbol Equation Unit Ref.
𝑟 , 𝑘 −22600 1 Voltz,
1.83 × 10 exp
𝑇 𝑠[𝑂 ] 1973
𝑟 𝑘 −10800 1 Voltz,
3.80 × 10 exp
𝑇 𝑠[𝐶 𝐻 ]
1973
𝑅(𝜃) 𝑘 1730 1 Voltz,
6.55 × 10 exp
𝑇 [𝐶𝑂 ]
1973
𝑅(𝜃) 𝑘 650 1 Voltz,
2.08 × 10 exp
𝑇 [𝐶 𝐻 ]
1973
𝑅(𝜃) 𝑘 20900 1 Voltz,
3.98 × 10 exp
𝑇 [𝐶 𝐻 ][𝐶𝑂 ]
1973

Table B.4: Kinetic Parameters of the Basic Model

Constant Parameters for Basic Model


Parameter Symbol Value Unit Ref.
Sticking Coefficient 𝑆 0.15 Salomons, 2008
of O2
Surface Reaction 𝑘 1.0 × 10 1 Herz and Marin, 1980
𝑠
𝐸 104.6 𝑘𝐽 Salomons, 2008
𝑚𝑜𝑙
𝛽 46.024 𝑘𝐽 Salomons, 2008
𝑚𝑜𝑙
Heat of Reaction ∆𝐻 −282.55 𝑘𝐽 Salomons, 2008
𝑚𝑜𝑙
115

Table B.5: Kinetic Parameters of the Compressed Oxygen Model

Constant Parameters for Compressed Oxygen Model


Parameter Symbol Value Unit Ref.
𝑘 𝐴 1 × 10 𝑚𝑜𝑙 Salomons, 2008
𝑠
𝑐𝑚
𝐸 50 𝑘𝐽 Salomons, 2008
𝑚𝑜𝑙
𝑘 𝐴 1 × 10 𝑚𝑜𝑙 Salomons, 2008
𝑠
𝑐𝑚
𝐸 115 𝑘𝐽 Salomons, 2008
𝑚𝑜𝑙
𝑘 𝐴 1 × 10 𝑚𝑜𝑙 Salomons, 2008
𝑠
𝑐𝑚
𝐸 115 𝑘𝐽 Salomons, 2008
𝑚𝑜𝑙
116

5 APPENDIX C: List of Miscellaneous Equations

All the equations below are obtained from ‘Introduction to Catalytic Combustion’ by Hayes
and Kolaczkowski (1997)

Density of air
𝑃 × 𝑀𝑟
𝜌 =
1000𝑅 𝑇

Equation C.1
Specific mass heat capacity of air

1
𝐶 , = 28.09 + (0.1965 × 10 )𝑇 + (0.4799 × 10 )𝑇 − (1.965 × 10 )𝑇
0.02896
Equation C.2
Thermal conductivity of air

𝑘 , = 1.679 × 10 + (5.973 × 10 )𝑇

Equation C.3
Thermal diffusivity of air

𝑘 ,
𝛼=
𝜌 𝐶 ,

Equation C.4
Specific mass heat capacity of solid phase

𝐶 , = 948 + 0.2268𝑇

Equation C.5

Thermal conductivity of solid phase


117

𝑘 , = 0.9558 − (2.09 × 10 )𝑇

Equation C.6
Viscosity of air

𝜇 = 7.701 × 10 + (4.166 × 10 )𝑇 − (7.531 × 10 )𝑇

Equation C.7
Reynolds number

𝜌 𝑣 𝐷
𝑅𝑒 =
𝜇

Equation C.8
Prandtl number

𝐶 , 𝜇
𝑃𝑟 =
𝑘 ,

Equation C.9
Graetz number

𝐷
𝐺𝑧 = 𝑅𝑒 × 𝑃𝑟 ×
𝐿
Equation C.10
Nusselt number for constant wall temperature

.
48.2 1000
𝑁𝑢 = 3.657 + 8.827 × 𝑒𝑥𝑝 − ×
𝐺𝑧 𝐺𝑧
Equation C.11
Nusselt number for constant wall flux

.
60.2 1000
𝑁𝑢 = 4.364 + 13.18 × 𝑒𝑥𝑝 − ×
𝐺𝑧 𝐺𝑧
Equation C.12
118

Nusselt number

𝑁𝑢 + 𝑁𝑢
𝑁𝑢 = +1
2
Equation C.13
Heat transfer coefficient of air

𝑁𝑢 × 𝑘 ,
ℎ =
𝐷ℎ
Equation C.14
Diffusion coefficient of species A in air

1 1
+
𝑀𝑟 𝑀𝑟
𝐷 , , = (1.013 × 10 )𝑇
𝑃 𝑉 , + 𝑉 ,

Equation C.15
Convective mass transfer coefficient of species A in air

𝑆ℎ × 𝐷 ,
𝑘 , =
𝐷

Equation C.16
Knudsen diffusion coefficient of species A

𝑇
𝐷 = 97𝑟
𝑀𝑟

Equation C.17
Diffusion coefficient of species A in air in solid phase

1 1
𝑀𝑟 + 𝑀𝑟
𝐷 , , = (1.013 × 10 )𝑇
𝑃 𝑉 , + 𝑉 ,

Equation C.18
119

Diffusion in pores

1
𝐷 , =
1 1
𝐷 +𝐷
, ,

Equation C.19
Effective Diffusivity

𝜙𝛿𝐷 ,
𝐷 =
𝜏
Equation C.20
Heat of Reaction of C3H6

∆𝐻 = −2.059 × 10 + (72.3)𝑇 − (9.69 × 10 )𝑇 + (4.34 × 10 )𝑇 + (7.56 × 10 )𝑇

Equation C.21
120

APPENDIX C: Python Programming Code


121
122
123
124
125
126
127
128
129
130
131
132
133

Das könnte Ihnen auch gefallen