hydrogen production by the catalytic auto-thermal

209
HYDROGEN PRODUCTION BY THE CATALYTIC AUTO-THERMAL REFORMING OF SYNTHETIC CRUDE GLYCEROL IN A PACKED BED TUBULAR REACTOR A Thesis Submitted to the Faculty of Graduate Studies and Research In Partial Fulfillment of the Requirements For the Degree of Master of Applied Science In Process Systems Engineering University of Regina By Ahmad Mahmoud Abdul Ghani Regina, Saskatchewan July, 2014 Copyright 2014: A. M. Abdul Ghani

Upload: others

Post on 08-Jan-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

HYDROGEN PRODUCTION BY THE CATALYTIC AUTO-THERMAL REFORMING

OF SYNTHETIC CRUDE GLYCEROL IN A PACKED BED TUBULAR REACTOR

A Thesis

Submitted to the Faculty of Graduate Studies and Research

In Partial Fulfillment of the Requirements

For the Degree of

Master of Applied Science

In

Process Systems Engineering

University of Regina

By

Ahmad Mahmoud Abdul Ghani

Regina, Saskatchewan

July, 2014

Copyright 2014: A. M. Abdul Ghani

UNIVERSITY OF REGINA

FACULTY OF GRADUATE STUDIES AND RESEARCH

SUPERVISORY AND EXAMINING COMMITTEE

Ahmad Mahmoud Abdul Ghani, candidate for the degree of Master of Applied Science in Process Systems Engineering, has presented a thesis titled, Hydrogen Production by the Cataylitic Auto-Thermal Reforming of Synthetic Crude Glycerol in a Packed Bed Tubular Reactor, in an oral examination held on July 23, 2014. The following committee members have found the thesis acceptable in form and content, and that the candidate demonstrated satisfactory knowledge of the subject material. External Examiner: Dr. Shahid Azam, Environmental Systems Engineering

Co-Supervisor: Dr. Hussameldin Ibrahim, Process Systems Engingeering

Co-Supervisor: *Dr. Farshid Torabi, Petroleum Systems Engineering

Committee Member: Dr. Amr Henni, Industrial Systems Engineering

Committee Member: Dr. Raphael Idem, Process Systems Engineering

Chair of Defense: Dr. Christopher Street, Faculty of Business Administration *Not present at defense

i

ABSTRACT

The target of this work was to develop an efficient autothermal reforming (ATR)

process for the production of renewable hydrogen from synthetic crude glycerol (CG).

Hence, the work was divided into three phases: (1) development of a high performance

catalyst, (2) optimization of process variables, and (3) investigating the kinetics of the

involved reactions. A portfolio of ternary oxide catalysts with a nominal composition of

5Ni/CeZrM (where M= Ca, Gd, Mg) was prepared, characterized and tested in the

process. A series of experiments was conducted in a Packed Bed Tubular Reactor

(PBTR) using a factorial design technique to investigate the effects of the different

operating parameters. A rate model expression was then developed based on the

experimental kinetic data.

Synthetic CG was reformed over a modified cerium-zirconium support loaded

with nickel catalyst (5%Ni/CeZrM) by a combination of partial oxidation and steam

reforming reactions to generate hydrogen via an overall auto-thermal process. Amongst

the tested catalysts, calcium promoted one showed the highest catalytic activity due

mainly to its reducibility and nickel dispersion properties. The prepared catalysts were

characterized by N2 physisorption (BET), thermogravemetric analysis (TGA),

temperature programmed oxidation (TPO), temperature programmed reduction (TPR),

inductively coupled plasma-mass spectrometry (ICP-MS) and x-ray diffraction (XRD)

techniques.

Likewise, the composition of crude glycerol mixture generated at biodiesel plants,

free glycerol, methanol, soap, free fatty acids and ashes (NaCl and KCl), were contained

ii

in the synthetic crude glycerol. The catalytic performance was evaluated based on

conversion, hydrogen selectivity, hydrogen yield, turnover frequency and rate of coke

formation. A reforming temperature of 575°C, steam-to-carbon ratio (S/C) of 2.6,

oxygen-to-carbon ratio (O/C) of 0.125, reduction temperature of 600°C and calcination

temperature of 550°C were experimentally found to be the best operating conditions

based on hydrogen yield and process stability. Analysis of Variance (ANOVA) was

performed to study the main effects and interactions among the different parameters and

quantify the significance of each parameter on the overall performance. Reaction

temperature and S/C ratio were found to be the most influential variables on conversion

and H2 selectivity.

The kinetics of synthetic CG ATR reactions were studied in a temperature range

of 500-650°C, steam-to-carbon (S/C) range of 1.6-3.6, oxygen-to-carbon (O/C) range of

0.05-0.2, weight space velocity (WSV) range of 0-158.2 gcat.min/mol C and at

atmospheric pressure. In preparation for collecting intrinsic kinetic data, a region free of

heat and mass transfer limitations was established by varying catalyst particle size and

inlet flow rates in the ranges of 0.55-1.27 mm and 0.0019-0.0033 mol C/min,

respectively, and the stability of the catalyst was tested in an extended period run for 15

hours time on stream (TOS) experiment. The integral method of kinetic analysis was then

applied to estimate the parameters of the proposed power law model. The activation

energy was found to be 93.7 kJ/mol, alongside with reaction orders of 1, 0.5 and 2 with

respect to synthetic CG, steam and oxygen, respectively. Excellent agreement between

the experimental conversion results and those predicted by the model was observed with

an absolute average deviation (AAD) of 5.2%.

iii

ACKNOWLEDGMENTS

I would especially like to express my deep gratitude to my supervisor Dr.

Hussameldin Ibrahim for his guidance, support and encouragement throughout the course

of my research work that allowed me to complete this research work successfully. Also, I

would like to extend my gratitude to my co-supervisor Dr. Farshid Torabi for the

continuous support he has provided throughout my research work. I would also like to

thank Dr. Raphael Idem for his expert recommendations and helpful suggestions.

I would like to thank the Natural Sciences and Engineering Research Council of

Canada (NSERC) for their financial support through the NSERC Discovery Grant held

by Dr. Hussameldin Ibrahim. Thanks to Dr. Farshid Torabi, the Faculty of Graduate

Studies and Research (FGSR) and the Faculty of Engineering at the University of Regina

for providing additional financial support and the International Test Centre for CO2

Capture (ITC) for giving me the opportunity to perform the experimental phase of my

work in their laboratories.

I am very grateful to my family and friends for their motivation and support

throughout the course of this program. Special thanks go to my parents and to my brother

Mohammad for their assistance and suggestions. I also would like to thank my friends

Faezeh Sabri, Judith Oluku, Harry Inibehe, Naveed Qamar, Zeeshan Shahid, Anku Edet,

Bandar and Laila Jaafary for their encouragement and helpful discussions.

iv

TABLE OF CONTENTS

ABSTRACT i

ACKNOWLEDGMENTS iii

TABLE OF CONTENTS iv

LIST OF TABLES x

LIST OF FIGURES xi

NOMENCLATURE xvi

CHAPTER 1 INTRODUCTION 1

1.1 Hydrogen: The Global Perspective 2

1.2 Hydrogen Production Techniques 4

1.3 Justification 6

1.3.1 Biodiesel in Canada 6

1.3.2 Crude Glycerol: Biomass for Hydrogen 7

Production

1.4 Knowledge Gap and Problem Identification 10

1.5 Research Objectives and Scope 12

1.5.1 Phase 1: Catalyst Preparation and

Characterization 12

1.5.2 Phase 2: Experimental Investigation

of Process Performance in a PBTR 13

1.5.3 Phase 3: Kinetic Study 14

1.6 Original Contributions 15

v

CHAPTER 2 LITERATURE REVIEW 16

2.1 Precis 16

2.2 Hydrogen as a Fuel 16

2.3 Crude Glycerol 17

2.4 Hydrogen Production from Glycerol Biomass 19

2.4.1 Steam Reforming 20

2.4.2 Partial Oxidation 22

2.4.3 Auto-Thermal Reforming 23

2.4.4 Hydrogen Peroxide Reforming 30

2.4.5 Dry Reforming 30

2.4.6 Dry Auto-Thermal Reforming 31

2.4.7 Aqueous Reforming 31

2.5 Kinetics of Glycerol Reforming 32

2.6 Statistical Approaches 34

2.7 Catalysts for Glycerol Reforming 35

vi

CHAPTER 3 EXPERIMENTAL SECTION 37

3.1 Safety Precautions 37

3.2 Catalyst Preparation 38

3.2.1 Chemicals Used 38

3.2.2 Equipment Utilized 37

3.2.3 Support Preparation 39

3.2.4 Nickel Impregnation 41

3.3 Catalyst Characterization 42

3.3.1 Thermo-Gravimetric Analysis 42

3.3.2 N2 Physisorption 43

3.33 Temperature Programmed Oxidation 43

3.3.4 Temperature Programmed Reduction 44

3.3.5 Powder X-Ray Diffraction 45

3.3.6 Inductively Coupled Plasma Mass 45

Spectrometry

3.4 Reaction Feed Stock 46

3.4.1 Synthetic CG Feed 46

vii

3.4.2 Steam Feed 47

3.4.3 Oxygen Feed 47

3.5 Performance Evaluation 49

3.5.1 Experimental Setup 49

3.5.2 Catalytic Activity Evaluation 51

3.6 Performance Evaluation Criteria 54

3.7 Design of Experiments 59

CHAPTER 4 RESULTS AND DISCUSSION 61

4.1 Catalyst Characterization 61

4.1.1 Thermo-gravimetric Analysis 61

4.1.2 N2 Physisorption on Fresh Catalysts 64

4.1.3 N2 Physisorption on Spent Catalysts 67

4.1.4 Temperature Programmed Oxidation 70

4.1.5 Temperature Programmed Reduction 77

4.1.6 X-ray Diffraction 79

4.2 Catalytic Activity 81

4.2.1 Effects of Synthetic CG Components 81

viii

4.2.2 Catalysts Screening- Promoter Effect 87

4.2.3 Structure-Activity Relationship 91

4.3 Parametric Study 96

4.3.1 Effect of Reduction Temperature 96

4.3.2 Effect of Calcination Temperature 100

4.3.3 Effect of Steam-to-Carbon Ratio 103

4.3.4 Effect of Oxygen-to-Carbon Ratio 107

4.3.5 Effect of Reaction Temperature 110

4.4 Statistical Analysis 112

4.4.1 Main Effects and Interactions 112

4.4.2 Model Development 118

4.5 Kinetics 123

4.5.1 Heat and Mass Transport Limitations 123

4.5.1.1 Effect of Pellets Size 123

4.5.1.2 Effect of flow Rate 126

4.5.2 Extended TOS Stability 128

4.5.3 Mass Distribution 131

ix

4.5.4 Results of Kinetics Experiments 135

4.5.5 Rate Model 139

4.5.6 Estimation and Validation of Model 143

Parameters

Chapter 5 CONCLUSIONS AND RECOMMENDATIONS 145

5.1 Conclusions 148

5.2 Recommendations 151

REFERENCES 152

APPENDICES 161

Appendix A: Representative GC Data Sheets 164

Appendix B: Regression Results from NLREG for Conversion Model 184

Appendix C: Regression Results from PLKA software (Kinetics) 187

Appendix D: Carbon Analysis Results from Loring Laboratories LTD. 188

Appendix E: Synthetic CG Average Molecular formula Calculations 189

x

LIST OF TABLES

Table 1.0 Composition of crude glycerol from Milligan Biofuels Inc. 8

Table 2.1 Crude glycerol composition in wt% 19

Table 3.1 Compositions of catalysts prepared 41

Table 3.2 Physical properties and percentage composition of the synthetic CG 48

Table 3.3 Different reaction stoichiometries at different S/C and O/C ratios 56

Table 3.4 (a) Parametric experiments 59

Table 3.4 (b) Kinetic experiments (variables: reactant concentrations) 59

Table 3.4 (c) Kinetic experiments (variables: temperature and W/FA0) 59

Table 4.1 BET and ICP-MS results of fresh supports and catalysts 64

Table 4.2 BET results of spent catalysts at different reaction temperatures 67

Table 4.3 Pearson correlations coefficients for structure-activity relationship 92

Table 4.4a Experimental data for model development 119

Table 4.4b Experimental and predicted conversions 120

Table 4.5 Carbon distribution in the system (basis: 1 hour of operations) 131

Table 4.6 Reaction stoichiometries at different operating conditions 138

Table 4.7 Estimation results of model parameters 142

Table 4.8 Experimental kinetic data with the predicted conversion results 143

xi

LIST OF FIGURES

Figure 3.1 Schematic flow diagram of the experimental setup for synthetic CG 50

autothermal reforming in a PBTR

Figure 4.1 TGA profiles of CeZrCa support and the corresponding 5wt% catalyst 62

Figure 4.2 (a) N2 Isotherms of fresh CeZrM supports and corresponding 5Ni/CeZrM 65

catalysts

Figure 4.2 (b) N2 Isotherms of spent 5Ni/CeZrCa catalysts at different reaction 68

temperatures

Figure 4.3 Coke deposition rates at different O/C ratios 71

Figure 4.4 Coke deposition rates at different S/C ratios 71

Figure 4.5 Coke formation rates at different reaction temperatures 72

Figure 4.6 (a) TPO profiles of spent catalysts at different oxygen-to-carbon ratios 73

Figure 4.6 (b) TPO profiles of spent catalysts at different steam-to-carbon ratios 74

Figure 4.6 (c) TPO profiles of spent catalysts at different reaction temperatures 75

Figure 4.7 TPR profiles of the fresh supports and catalysts 77

Figure 4.8 XRD patterns of the fresh supports and catalysts 79

Figure 4.9 Evaluation of the individual effects of the different components of 83

CG on the ATR reactions of synthetic CG at 600°C, 1.0 atm, S/C= 3.6 and O/C= 0.125

using 5Ni/CeZrGd catalyst in a PBTR. (GM: Glycerol + Methanol)

xii

Figure 4.10 Individual effects of the different components of CG on the 85

performance of ATR reactions (a) Main effects plot for synthetic CG Conversion [mol

%] (b) Main effects plot for hydrogen selectivity [mol%] (c) Main effects plot for

hydrogen yield [mol H2/ mol C] (d) Main effects plot for turnover frequency [1/s].

Figure 4.11 Performance evaluations of 5Ni/CeZrM catalysts for synthetic 88

CG ATR at S/C= 3.6, O/C= 0.125 and reaction T=600°C in a PBTR.

Figure 4.12 Effects of the developed catalysts on the ATR of synthetic CG; 89

(a) Main effects plot for synthetic CG Conversion [mol%] (b) Main effects plot for

hydrogen selectivity [mol%] (c) Main effects plot for hydrogen yield [mol H2/ mol C] (d)

Main effects plot for Turnover Frequency [1/s].

Figure 4.13 Activity-structure correlations obtained for 5Ni/CeZrM catalysts 94

where M=Mg, Gd, Ca; (a) Main effects plot for synthetic CG conversion [mol%] (b)

Main effects plot for hydrogen selectivity [mol%] (c) Main effects plot for Turnover

frequency [1/s].

Figure 4.14 Effect of reduction temperatures on the activity of 5Ni/CeZrCa 97

catalyst at S/C= 3.6, O/C= 0.125 and reaction T=600°C in a PBTR.

Figure 4.15 Performance variations with reduction temperature 98

xiii

(a) Main effects plot for synthetic CG conversion [mol%] (b) Main effects plot for

hydrogen selectivity [mol%] (c) Main effects plot for hydrogen yield [mol H2/mol Cin)

(d) Main effects plot for turnover frequency [1/s].

Figure 4.16 Effect of calcination temperatures on the activity of 5Ni/CeZrCa 100

catalyst at S/C= 3.6, O/C= 0.125 and reaction T=600°C in a PBTR.

Figure 4.17 Performance variations with calcination temperatures 101

(a) Main effects plot for synthetic CG conversion [mol%] (b) Main effects plot for

hydrogen selectivity [mol%] (c) Main effects plot for hydrogen yield [mol H2/ mol Cin]

(d) Main effects plot for Turnover frequency [1/s].

Figure 4.18 Effect of S/C ratio on synthetic CG ATR reactions using 104

5Ni/CeZrCa catalyst at O/C= 0.05 and reaction T=600°C in a PBTR.

Figure 4.19 Performance variations with steam-to-carbon (S/C) ratio 105

at O/C= 0.05 and reaction T=600°C in a PBTR.

Figure 4.20 Effect of O/C ratio on synthetic CG ATR reactions using 107

5Ni/CeZrCa catalyst at S/C= 2.6 and reaction T=600°C in a PBTR.

Figure 4.21 Performance variations with oxygen-to-carbon (O/C) ratio 108

at 600°C and W/FA0=127.4 gccat.min/ mol C using 5Ni/CeZrCa catalyst.

Figure 4.22 Performance variations with reaction temperature at 110

xiv

S/C= 2.6, O/C= 0.125, W/FA0= 127.4 gcat.min/ mol C using 5Ni/CeZrCa catalyst.

Figure 4.23 Main effects plots 114

(a) Synthetic CG Conversion (b) Hydrogen Selectivity (c) Turnover Frequency.

Figure 4.24 Interactions plots 115

(a) Synthetic CG Conversion (b) Hydrogen Selectivity (c) Turnover Frequency.

Figure 4.25 Pareto Charts 116

(a) Synthetic CG Conversion (b) Hydrogen Selectivity (c) Turnover Frequency.

Figure 4.26 Parity plot of experimental versus predicted conversion 120

Figure 4.27 Effects of catalyst pellets size on the activity 123

(a) Synthetic CG Conversion (b) Hydrogen Selectivity (c) Hydrogen Yield

Figure 4.28 Effect of feed molar flow rate on the activity 125

(a) Synthetic CG Conversion (b) Hydrogen Selectivity (c) Hydrogen Yield

Figure 4.29 Long term stability run for synthetic CG ATR over 5Ni/CeZrCa 127

catalyst at S/C= 3.6; O/C= 0.125; reaction T= 600°C and W/FA0= 2.64 gcat. h/mol C

Figure 4.30 I-MR control charts of conversion and turnover frequency 128

Figure 4.31 Material flow diagram of the synthetic CG reformer 132

Figure 4.32 (a) Conversion variations with W/FA0 at different reaction 134

xv

temperatures (500 and 550°C)

Figure 4.32 (b) Conversion variations with W/FA0 at different reaction 135

temperatures (600 and 650°C)

Figure 4.32 (c) Conversion of synthetic CG versus W/FA0 136

Figure 4.33 Parity plot of predicted data versus experimental data 142

xvi

NOMENCLATURE

Å Angstrom [10-10

m]

AAD

Ag

Al2O3

ASTM

ATR

atm

BET

Average absolute deviation [%]

Silver

Aluminium oxide

American Society for Testing and Materials

Auto-thermal Reforming

Atmosphere

Brunauer, Emmett and Teller

C Carbon

CH4 Methane

CO2 Carbon dioxide

Co Cobalt

CO Carbon monoxide

CG Crude glycerol

oC

Cl

ΔH

Degrees Celsius

Chlorine

Heat of reaction [kJ/mol]

Ca

CZM

d

dp

e

Calcium

Promoted Ceria-zirconia-metal oxide

Internal diameter of the reactor [m]

Particle size of the catalyst [m]

Error [%]

xvii

EA

FA0

Activation energy [kJ/mol]

Inlet flow rate [mol C/min]

Fe Iron

GC

Gd

Gas chromatography

Gadolinium

H2

H2O

H2S

HP

ICP-MS

IUPAC

K

kJ

k0

kPa

L

Mg

mg

Na

Ni

N2

O

P

Hydrogen

Water

Hydrogen sulfide

High Pressure

Inductively Coupled Plasma-Mass Spectroscopy

International Union of Pure and Applied Chemistry

Potassium

Kilojoule

Frequency factor [mol C gcat-1

min-1

atm-3.46

]

Kilo Pascal

Catalyst bed height [m]

Magnesium

Milligram

Sodium

Nickel

Nitrogen

Oxygen

Pressure [Pa]

pH Negative logarithm of hydrogen ion concentration

xviii

PV

Pt

psi

Rh

S

SA

SAR

SSM

SiO2

S/M

S/C

Tmax

TCD

TGA

TOS

TPO

TPR

UHP

W/FA0

WGS

wt.

XRD

Specific Pore volume [cm3/g]

Platinum

Pound per square inch

Rhodium

Steam

Specific surface area [m2/g]

Structure activity relationship

Squared sum of residuals

Silica

Surfactant/metal ratio

Steam/Carbon

Temperature at maximum peak [°C]

Thermal conductivity detector

Thermo gravimetric analysis

Time on stream

Temperature programmed oxidation

Temperature programmed reduction

Ultra high purity

Weight time [gcat.min/mol C]

Water-gas-shift

Weight [g]

X-ray diffraction

Density of catalyst [g/cm3]

1

CHAPTER 1

INTRODUCTION

The growing science of climate change research is increasingly demonstrating the need to

produce more environmentally sustainable energy sources. The three confirmed facts

according to (Rand & Dell, 2008) are: (1) the climate on earth is changing to become

warmer; (2) the percentage of carbon dioxide in the atmosphere has seen a dramatic

increase since the beginning of the Industrial Revolution; (3) carbon dioxide released

from burning fossil fuels is a greenhouse gas that increases the heat capacity of air and,

thus, the energy reserved in the atmosphere. Not only carbon dioxide, but also methane,

nitrous oxides, hydrofluorocarbons (HFCs), perfluorocarbons (PFCs) and sulphur

hexafluoride have been classified as greenhouse gases that contribute to the global

warming. The Intergovernmental Panel on Climate Change (IPCC, 2014) reported net

GHG emissions of more than 49 billion tonnes of CO2 equivalents in 2010, which was

the maximum level in human history. For these reasons and more, it is now widely

agreed that hydrogen energy is a promising alternative capable to replace carbon-based

one. A comprehensive strategic plan is required to account for the associated

technological, economical, political and social considerations.

2

1.1 HYDROGEN:THE GLOBAL PERSPECTIVE

In 1997, the U.S. President’s Council of Advisors on Science and Technology

expected that hydrogen will become a significant energy carrier compared to electricity

not in the short term, but at least in the mid- to long term research work plans. In May

2009, Steven Chu, the Secretary of Energy said “We asked ourselves, is it likely in the

next 10, 15 or 20 years that we will convert to hydrogen car economy? The answer, we

felt, was no”. McKinsey’s study in 2010 showed that Fuel Cell Electric Vehicles

(FCEVs) are technologically ready and can be commercialised to the market in the next

five years; the only step left was to develop an inclusive market plan in order to build

hydrogen infrastructure in Europe.

Hydrogen has long been recognized to have the highest amount of energy released

per unit mass burned, with -141 kJ/g enthalpy of combustion. Due to this fact, significant

efforts have been spent towards increasing the efficiency of hydrogen fuel (Mills et al.,

2002). Based on fuel-to-wheel efficiency values reported by (Sørensen, 2012), the

efficiency of hydrogen operated fuel cell cars is about 36%, while it is 26% for the

electric vehicles, 27% for high efficiency diesel engines and 17% for the conventional

gasoline engines.

As a clean, abundant and renewable source of energy, hydrogen can be utilised in

different ways to become a significant source of energy in the near future. According to

Hoffman (2012), the international world standards are getting to be zero-emissions from

automobiles, industry, residential furnaces and general transportations. Technically,

taking the carbon out of fossil fuel emissions is hard especially when considering mobile

3

or small emitters. The fossil fuels upon combustion release CO2, CO, NOx and SOx to the

atmosphere. These resulting pollutants contribute directly to serious environmental

problems such as global warming and climate change. According to McMichael et al.

(2003), fossil fuel emissions degrade air quality and continually change world’s climate;

and this in turn will impose several negative effects on human health and ecosystem.

Unlike fossil fuels, the hydrogen has minimum environmental impacts due to its zero

harmful emissions upon combustion. Moreover, considering the availability and

limitations of other alternative renewable energy sources; hydrogen has been always

preferred as no one can guarantee the wind to blow or the sun to shine.

4

1.2 HYDROGEN PRODUCTION TECHNIQUES

Hydrogen can be produced via various processes either from conventional fossil fuel

sources such as natural gas and coal, or alternative sources such as biomass, wind,

nuclear, solar and others. Typical reforming techniques for hydrogen production from

glycerol are: steam reforming, partial oxidation, auto-thermal reforming, hydrogen

peroxide reforming, dry reforming, dry autothermal reforming and aqueous phase

reforming. These techniques are discussed in more details in the literature review section.

By combining the functions of both steam reforming (endothermic) and partial oxidation

(exothermic) in one reactor, an auto-thermal process for glycerol reforming can be

developed to enhance hydrogen production and the overall efficiency. The main

advantageous factors in this process can be summarized as:

I. Energy efficiency: energy required for steam reforming will be supplied by the

highly exothermic oxidation reactions.

II. Process stability: oxidation of carbon molecules hinder coke formation over the

surface of the catalyst and subsequently maintain its activity (Hardiman et al.,

2006).

III. Practicality: auto-thermal reforming is a more pragmatic technique in real

biomass reformers, plus it can yield higher hydrogen productivity than steam

reforming (Ahmed and Krumpelt, 2001).

It is now very promising that renewable hydrogen from biomass reforming will play

a major role in the near future to compensate the dramatic increase of energy demands,

concurrent with the tendency of establishing an environment friendly energy system. The

5

cost of producing these renewable energy carriers as feasible alternatives for the fossil

fuels is still the main drawback. Crude glycerol mixture, the main biodiesel by-product

from transesterification process, worsens the cost effectiveness of biodiesel production.

The current market is flooded with glycerol and any further refining/purification is not

economical. On the other hand, glycerol mixture can be a significant biomass to be

utilised for hydrogen production for three main reasons: (1) to reform a toxic waste; (2)

to produce renewable hydrogen; (3) and to enhance biodiesel production economy.

6

1.3 JUSTIFICATION

1.3.1 BIODIESEL IN CANADA

Biodiesel utilization for supplying energy demands is a potential way to reduce

GHG emissions. As compared to the fossil diesel, the biodiesel emits 41% less GHG

upon combustion (Hill et al., 2006). A study by Dyer et al. (2010) revealed a significant

decrease in GHG emissions after blending petro-diesel with 2% and 5% biodiesel. The

net reduction in GHG emissions was measured to be 2.6 MgCO2eha-1

and 0.94

MgCO2eha-1

in the eastern and western regions of Canada; respectively.

Biodiesel manufacturing process is strongly influenced by the applied operating

conditions, raw materials and catalyst used. Nowadays, the most common technique to

produce biodiesel is the tranesterification of vegetable and animal oils (Chisti, 2007).

Using different feedstocks in the production process, the properties of biodiesel will be

mainly affected by the content of FFA in the fatty material used. The efficiency of using

waste cooking oil to produce biodiesel in Canada was studied by [Kulkarni & Dalai

(2006); Chhetri et al. (2008)]. Based on their results, waste cooking oil is an economical

feedstock to be utilised, and it yields a high performance fuel for diesel engines compared

with biodiesel generated from other feedstocks. Additionally, biodiesel production from

waste cooking oil reduces environmental problems and helps to avoid ethical debates

over using agricultural crops for the purpose.

Canadian governments have always motivated the production of biofuels through

low-tax, tax-refund and non-taxable biofuel programs. Besides that, in Saskatchewan,

Manitoba and Ontario, petro-fuels must contain 5 wt% biofuels before selling to the

7

customer. A study by (Smith et al. 2007) showed that the overall energy balance in

Canada can also be enhanced by producing biodiesel from canola and soybean oils. The

energy obtained from biodiesel was reported to be 2.1 to 2.4 times more than the energy

required for producing vegetable oils.

1.3.2 CRUDE GLYCEROL: BIOMASS FOR HYDROGEN PRODUCTION

The by-product of the transesterification process is the glycerol. The process

yields 1 mole of glycerol for every three moles of biodiesel produced. At present, the

worldwide production of crude glycerol is constantly increasing due to the continuous

developments in biodiesel economy. According to 2007 market analysis by ABG Inc.,

the estimated production of crude glycerol from biodiesel industry is expected to exceed

five billion pounds by 2020, which is really a huge volume. At the same time, demand on

free glycerol is not growing due to the saturation of the market, and this affects the value

of crude glycerol and consequently the efficiency of biodiesel production. According to

Yazdani and Gonzalez (2007), a sharp diminution in crude glycerol value has been

observed between 2004 and 2006 when its price dropped from 25 cents/lb to less than 3

cents/lb. During the same period, the production of biodiesel increased from 25 million

gallons in 2004 to about 250 million gallons in 2006. Subsequently crude glycerol

became a waste product with the associated disposal cost. As recently reported by Ayoub

& Abdullah (2012), the existing market of crude glycerol is very unsteady. Zhang et al.

(2003) reported a production rate of 120,000 ton/year of yellow grease, cooking oil with

less than 15 wt% FFA, in Canada. Based on the future scenarios for crude glycerol

treatment, supply and demand relationship will automatically control the prices of this

product.

8

Unlike the purified one, the high amount of impurities in crude glycerol makes it

unfavourable for commercial purposes. These impurities are usually methanol, ash, soap,

water, FFA and FAME residues. Crude glycerol pertinent to Saskatchewan biodiesel

production has been collected from Milligan Biofuels Inc., the first Supplier of biodiesel

in the province. Excluding the high percentage of potassium hydroxide present in

Milligan’s crude glycerol, the composition given in Table 1.0 was very similar to the

composition reported by Hu et al. (2012) in the United States.

Table 1.0 Composition of Crude Glycerol From Milligan Biofuels Inc. (MSDS)

Ingredients Weight %

Potassium Hydroxide 10.00-30.00

Methanol 1.00-5.00

Glycerine 40.00-70.00

Fatty acids, canola, methyl esters 7.00-13.00

Organic material and soap 10.00-30.00

Crude glycerol generated from biodiesel operations should always be handled with

caution. According to material safety data sheet (MSDS) of stripped glycerin by Milligan

(2013), crude glycerine is a controlled product that contains toxic, flammable and

corrosive materials. A short-term exposure to humans could result in skin damage,

dizziness, serious problems in the nervous and respiratory systems, blindness and death.

For a safe storage, crude glycerine should be stored in a proper container at a temperature

below 38°C, and in a well ventilated area away from metals, acids and sunlight. In

9

addition, shipping of crude glycerol in Canada must be in accordance with the transport

of dangerous goods regulations TDGR.

Since the refining of crude glycerol to obtain pure products is very costly and cannot

be economically justified in the current market, CG generated at Canadian biodiesel

plants are mostly shipped to the United States for further processing (Milligan biofuels).

Many researchers these days are working hard to develop practical techniques for crude

glycerol reforming to improve the cost effectiveness of biodiesel production as an

efficient renewable energy for a sustainable future.

10

1.4 KNOWLEDGE GAP AND PROBLEM IDENTIFICATION

The major techniques for hydrogen production from glycerol reforming are: steam

reforming, partial oxidation and the combination of both reactions to have an auto-

thermal reforming process. Each technique has its own advantages and drawbacks related

to energy efficiency, hydrogen yield and process stability. The main reasons of interest in

autothermal reforming process are: (1) no external heat is required and consequently

simpler and lower capital cost technology (Rand & Dell, 2008); (2) stable operations with

the least coke formation (Hardiman et al., 2006); (3) practical technique that gives high

hydrogen production yields (Ahmed and Krumpelt, 2001). Thermodynamics of pure

glycerol autothermal reforming was investigated by [Wang et al. 2009; Yang et al. 2011].

Some thermodynamic studies addressed crude glycerol composed only of free glycerol

and methanol [Authayanun et al. 2010; Ortiz et al. 2011]. Thermodynamics of both pure

and crude glycerol feeds emphasized on the feasibility of autothermal reforming reactions

to produce hydrogen.

On the catalyst side, the failure of mechanical stability of Rhodium catalysts plus the

formation of several non-equilibrium components with Platinum catalysts were observed

by Rennard et al. (2009) upon testing the catalytic activity of these materials in the

process of autothermal reforming of glycerol. Unlike the results from yellow glycerol

autothermal reforming; nickel supported on promoted alumina carrier showed poor

performance and low hydrogen yield upon using crude glycerol from biodiesel plants due

to high coke formation and subsequently catalyst deactivation (Kamonsuangkasem et al.

2011). Most of the catalysts utilised for the purpose of glycerol reforming were Nobel

metals and very expensive [Pompeo et al., 2010; Dauenhauer et al., 2006; Rennard et al.,

11

2009; Slinn et al., 2008; Kunkes et al., 2008; Zhang et al., 2007; Hirai et al., 2005]. A

serious gap in the experimental work and reaction kinetics related to the process of

glycerol autothermal reforming has been discovered. Strong efforts were made in our

group by Sabri (2013) to experimentally optimise the different parameters in the process.

Pure glycerol was the feed reformed in this study. Nickel supported over cerium-

zirconium promoted carrier showed high catalytic performance in the process. At the end,

the kinetics of pure glycerol autothermal reforming was investigated and a rate model

equation was developed.

12

1.5 RESEARCH OBJECTIVES AND SCOPE

The aim of this study was to generate hydrogen via an efficient process capable of

reforming synthetic crude glycerol similar to the one generated at biodiesel plants. The

main objectives of the work were (1) to develop a proper catalyst capable of handling the

different impurities of crude glycerol, (2) to optimize different operating variables in the

autothermal reforming process in order to attain a stable hydrogen yield, (3) to investigate

the main effects and interactions between these parameters, and (4) to study the kinetics

of autothermal reforming reactions and develop an appropriate rate model expression.

The performance of different catalysts under similar operating conditions was

evaluated based on glycerol conversion, hydrogen selectivity, turnover frequency and

coke formation criteria. The stability of the process was studied based on an extended

term run for 16 hours time-on-stream. The screening experiments at different operating

conditions were conducted in a PBTR for 6 hours time-on-stream. Both the fresh and the

spent catalysts were characterized for their physical and chemical properties. A statistical

analysis was performed to investigate the effects and interactions among the different

operating parameters and to generate a general model for glycerol conversion. The

objectives of this work were mapped in into three phases.

1.5.1 Phase I: Catalyst Preparation and Characterization

i. Preparation of cerium-zirconium supports promoted with gadolinium, magnesium

and calcium according to surfactant-assisted technique.

ii. Impregnation of 5 wt% Nickel over the prepared supports according to the

standard wet impregnation technique.

13

iii. Characterizing the prepared supports and catalysts for their physical and chemical

properties using the following techniques:

a. Thermogravimetry Analysis (TGA)

b. Temperature Programmed Reduction (TPR)

c. Temprature Programmed Oxidtion (TPO)

d. N2 Physisorption (BET)

e. X-ray Diffraction (XRD)

f. Inductively Coupled Plasma Mass Spectrometry (ICP-MS)

1.5.2 Phase Π: Experimental Investigation of Process Performance in a PBTR

i. Conducting the performance screening experiments in a PBTR over Ni-based

catalysts to discover the effects of different operating variables on the reforming

process. The optimized operating parameters in this study were:

a. Promoter element (Ca, Mg & Gd)

b. Reduction temperature (500, 600 & 700°C)

c. Calcination temperature (550, 600 & 650°C)

d. Reaction temperature (500, 550,600 & 650°C)

e. Steam-to-carbon molar ratio (1.6, 2.6 & 3.6)

f. Oxygen-to-carbon molar ratio (0.05, 0.125 & 0.2)

ii. Statistically analyzing the collected experimental data to better correlate the

effects of the different operating parameters to the overall performance. These

effects were also quantified and regressed to generate a mathematical model for

crude glycerol conversion.

14

iii. Characterizing the spent catalysts to investigate the impacts of the major

parameters on the catalyst side. This was done by applying BET and TPO analysis

on the spent catalysts.

iv. Analyzing the carbon contained in the different streams of the process and

performing a material balance to validate the experimental runs. The carbon in the

liquid stream was quantified by a TOC analysis. The results are given in

Appendix D and interpreted in the results section.

1.5.3 Phase Ш: Kinetic Study

i. Establishing the kinetic regions to avoid any possible heat or mass transfer

limitations. This was achieved by varying catalyst particle size and inlet flow of

crude glycerol.

ii. Testing the stability of the process through an extended term run for 16 hours

time-on-stream experiment.

iii. Collecting the intrinsic kinetic data by varying reforming temperature, catalyst

weight time (W/FA0), steam-to-carbon and oxygen-to-carbon ratios as the main

four kinetically controlled parameters in the process.

iv. Development and validation of a rate expression (power law model) to

mathematically describe the kinetics of ATR of crude glycerol.

The next chapter in this study will present a thorough literature review about the

hydrogen production from glycerol reforming. The addressed topics include hydrogen

energy, crude glycerol from biodiesel plants, techniques of hydrogen production,

catalysts used, and kinetics and statistical approaches in the field.

15

1.6 ORIGINAL CONTRIBUTIONS

Crude glycerol generated at biodiesel plants has a great potential for hydrogen

production. The high concentration of free glycerol and methanol in the crude mixture

makes it a desirable source for renewable hydrogen. The overall efficiency of hydrogen

production from glycerol is directly related to the reforming process. This includes the

employed catalyst, operating conditions, reactants concentrations, etc. Most of the

previous studies were focused on producing hydrogen from pure glycerol via

conventional steam reforming or partial oxidation techniques. However, these techniques

have basic limitations related to energy requirement, process stability and hydrogen yield.

In addition, the studied feedstock was mostly pure glycerol. The several impurities

present in crude glycerol were neglected due to their negative effects on catalyst activity

and process stability. Refining crude glycerol to obtain pure products is very expensive.

The low prices of pure glycerol and methanol make it unfeasible to go for further

refining.

The current study deals with synthetic crude glycerol having impurities similar to

the ones contained in the industrial mixture. The autothermal reforming technique was

employed due the basic advantages mentioned before. A proper stable catalyst was then

developed. The effects of the different operating variables were investigated through an

inclusive experimental work. The main interactions among these variables and their

magnitudes were statistically analyzed. The kinetics of the involved reactions was also

studied and a representative rate model was developed. The employed realistic situations

of feedstock besides the comprehensive experimental and mathematical analyses in this

study are the key factors to implement this process at the industrial scale.

16

CHAPTER 2

LITERATURE REVIEW

2.1 PRECIS

The main contributions of the current research work to the pool of scientific

knowledge can be identified as: (1) developing an inexpensive catalyst for hydrogen

production via auto-thermal reforming of synthetic crude glycerol; (2) reforming crude

glycerol without pre-treatment and further expenses; (3) characterising and screening

catalyst performance under different operating conditions; (4) simplifying the employed

reactions and reactor design; (5) investigating the main effects and interactions in order to

attain a stable hydrogen yield; and (6) conducting a kinetic study for the ATR of

synthetic crude glycerol. This chapter includes the extensive literature survey that was

conducted to fully track the progress in the area of hydrogen production from glycerol

reforming.

2.2 HYDROGEN AS A FUEL

In 1766, British philosopher Henry Cavendish discovered the “inflammable air”

which produces water upon combustion. “Inflammable air” was the common name of

Hydrogen before 1783 until it was recognised and named by the French chemist Antoine

Lavoisier, Father of Modern Chemistry. After he burned hydrogen in the presence of

oxygen, Lavoisier proved that water is not a pure element as what had been known

before, but a combination of hydrogen and oxygen atoms. In 1787, Lavoisier with

Guyton de Morveau, Claude-Louis Berthollet, and Antoine François de Fourcroy

developed a new approach to nomenclature and submitted their reports to the Academy

17

of Sciences. Since that time, oxygen and hydrogen gases have been officially identified

and considered as two distinct compounds. Hydrogen gas is composed of the most

abundant element in the universe. It is a diatomic non-metal compound formed of two

hydrogen atoms. As a colourless, odourless, non-toxic, very light, carbon free and highly

combustible gas, hydrogen is a clean energy carrier that yields 2.75 times more energy

than hydrocarbon fuels (Kapdan and Kargi, 2006). Several techniques have been applied

to produce hydrogen from both fossil and renewable resources. According to Holladay et

al. (2009), hydrogen production technologies are divided into two major groups: fuel

processing and non-reforming techniques. Hydrogen from fuel processing includes

hydrocarbon reforming, pyrolysis, plasma reforming, aqueous reforming and ammonia

reforming. On the other side, non-reforming techniques go under two main branches:

Hydrogen from biomass such as; biomass gasification and biological hydrogen

Hydrogen from water such as; electrolysis, thermochemical water splitting and

photo-electrolysis.

2.3 CRUDE GLYCEROL

The increasing tendency toward replacing fossil fuels with cleaner energy carriers

keeps pushing the need of alternative energy sources to the top. Biodiesel is a promising

alternative produced mainly via transesterification of vegetables or animal fats with

alcohol in the presence of a basic catalyst, usually NaOH or KOH (Ma and Hanna, 1999).

The weight ratio of crude glycerol to fatty acid methyl esters ‘FAME’ produced from

transesterification reactions at biodiesel plants is about 9:1 (Johnson and Taconi, 2007). The

composition of crude glycerol strongly depends on the oil source, alcohol feed, catalyst

18

used, reactions utilised, process efficiency, operating conditions and any pre or post

treatments. In 2006, Thomson and He characterized several crude glycerol samples

generated from different vegetable oil feedstocks: canola, crambe, mustard, rapeseed,

soybean and waste cooking oils; the last showed a big difference with respect to crude

glycerol chemical and physical properties compared to the rest of samples that showed

little variations from each other. According to Wijesekara et al. (2008), the chemical

composition (w/w) of crude glycerol from sunflower oil looks like: 50% methanol, 30%

free glycerol, 13% soap, 2% moisture, 2-3% salts (mostly sodium and potassium) and

same 2-3% for other impurities. Furthermore, Hansen et al. (2009) studied the

characteristics of eleven crude glycerol samples generated at seven different biodiesel

plants in Australia. As a part of their results, the researchers gave information about pH,

density and composition of raw glycerol. After calculating the average of 11

measurements, raw glycerol was found to have a value of 5.9 for pH, 1.2 g/cm3 for

density, 72.4% (w/w) for glycerol content, 5.5% for moisture, 5.2% for ash, 4.12% for

methanol and 13.9% for the matter organic non-glycerol (MONG). In a follow up, the

most recent and comprehensive work by Hu et al. (2012) presented a detailed study

corresponding to the physical and chemical characteristics of five raw glycerol samples

obtained from biodiesel manufacturing processes. The determined physical properties

were: density, pH and viscosity. Besides the results of physical properties and elemental

composition, Hu et al. (2012) reported the chemical composition of crude glycerol as

weight composition (Table 2.1) of free glycerol, methanol, free fatty acids (FFAs), fatty

acids methyl esters (FAMEs), soap, glycerides, water and ash.

19

Table 2.1 Crude Glycerol Composition in wt%

CG Samples A B C D E

Free Glycerol 63.0 ± 0.3 22.9 ± 0.2 33.3 ± 0.1 27.8 ± 0.2 57.1 ± 0.0

Methanol 6.2 ± 0.0 10.9 ± 0.2 12.6 ± 0.1 8.6 ± 0.0 11.3 ± 0.0

Water 28.7 ± 0.3 18.2 ± 0.1 6.5 ± 0.1 4.1 ± 0.1 1.0 ± 0.1

Soap BDL 26.2 ± 0.2 26.1 ± 0.1 20.5 ± 0.1 31.4 ± 0.1

FAMEs BDL 21.3 ± 0.2 19.3 ± 0.3 28.8 ± 1.1 0.5 ± 0.1

Glycerides BDL 1.2 ± 0.2 1.6 ± 0.3 7.0 ± 0.5 0.4 ± 0.1

FFAs BDL 1.0 ± 0.1 1.4 ± 0.1 3.0 ± 0.1 BDL

Ash 2.7 ± 0.1 3.0 ± 0.0 2.8 ± 0.1 2.7 ± 0.0 5.7 ± 0.2

All numerical values given in the table are in weight percentages

Data are expressed as mean of three replicate measurements ± standard deviation. BDL: Below the

detection limit

A, B, C: Soybean oil based. D: Waste vegetable oil based. E: soybean oil−waste vegetable oil mixture.

2.4 HYDROGEN PRODUCTION FROM GLYCEROL BIOMASS

Most of hydrogen on our planet is found as atoms participating in the formation of

different molecules such as water, fossil fuels, organic and inorganic materials. The

process of hydrogen production includes both extracting and isolating hydrogen

molecules to a desired level of purity (Sørensen, 2012). As mentioned earlier, hydrogen

can be produced through different processes either from conventional fossil fuel sources

such as natural gas and Coal, or alternative sources such as biomass, wind, nuclear, solar

and others. An economical analysis by Bartels et al. (2010) showed that the hydrogen

production from natural gas and coal is currently the most economical. However, this will

not be the case in the near future with the expected technological progress in utilising

20

new alternative sources. Due to this fact and the thorough contributions to the sustainable

world energy plan, renewable hydrogen from biomass is becoming a portentous fuel to

supply future energy demands. Considering the large agricultural countries, such as

United States and Germany, hydrogen generation from biomass comes in series with

agricultural activities in order to supply energy demands from one side, and to reform

organic wastes and undesired by-products from the other; and thus enhancing the

efficiency of clean renewable energy by completing the loop of sustainability from both

economical and environmental sides (Kotay and Das, 2007).

2.4.1 STEAM REFORMING

Steam reforming of glycerol is a highly endothermic reaction that produces 7

moles of H2 per 1 mole of glycerol decomposed according to the stiochiometric ratio.

Over certain catalysts, glycerol can react with steam to produce hydrogen and carbon

dioxide according to reaction 2.1.

C3H8O3 (g) + 3 H2O → 7H2 (g) + 3 CO2 (g) Δ = +123 KJ/mol (Adhikari et. al,

2007) (2.1)

Using a commercial Nickel-based catalyst, Czernik et al. (2002) was able to

reform crude glycerol and generate hydrogen via steam reforming technique. Even

though the process had not been optimized by that time, their results were very promising

with 77% hydrogen production yield.

Adhikari et al. (2006) conducted a thermodynamic analysis on steam reforming of

glycerol and found that operating conditions such as atmospheric pressure, temperature >

900 K and steam to glycerol ratio of 9:1 are the optimum in terms of hydrogen production

21

and coke and methane formation. In line with that, Chen et al. (2012) presented an

extensive work that included results of both thermodynamic analysis and pilot-scale

experiments. A fair agreement between thermodynamic results and experimental data

was found for different operating parameters such as: temperature, pressure, Steam-to-

glycerol, inert gas-to-glycerol and residence time. The study reported an optimum water-

to-glycerol ratio of 9.0 for hydrogen production at high temperature and low pressure.

Several catalysts such as; Ni, Co, Ir, Ru, Pt and Rh have been tested for their

capabilities of promoting the process of hydrogen production via glycerol steam

reforming. Slinn et al. (2008) studied the ability of using Pt-based catalyst supported on

Al2O3 for glycerol steam reforming. The optimum reforming temperature was found to be

880°C at 0.12 mol glycerol/ min per kg catalyst and with steam/carbon ratio of 2.5. In a

follow up, Pompeo et al. (2010) tested the performance of platinum catalyst impregnated

over ZrO2, SiO2, γ-Al2O3 and α-Al2O3 supports and proved that the acidity of support

increases the rate of coke formation and thus catalyst deactivation. A comparative study

by Zhang et al. (2007) showed a superior performance for the Ir/CeO2 catalyst compared

to Cobalt and Nickel ones. The study reported a complete glycerol conversion with

hydrogen selectivity of more than 85% at a relatively low temperature of 400°C. The

performance of Nickel catalyst supported on MgO, TiO2 and CeO2 was investigated by

Adhikari et al. (2007). Ni/CeO2 mixture was found to give the best results with a

hydrogen selectivity of 74.7% and glycerol conversion up to 99% at 600°C and steam to

glycerol ratio of 12:1.

According to Buffoni et al. (2009), Ni-based catalysts are very active and

selective in the process of glycerol steam reforming; the study suggested a minimum

22

reforming temperature of 550°C and a Ceα support based on hydrogen selectivity and

catalyst stability performance. In addition, the catalytic activity of Nickel supported on

CeO2, Al2O3 and CeO2-promoted Al2O3 was investigated by Iriondo et al. (2010). This

work reported a poor performance for pure ceria support compared to the conventional

and modified alumina ones that showed high conversion and stability results, possibly

due to the low amount of nickel deposits on pure ceria.

Using Ruthenium catalysts, steam reforming of crude glycerol was studied by

Hirai et al. (2005); based on their results, steam reforming of glycerol at 600°C is optimal

with 3 wt% of ruthenium impregnated over Y2O3. Again, magnesia showed poor glycerol

conversion and hydrogen selectivity results compared with yttria and zirconia supported

catalysts.

2.4.2 PARTIAL OXIDATION

Oxidation chemically means the reaction of any material with oxygen; the partial

term accounts for the smaller than stoichiometric amount of oxygen fed to the reformer

compared with the full oxidation. Partial oxidation and auto-thermal reforming are more

practical techniques for real reformers compared with steam reforming, due to energy

efficiency and processor simplicity considerations (Ahmed and Krumpelt, 2001). A

review on hydrogen production techniques from glycerol by Adhikari et al. (2009)

suggested further research work on crude glycerol partial oxidation and at supercritical

steam reforming conditions. Catalyst deactivation was pointed out to be the main

challenge in the process of crude glycerol reforming, because of the considerable amount

of impurities in the crude waste generated at biodiesel plants.

23

In order to identify partial oxidation as a practical technique for hydrogen

production from glycerol reforming, the first need was to investigate the effects of

different process parameters, establish reaction regions and then see how efficient the

process could be under different conditions. In line with that, a thermodynamic analysis

on glycerol partial oxidation was conducted by Wenju (2010) to determine the effects of

different operating variables on process performance and establish regions for coke

formation and optimum conditions. Based on Gibbs free energy minimization method,

the effects of temperature, pressure, oxygen to glycerol and nitrogen to glycerol ratios on

glycerol partial oxidation reactions were assessed. The optimum conditions for hydrogen

production were found to be: 727-827°C temperature, 1.0 atmosphere pressure, and 0.4-

0.6 O2/C3H8O3 molar ratio. Besides that, low temperatures and low O2/C3H8O3 ratios

were favourable for carbon formation and thus catalyst deactivation; and even though the

influence of inert gases on hydrogen yield was positive, the author recommended

avoiding it in practical applications. In addition, Wenju established the regions of

temperature, pressure and O2/C3H8O3 molar ratio where partial oxidation reactions of

glycerol can thermodynamically occur.

2.4.3 AUTO-THERMAL REFORMING

Swami and Abraham (2006) examined the performance of auto-thermal reforming of

glycerol over Nickel, Palladium, Copper and Potassium catalysts in temperature range of

550-850°C. The results obtained from ATR process revealed higher hydrogen production

rates and lower coke formation over the surface of catalyst than that of steam reforming.

In line with this work, auto-thermal reforming of glycerol over Rhodium catalysts was

studied by Dauenhauer et al. (2006); a high reforming temperature of 862°C; S/C ratio of

24

4.5 and C/O ratio of 0.9 were reported as the optimum operating conditions to completely

convert glycerol and reach 79% hydrogen selectivity. In addition, their study

demonstrated that undesirable products formation such as CO and CH4 has been

suppressed to a selectivity of less than 2% under optimum conditions.

In a different case, Luo et al. (2007) studied the thermodynamics of glycerol auto-

thermal reforming as a combination of oxidation and aqueous steam reforming in a

temperature range of 300-500 K. Methanation reactions were promoted in the tested

regions leading to a dramatic drop in the yield of hydrogen production. In addition, CO

formation was mostly controlled by reaction temperature, while H2 production was

mostly affected by oxygen to glycerol ratio with a gas product mainly composed of CO2

and CH4 under the tested conditions.

Vaidya and Rodrigues (2009) presented an extensive review on glycerol reforming

and the techniques employed for this purpose. A particular attention was given to CeO2

and MgO supports for their stability in practical reforming tasks compared with Al2O3

that showed quick deactivation due to dehydration. Further investigation on the effects of

promoter elements, particle size and metallic loadings was recommended, and a serious

gap in literature related to the kinetics of glycerol reforming has been reported.

In consonance with these recommendations, Kamonsuangkasem et al. (2011) tested

the performance of Ni/Al2O3 promoted with CeO2-ZrO2 catalyst in the oxidative steam

reforming process using yellow and crude glycerol feedstock at 923 K, water to glycerol

ratio of 9.0 and oxygen to glycerol ratio of 0.5. In spite of the promising results obtained

from yellow glycerol, crude glycerol from biodiesel plants was reported as a poor

25

feedstock for hydrogen production under these conditions due to catalyst deactivation;

applying a set of proper operating conditions or pre-treatment steps were recommended

for further investigation. The enhanced performance of AL2O3 promoted carrier

compared to pure one has been linked to both, the oxidative reforming reactions and

CeO2-ZrO2 promoter that probably reduced coke precipitation.

Rennard et al. (2009) studied the efficiency of glycerol reforming via autothermal

steam technique using Rh-cerium and Platinum catalysts. According to the results of Rh-

Ce catalyst, hydrogen yield can be increased by feeding more steam to the system up to

steam/carbon value of 2/3, the ratio that amends water gas shift reaction within the

reformer without suppressing the main reforming reaction. The study reported a

significant loss in mechanical stability for Rhodium catalysts due to obscure reasons. In

addition, the production of different non-equilibrium components such as acrolein,

hydroxyacetone and acetaldehyde was observed by using platinum catalysts.

According to [Authayanun et al. 2010; Wang et al. 2009; Vaidya & Rodrigues, 2009], the

main reactions expected to take place inside the autothermal reformer of crude glycerol

are:

Partial oxidation of glycerol

C3H8O3 + 2O2 ⇄ 3CO + 4H2O (2.2)

Partial oxidation of methanol

CH3OH + O2 ⇄ CO + 2H2O (2.3)

Steam reforming of glycerol

26

C3H8O3 + 3H2O ⇄ 3CO2 + 7H2 (2.4)

Steam reforming of methanol

CH3OH + H2O ⇄ CO2 + 3H2 (2.5)

Water gas shift reaction

CO + H2O ⇄ CO2 + H2 (2.6)

Dry reforming of methane

CH4 + CO2 ⇄ 2H2 + 2CO (2.7)

Steam reforming of methane

CH4 + H2O ⇄ 3H2 + CO (2.8)

Decomposition of methanol

CH3OH ⇄ CO + 2H2 (2.9)

Methanation Reactions

CO + 3H2 ⇄ CH4 + H2O (2.10)

CO2 + 3H2 ⇄ CH4 + 2H2O (2.11)

Based on Gibbs free energy minimization technique, Wang et al. (2009) studied the

thermodynamics of autothermal reforming of pure glycerol in a temperature range of

(700-1000 K), (1-12) for S/G molar ratio and (0-3) for O/G molar ratio. Glycerol overall

autothermal reforming reaction was expressed as: C3H8O3 + aO2 + bH2O → cCO + dCO2

27

+eH2 + fCH4. In order to maximize hydrogen yield and minimize the formation of

undesired products such as CO2, CO, CH4 and C species inside the reformer,

CHEMCAD-5 software was used to simulate the process and thermodynamically

establish the proper operating regions. Results for optimum temperature, steam/glycerol

and oxygen/glycerol molar ratios were reported as 627-727°C, 9-12, 0.0-0.4,

respectively. By investigating the thermodynamics of glycerol autothermal reforming for

hydrogen production, this study emphasized on the capability and feasibility of the

process and suggested more experimental work for further comprehension.

Directing research towards a more pragmatic route, Authayanun et al. (2010)

conducted a thermodynamic analysis on crude glycerol autothermal reforming using

HYSYS software. Free glycerol and methanol were considered to mainly compose crude

glycerol. Taking into account steam and oxidation reactions of both components beside

the possibility of methanation, methane dry reforming and water-gas shift reactions,

different profiles for process parameters were thermodynamically achieved at

atmospheric pressure. As expected, hydrogen yield increased with temperature and steam

to glycerol ratio, while an opposite trend was exhibited with oxygen to glycerol ratio, and

since the hydrogen capacity of methanol reforming is lower than that of glycerol, pure

glycerol showed a more favourable performance compared to the crude one, with a direct

dependency on the fraction of methanol in the crude glycerol.

Working in a pressure range of 200-300 atm this time, Ortiz et al. (2011) investigated

the thermodynamics of glycerol autothermal reforming at supercritical water conditions

for pure and “pre-treated” glycerol. Pre-treated glycerol was simply identified as 80% by

weight glycerol with the balance methanol. Using AspenPlusTM

software and again Gibbs

28

free energy minimization method, the optimum parameters for highest hydrogen yield

(88.4%) were obtained at: temperature of 900°C, pressure of 240 atm, water-to-glycerol

ratio of 99, oxygen-to-glycerol ratio of 0.41 and with a pure glycerol feed. Although the

increase in steam-to-glycerol ratio enhanced hydrogen yield, it also decreased hydrogen

molar flow rate due to the low concentration of glycerol in the feed and thus decreased

the overall production efficiency. To overcome this limitation and avoid the need of

external heat supply, thermoneutral conditions were established at different reforming

temperatures and water-to-glycerol ratios by finding out the amount of oxygen required

to internally produce this energy. Oxygen flow needed to achieve thermoneutral

conditions increased with water-to-glycerol ratio up to a maximum value of 0.407 at

800°C reaction temperature and 99 water-to-glycerol ratio.

In a follow up work by the same group and again with the aid of AspenPlusTM

, Ortiz

et al. (2012) conducted an energy integration and exergy analysis on the process of

autothermal reforming of glycerol using supercritical steam at 240 atm pressure. Initially,

the optimum conditions for hydrogen generation from a previous study (temperature of

900°C, pressure of 240 atm, water-to-glycerol ratio of 99, oxygen-to-glycerol ratio of

0.41) were implemented to simulate the process. Simulation results showed a significant

energy shortage and eventually a non-feasible process at these conditions; so adding the

heat generated from burning effluent gas in a separate combustor was the modification to

improve energy efficiency and take the process to a workable zone at lower reforming

temperatures. The counter trends of hydrogen yield and the combined efficiency (thermal

and exergy) at high reforming temperatures suggested a preheating temperature of less

than 500°C and a reforming temperature of not more than 800°C.

29

Yang et al. (2011) also carried out a thermodynamic analysis on pure glycerol

autothermal reforming using AspenPlusTM

for operating parameters varied within the

ranges of 0.5-8.0 for steam/carbon ratio, 0.5-3.0 for carbon/oxygen ratio and 400-850°C

for reforming temperature. Based on hydrogen yield and process feasibility, the optimum

ranges of carbon-to-oxygen ratio and reaction temperature were found out to go from 0.8

to 1.2 and 600 to 700°C, respectively. A steam-to carbon ratio of less than 3.0 was

recommended based on the fact that more energy consumption and less hydrogen

production will be caused by diluting glycerol feed, and consequently this might result in

a non-efficient process. The study suggested developing a new catalyst that can handle

ATR operations and avoid kinetic limitations in the presence of metal ions and bigger

biomass molecules within the crude glycerol feed.

2.4.4 HYDROGEN PEROXIDE REFORMING

This reforming technique is basically the reaction of biomass with hydrogen

peroxide to produce hydrogen and some other side products as reported by Luo et al.

(2007):

C3H8O3 + 1.5 H2O2 = 3CO2 + 5.5H2 (2.12)

In contrast with the upper mentioned techniques, hydrogen peroxide reforming has not

been an attractive method for research in the field of hydrogen generation from biomass,

and this is obvious from the very few publications in this regard. Based on the

thermodynamics of three reforming techniques, Luo et al. (2007) compared the

performance of water aqueous reforming, aqueous hydrogen peroxide reforming and

30

autothermal reforming. Hydrogen peroxide reforming showed the lowest capability for

hydrogen generation compared to the other two methods.

2.4.5 DRY REFORMING

Virtually, carbon dioxide reforming or dry reforming is the case when CO2 reacts

with glycerol to produce a mixture of carbon monoxide and hydrogen known as syngas.

Wang et al. (2009) utilised the power of Matlab software to perform a thermodynamic

study over ranges: temperature 327-727°C, pressure 1-5 atm, and CO2-to-glycerol ratio 0-

5. The results showed depletion in hydrogen yield with increasing the CO2 feed to the

reactor, while an opposite behaviour was reported for syngas. The maximum yield for

synthesis gas formation, 6.4 moles of syngas per mole of glycerol, was obtained at 1000

K with CGR of 1.

2.4.6 DRY AUTOTHERMAL REFORMING

As a combination of partial oxidation and dry CO2-reforming reactions, dry

autothermal reforming of glycerol has been studied by Kale and Kulkarni (2010) to

investigate the capability of employing this technique in the area of hydrogen and syngas

production. With the aid of HSE Chemistry software, a thermodynamic analysis was

conducted on the dry autothermal reforming of glycerol in a temperature range of 600-

100 K, oxygen-to-glycerol carbon “OCGR” of 0.1-0.5, CO2-to-glycerol “CGR” of 1-5

and at ambient pressure. T=926.31 K, P= 1bar, OCGR= 0.3 and CGR= 1 was determined

as the optimum thermoneutral operating conditions for DATR of glycerol. Maximum

capacity of hydrogen yield (2.88 mol of H2/mol of glycerol) was thermodynamically

31

obtained at T= 950 K, OCGR= 0.1 and CGR=1. The process was found to be more

feasible for syngas production with an ability to be utilised for hydrogen production.

2.4.7 AQUEOUS PHASE REFORMING

Many studies have been conducted on the aqueous phase reforming of glycerol

for hydrogen production [Shabaker et al. 2004; Wawrzetz et al. 2010; King et al. 2010;

Wen et al. 2008]. Davda et al. (2005) showed that aqueous phase reforming of biomass

such as glycerol or ethanol can play a promising role in hydrogen production in case of

developing proper catalysts and optimising the operating conditions. Lehnert & Claus

(2008) studied the aqueous reforming of glycerol over Platinum catalysts; after applying

different catalyst pre-treatments, the conversion glycerol was observed to stay around a

value of 20% with a significant increase in hydrogen selectivity to 95%. The premium

performance of MgO and ZrO2 supports in prompting Platinum catalyst activity for

hydrogen production via aqueous phase reforming of glycerol has been discovered by

Menezes et al. (2011). The fact of generating many side products in the liquid side by

aqueous phase reforming of glycerol makes the process more favourable for liquid target

products. Co-products such as methanol, ethanol, acetone, acetic acid, propanal 2-3

dihydroxyl, diglycerol and propylene glycol has been detected by Luo et al. (2008) and a

maximum of 42% carbon conversion was determined based on gas phase carbon; the fact

that confirms the formation of more liquid side by-products in the process.

32

2.5 KINETICS OF GLYCEROL REFORMING

In order to investigate the different effects of operating parameters on the rate of

glycerol reforming and hydrogen production, many scholars have studied the kinetics of

the corresponding reforming reactions. A kinetics analysis on glycerol aqueous

decomposition over bimetallic Pt-Re/C catalysts to produce synthesis gas has been

carried out by Kunkes et al. (2008). The study revealed the positive effect of integrating

Re catalyst into Pt supported on carbon carrier through increasing the value of turnover

frequency for synthesis gas production. Although the values of turnover frequencies for

H2, CO and CO2 production were reported based on the outlet gas composition and

number of active sites, the study did not generate a mathematical model to describe the

rate of glycerol reforming and its dependency on the major operating conditions.

On the other hand, the kinetics of crude glycerol pyrolysis was investigated by

Dou et al. (2009). According to results from thermogravimetric experiments, four distinct

phases in the thermal decomposition of crude glycerol were detected at 5 K/min heating

rate. During the first Phase of pyrolysis between 322 and 343 K, mass loss of about 10-

15% was observed due to the evaporation of water and methanol present in crude

glycerol. The main weight loss (67-69%) occurred during the second phase (322-440 K)

due to the removal of more than 95% of free glycerol, and liquid nitrate esters. The third

phase of crude glycerol pyrolysis (501-774 K) has been attributed to the decomposition

of fatty acids methyl esters and a small part of leftover free glycerol, and this contributed

to about 10.2-12.7 wt% loss. The fourth phase was found to go between 760 and 1123 K

in which less than 5.8% weight loss has been observed with less than 4.5 wt% residue at

the end. The activation energies and frequency factors of the four phases at four different

33

heating rates were reported by the study; as expected relatively low activation energies

were determined for phases 1&2 compared with phases 3&4.

The kinetics of glycerol steam reforming over Ni/CeO2 catalysts was studied by

Adhikari et al. (2009). A power law model of the form -ra=k0exp(-E/RT)[CA]n has been

fitted to illustrate the change of reaction rate with reforming temperature and glycerol

concentration. With the aid of SAS 9.1 software and the non-linear regression technique,

the activation energy and reaction order were determined as 103.4kJ/mol and 0.233,

respectively. The results showed a fair agreement between experimental and predicted

conversions with an average absolute deviation of about 6.7%. In line with this work,

Sutar et al. (2010) investigated the kinetics of glycerol steam reforming over Pt/C catalyst

in a temperature range between 623 and 673 K. The intrinsic kinetic data was again

collected in a fixed bed reactor by varying inlet flow (space velocity) and temperature

values. The reaction was found to have an order of one with respect to glycerol with a

good comparison between experimental and predicted conversions. In addition, Cheng et

al. (2010) studied the kinetics of glycerol steam reforming over a bimetallic Co-Ni

supported on alumina catalyst in a reaction temperature range between 773 and 823 K. A

power law model of the form –rGSR=Aexp(-EA/RT)[Pglycerol]β[Psteam]

γ was proposed to

describe the change in rate of glycerol steam reforming over the studied ranges. The

results of data regression demonstrated a reaction order of 0.36 with respect to steam and

0.25 with respect to glycerol with activation energy of 63.3 kJ/mol.

34

2.6 STATISTICAL APPROACHES

In many cases, statistics could be an effective tool to analyse, optimise and

interpret the behaviours of different variables in a chemical process. A statistical

approach to investigate the effects and interactions of different operating factors on

glycerol auto-thermal reforming process using nickel-based catalysts was conducted by

Douette et al. (2007). Using a fixed bed reactor, Douette et al. (2007) studied the effects

of different operating variables on process performance. Although a strong interaction

between reaction temperature and O/C ratio was revealed to significantly contribute to

the overall performance of ATR process, oxygen to carbon ratio was reported as the most

effective factor on conversion and H2 yield. Catalyst deactivation and coke formation was

observed after feeding crude glycerin obtained from biodiesel plants; this was again

interpreted by the high contaminants present in crude glycerin.

Recently, hydrogen and ethanol production from glycerol was investigated by

Varrone et al. (2012) in a biological reforming system. The authors utilised the power of

statistics and specifically design of experiments (DOE) section to optimize the process of

glycerol microbial fermentation. A stable fermentation conditions have been statistically

reached with an optimum ethanol capacity of 1 mol C2H5OH/ mol glycerol combined

with a hydrogen yield of 0.96 mol H2/ mol glycerol.

In a very recent study by Saha et al. (2014), the effects of introducing cobalt,

impregnating methods and promoting elements were statistically studied to better

understand the significance of these factors in enhancing the catalytic activity. According

to the results obtained from the analysis of variability (ANOVA), MgO and ZrO2

35

promoter elements, reverse impregnation and cobalt addition enhance the performance of

catalyst by modifying its textural properties to be more resistive to sulfur and more

suitable for heterogeneous catalysis. A comprehensive structure-activity analysis was

conducted to explain the impacts of these controlled variables on the activity of catalyst,

and a linear model to correlate catalysts’ textural properties to conversion has been

developed as follows: Conversion= 407.27 – 1.76 BET SA -13.76 PV/PS + 13.9 M SA –

16.46 M dispersion – 43.36 Ni Reducibility. In addition, a Pareto chart was constructed to

distinctly quantify the effects of these properties and as a result, nickel dispersion was

found to mostly contribute to catalyst performance.

2.7 CATALYSTS FOR GLYCEROL REFORMING

Catalyst is a substance that speeds up the rate of a chemical reaction without

being consumed itself. The accelerated rate of catalyzed reactions is due to the higher

contact frequency between reactants involved in the rate-limiting step (slowest step), and

this will consequently reduce the activation energy making the reaction easier to take

place (McNaught & Wilkinson, 1997). In general, catalysts can be classified into two

main categories; homogenous where the catalyst and reactants have the same phase and

heterogeneous where the phases are different.

Several metal-based catalysts have been used to promote glycerol reforming and

increase hydrogen productivity such as Ruthenium, Nickel, Rhodium, Cobalt, Platinum

and Iridium. The efficiency of Nobel-metal catalysts such as Ru, Rh, Ir and Pt in glycerol

reforming to produce hydrogen has been widely investigated [Pompeo et al., 2010;

Dauenhauer et al., 2006; Rennard et al., 2009; Slinn et al., 2008; Kunkes et al., 2008;

36

Zhang et al., 2007; Hirai et al., 2005]. On the other hand, base-metal catalysts such as

Nickel and Cobalt showed very promising results in glycerol reforming [Czernik et al.,

2002; Adhikari et al., 2007; Adhikari et al., 2008; Iriondo et al., 2010; Swami and

Abraham, 2006; Doette et al., 2007]. According to Iriondo et al. (2009), Nickel supported

on modified γ-Al2O3 revealed a superior catalytic activity compared to Platinum

supported on the same carrier in steam reforming of glycerol. In line with this

observation, Nickel supported on modified cerium was strongly recommended by

Buffoni et al. (2009) for glycerol steam reforming operations due to its activity, stability

and high selectivity. In addition, a study by Kamonsuangkasem et al. (2011) emphasized

on the stability of Ni/CeO2-ZrO2/Al2O3 catalyst in glycerol oxidative steam reforming

process. Moreover, Nickel catalyst and CeO2 based supports were recommended by

Vaidya and Rodrigues (2009) for hydrogen production from biodiesel by-product

glycerol.

Nickel catalyst over Ce-Zr mixed oxide support promoted with La, Ca, Mg, Gd

and Y elements was studied by Sengupta (2011) in order to investigate its efficiency in

hydrogen production from hydrocarbons. The results showed high thermal stability, low

coke formation and promising hydrogen yields. In a follow up study by same research

group, auto-thermal reforming of pure glycerol using nickel supported over CeZrM

catalysts was investigated by Sabri (2013) in a packed bed tubular reactor to

experimentally establish the optimum reforming regions. Based on her results, the

optimum conditions for nickel loading and surfactant-to-metal ratio were reported as 5

wt% and 0.5, respectively. Gadolinium was found to best perform as a promoter element

for pure glycerol reforming.

37

Chapter 3

EXPERIMENTAL SECTION

This chapter includes the procedures, chemicals, equipments and gases used to

carry out the experimental work and collect the required data for this research. Besides

experimental setup and design of experiments (DOE) parts, more details about the

methods of catalyst preparation, characterization and performance evaluation will also be

given.

3.1 SAFETY PRECAUTIONS

Adequate precautions have been taken to ensure a safe and healthy work

environment while conducting our experiments. As a strict perquisite to a vigorous

performance in the laboratories of the Process Engineering Department; all students,

researchers and employees must fully adhere to the following terms in the workplace;

otherwise penalty and/or dismissal steps will be taken.

1. Proper lab training, safety exams and safety orientation sessions such as WHIMS,

Chemical and Laboratory Safety Training, H2S Training and Radiation Safety

Training were completed prior to acquiring the access to the laboratories.

2. Personnel Protective Equipments (PPE) such as safety goggles, lab coats, gloves

and closed shoes were used inside the laboratories regardless if the person is

running an experiment or not.

3. Volatile toxic chemicals such as Ammonium Hydroxide utilized for catalyst

preparation were always handled in the fume hood.

4. All chemicals were labeled and stored properly.

38

5. Flammable chemicals in particular were stored in the designated cabinets of the

storage rooms.

6. Chemical Wastes generated during catalysts synthesis were stored, labeled and

disposed appropriately.

7. Carbon Monoxide (CO) and Hydrogen Sulfide (H2S) detectors were installed in

the labs where cylinders containing these gases were present.

8. Gas cylinders were transported and stored rightly. All cylinders were secured to a

wall or bench and capped well when not in use.

9. Neat and clean workplace was always ensured to keep a motivating healthy

working atmosphere.

3.2 CATALYST PREPARATION

3.2.1 Chemicals Used

1. Nickel (II) nitrate hexahydrate, Ni(NO3)2.6H2O, 99.99%; Sigma Aldrich

2. Zirconium (IV) oxynitrate hydrate, ZrO(NO3)2.xH2O, 99.99%; Sigma Aldrich

3. Cerium (III) nitrate hexahydrate, Ce(NO3)2.6H2O, 99%; Sigma Aldrich

4. Calcium (II) nitrate tetrahydrate, Ca(NO3)2.4H2O, 99%; Sigma Aldrich

5. Gadolinium (III) nitrate hydrate, Gd(NO3)3.xH2O, 99.99%; Alfa Aesar

6. Magnesium nitrate, Mg(NO3)2.6H2O, 98%; Sigma Aldrich

7. Ammonium hydroxide reagent ACS-Pure; (28-30% w/w); BDH Chemicals

8. Hexadecyltrimethylammonium bromide (CTAB), C19H42NBr; Sigma Aldrich

3.2.2 Equipment Utilized

1. Isotemp Muffle furnace, Model 550-126, Fisher Scientific Company

39

2. Hydraulic Press, model 3912, Carver

3. Weighing Balance, PB8001-S, MonoBloc

4. Magnetic Stirrer Hot Plate, Gyratherm II a, VWR Scientific Inc.

5. U.S.A standard test sieve, ASTM, E-11 standards, Fisher Scientific

3.2.3 Support Preparation

With the aid of Surfactant-Assisted method, the ternary oxide supports utilized in

this study were prepared under basic conditions to nominally have a composition of

Ce0.5Zr0.33M0.17, where M is the promoter element. Three different promoter elements

were tested in this study: calcium, gadolinium and magnesium (Ca, Mg and Gd). The

selection of the support was based on the experimental screening results obtained from a

pure glycerol study by our group (Sabri, 2013). In this study, cerium-zirconium based

supports promoted with gadolinium, calcium and magnesium elements showed promising

results in the process of pure glycerol autothermal reforming. The reasons behind the

enhanced performance of the promoted Ce-Zr supports have been discussed by Ibrahim

and Idem (2008).

As a rule, nitrates precursors (Ce, Zr and M) with prearranged loads were firstly

dissolved in separate beakers, each containing 500 ml of deionized (DI) water. Rigorous

stirring was applied to facilitate the diffusion of these salts into the aqueous face. In the

meantime, a predetermined amount of the surfactant (CTAB) was also dissolved in 1000

ml of deionized (DI) contained in a separate beaker heated to ~60°C and under a rigorous

stirring too. After that, the contents of the four beakers were mixed together to obtain a

2500 ml resultant solution. The measured weights of nitrate precursors and surfactant

40

were determined according to Ce0.5Zr0.33M0.17 nominal composition and 0.5 surfactant-to-

metal ratio [CTAB]/[Ce+Zr+M], respectively. Aqueous ammonium hydroxide (28-30%

w/w) was then added gradually to increase the pH of the mixture to ~11.8. The big beaker

containing the resultant solution was also maintained under continuous stirring while the

addition of 50 ml of NH4OH solution was taking place every 5 minutes. The pH of the

solution was measured repeatedly using standard pH papers. After reaching the desired

pH (~11.8), the mixture was left for about 20 min to clearly see yellowish gelatinous

slurry precipitated at the bottom. Then the mixture was transferred to Pyrex glass bottles

and aged in an oven at 90°C and atmospheric pressure for 5 days. After five days, the

bottles were taken out of the oven and left in a fume hood overnight to slowly cool down.

On the next day, the precipitated material was collected by filtering the mixture using 2

filter papers [125 mm diameter, #2, Whatman], and then rinsed with warm deionized

water to wash out any leftover surfactant and nitrates. After that, the retained solid was

dried out in the oven for 12 hours at 120°C. The collected dry material was then crushed

and collected in special ceramic crucible to be calcined at 650°C for another 3 hours. The

aim of calcination was to remove the remaining surfactant and nitrates; and thus to obtain

a support composed of free metal oxides. Indeed, similar preparation procedures were

followed by other studies in our group [Oluku, 2013; Sengupta, 2011; Sukonket et al.,

2011]. Next steps were to collect the prepared supports, impregnate the active metal

(Nickel), investigate their physical and chemical properties via several characterization

techniques and finally screen their performance in a PBTR for their catalytic efficiencies

in the autothermal reforming process.

41

3.2.4 Nickel Impregnation

With the help of the standard wet impregnation technique, 5 wt% Nickel

[mNi]*100/[mSupport] was impregnated over the prepared supports. Table 3.1 lists the

resultant catalysts with a composition of 5%Ni/Ce0.5Zr0.33M0.17 where M= Ca, Mg and

Gd. Nickel nitrate solution was prepared by dissolving a predetermined amount of

Nickel nitrate hexahydrate Ni(NO3)2.6H2O in a 1 litre of deionized water to obtain a 0.1

M aqueous solution. Then a volume of 25.5 ml of the prepared nickel solution was mixed

with 2.85 g of the support in a round bottom flask. The flask containing both the solid

support and the catalyst solution was then placed over a magnetic stirrer with heater

[Gyratherm Πa, VWR Scientific Inc.] at a temperature of 80°C and a stirring rate of 240

rpm for ~13 hours to autogenously dry out the water. Following this, the obtained solid

catalyst was calcined for 3 hours to remove any remaining moisture and nitrates.

Calination temperature was varied between 550 and 650°C in an Isotemp Muffle furnace

[Model 550-126]. Finally, the catalyst became ready to be characterized, pelletized and

tested in a PBTR. Different pellet sizes (0.5-1.27 mm) were prepared using hydraulic

press [model 3912, Carver] and standard sieves [ASTM U.S.A., E-11 standards].

Table 3.1 Compositions of catalysts prepared

Impregnation Technique Catalysts (CTAB= 0.5)

Standard Wet Impregnation

(5 wt% Nickel)

Ni /CeO2-ZrO2-GdOx

Ni /CeO2-ZrO2-MgOx

Ni /CeO2-ZrO2-CaOx

42

3.3 CATALYST CHARACTERIZATION

In order to understand the relation between the characteristics of the prepared

catalysts and their behaviours in the process, fresh supports and catalysts were

characterized for their physical and chemical properties. Spent catalysts were also

characterized to investigate the effects of different operating parameters on catalysts’

properties. The catalytic activity was then statistically correlated to the structural and

chemical properties based on the observed performance from the screening experiments.

3.3.1 Thermogravimetry Analysis (TGA)

Thermogravimetric analysis was conducted on fresh catalysts to determine the

calcination temperature. The uncalcined catalysts were subjected to a programmed

temperature increase under inert gas (N2) flow at atmospheric pressure to study the

weight loss and the rate of weight change due to the thermal removal of moisture,

surfactant and nitrates. TGA runs were performed using Shimadzu TGA-50 instrument

connected to a flow controller FC-60A.The applied temperature on the fresh catalyst

samples was 800°C controlled from TA-60WS Collection software. The flow of nitrogen

was adjusted at 50 ml/min from the FC-60A flow controller. The heating rate was set up

at 10°C/min. The cooling of the furnace assembly was done after each run using an

external fan (BLW-50, Shimadzu) purchased alongside with TGA-50 instrument.

43

3.3.2 N2 Physisorption (BET)

N2 physisorption technique was applied to investigate some structural properties

such as BET surface area, pore volume, pore size and pore size distribution of the

prepared supports and catalysts. The analysis was carried out under pure nitrogen flow

(99.99% pure N2; Praxair) using ASAP-2010 apparatus from Micromeritics Instruments

Inc. The amount of catalyst used in each run was about 0.25 g. The samples were first

degassed for about 4 hours at a temperature of 180°C to get rid of the adsorbed moisture

on the surface that can negatively affect the overall analysis and yield inaccurate results.

The isotherms of adsorption and desorption phases of nitrogen on the surface of the

catalyst samples were established at a temperature of -195°C using liquid nitrogen. The

previous mentioned properties related to BET area and porosity were determined based

on these isotherms.

3.3.3 Temperature Programmed Oxidation (TPO)

In order to determine the amount of carbon deposited on the spent catalyst during

the ATR of synthetic crude glycerol, TPO runs were performed using TGA-50

(Shimadzu, Japan) instrument. Catalyst particles were carefully separated from the solid

bulk found inside the PBTR after each screening run. The collected sample was then

oxidized at high temperature to burn the deposited carbon species. The oxidizing gas flow

rate was 50 ml/min (5% O2 balanced with N2; Praxair). The applied heating was

maintained at a constant 10°C/min rate to reach a maximum temperature of 800°C by the

end of the analysis. The coke formed over the surface of the catalyst was quantified by

the weight loss observed between 360 and 600°C. According to Le Minh et al. (1997),

44

formation of CO and CO2 species during TPO analysis on spent catalysts was observed

starting from a temperature around 380°C. In line with this observation, TPO peaks

observed by Nagaoka et al. (2000) for the carbon deposits over spent Pt/ZrO2 and

Pt/Al2O3 catalysts were also obtained between 342 and 567°C. In this work, the major

TPO peak was obtained around 550°C with a small shoulder peak around 400°C, similar

to the curve profiles and temperature ranges obtained also by [Li & Brown (1999); Li &

Brown (2001); Guo et al. (2007)]. The rate of carbon deposition over catalyst surface [mg

C/g cat.h] at different operating parameters was then determined by dividing the obtained

weight of carbon deposits over the weight of the catalyst used at that particular run and

the screening time on stream (6 hours).

3.3.4 Temperature Programmed Reduction (TPR)

The reducibility of the catalyst was determined by the temperature programmed

reduction (TPR) using ChemBET-3000 instrument [Quantachrome, USA] to find out the

optimum reduction temperature required to activate the catalyst before screening its

performance. A catalyst weight of 50 mg was used in each analysis. The sample was

firstly degassed with pure nitrogen flow (UHP 5.0; Praxair) at 180°C for 2 hours. Then,

the sample was moved to the analysis port where a flow of 45 ml/min of the reducing gas

(5% H2 balanced with N2; Praxair) was introduced. The baseline signal of the thermal

conductivity detector (TCD) was then adjusted to zero at ambient temperature. Next, the

temperature of the furnace was raised up to 900°C at a constant heating rate of 15°C/min.

The resultant TPR profiles were plotted by TPRWIN software as the intensity of thermal

conductivity of the effluent gas [mV] versus temperature [°C]. Reducibility was then

45

determined based on the temperature that maximizes the rate of hydrogen consumption

(highest TPR peak).

3.3.5 Powder X-Ray Diffraction Measurements (XRD)

Powder XRD patterns were established using D8 diffractometer (Bruker AXS,

USA) equipped with Ni-Kα radiation source at 40 mA and 40 KV to study the

crystallinity of supports and catalysts and investigate the uniformity of active sites

dispersion over the surface of the carrier material. A 2θ range of 10° - 90° with a step size

of 0.02° and a counting time of 1 second per point was applied to collect the intensity

data. Data from the international centre of Diffraction Data (ICDD) was used as a

reference to identify the crystalline phases. The peaks of nickel were identified by

analyzing a NiO sample and plotting the corresponding profile at the top supports and

catalysts profiles.

3.3.6 Inductively Coupled Plasma-Mass Spectrometry (ICP-MS)

ICP-MS analysis was conducted to quantify the nickel amount incorporated over the

support and obtain the elemental composition of the catalysts. The fundamental idea of

ICP-MS analysis is the ionization of metal atoms at a very high temperature (> 6000°C)

in order to separate them. The ionization part includes the digestion of a sample in a

mixture of specific acids, and then pumping it to a nebulizer in the presence of argon gas.

At the time the dissolved sample reaches the plasma region, the acids will eventually

evaporate and the solid part will break down into atoms and then into ions at the

extremely high plasma temperature. ICP-MS analysis was conducted at the department of

chemistry, University of Regina.

46

3.4 REACTION FEED STOCK

The main reactants in the autothermal reforming process of the current synthetic fuel are:

1. Synthetic Crude Glycerol

2. Steam

3. Oxygen

3.4.1 Synthetic Crude Glycerol Feed

The synthetic crude glycerol utilized in this study was prepared in our lab by mixing

the commonly found chemicals in the commercial crude glycerol generated at biodiesel

plants. In order to investigate the individual effects of crude glycerol components, and to

follow the scientific path of research, a synthetic feed was reformed in this study as the

next step after reforming pure glycerol in a previous study by our group. The next step

after that would be testing crude glycerol mixtures directly obtained from biodiesel

plants. A batch of about 120 ml containing free glycerol, water, methanol, soap, free fatty

acids (FFA) and ashes was prepared prior to each screening run. These are the common

chemicals found in crude glycerol mixtures. An excellent separation of biodiesel was

assumed in the production process so as, glycerol stream does not contain any significant

amounts of methyl esters. This assumption was based on the composition of the crude

glycerol collected from Milligan biofuels, where the weight percentage of fatty acids,

canola oil and methyl esters all together was less than 10 wt%. Potassium hydroxide was

also exempted from the synthetic CG and just ashes of NaCl and KCl were considered.

According to a recent comprehensive study conducted by Hu et al. (2012) on five

different samples of crude glycerol mixtures, sodium and potassium hydroxide salts were

47

not present in the mixtures, but only traces of ashes. The ashes were provided by the

addition of sodium and potassium chlorides (NaCl and KCl) [Ito et al. (2005); Bournay et

al. (2005); Thompson & He (2006)]. The synthetic CG was then prepared by mixing the

individual components in one beaker to yield a mixture with an average molecular

formula of C2.5H7O2 and an average molecular weight of 69.08 g/mol. Average molecular

formula calculations are given in Appendix E. Table 3.2 includes the physical properties

and weight percentages of the below listed chemicals used in synthetic CG preparation.

Glycerol (56-81-5; Sigma Aldrich)

Methanol (67-56-1; Fisher Scientific)

Potassium Palmitate Standard Solution (2624-31-9; Sigma Aldrich)

Oleic Acid (112-80-1; Alfa Aesar)

Potassium Chloride (7447-40-7; Alfa Aesar)

Sodium Chloride (7647-14-5; Alfa Aesar)

Deionized water (ITC labs)

48

Table 3.2 Physical properties and percentage composition of the synthetic CG

Chemical

Compound

Mass

Fraction

(%)

Density

(g/ml)

Mol.

weight

(g/mol)

Purity

(%)

Chemical

Formula

Glycerol 45.6 1.26 92.09 >99.5 C3H8O3

Methanol 11.2 0.79 32.04 >99.9 CH4O

Water 8.3 1.00 18.02 >99.9 H2O

Soap 29.1 0.87 294.51 > 99.0 C3H31O2K

Oleic Acid 3.8 0.89 282.47 >90.0 C18H34O2

Sodium Chloride 1.0 2.17 58.44 >99.0 NaCl

Potassium Chloride 1.0 1.98 74.55 99.0-100.5 KCl

3.4.2 Steam Feed

The steam was provided by adding a predetermined amount of water to the

mixture of CG. This allows the formation of high temperature steam inside the PBTR.

The amount of water was determined based on the steam-to-carbon ratio that needs to be

adjusted at each run.

3.4.3 Oxygen Feed

Oxygen was fed to the system from a compressed air cylinder obtained from Praxair

Inc. According to the compressed air MSDS from Praxair; the concentration of oxygen is

ranged between 19.5 and 23.5% balanced with nitrogen. The percentage of oxygen was

reasonably taken to be 21% balanced with nitrogen to ease the calculations.

49

3.5 PERFROMANCE EVALUATION

Catalytic activity and process performance have been experimentally investigated in

a packed bed tubular reactor at various operating conditions. The effects of the different

components of crude glycerol were first tested by adding them one at a time to the

synthetic feed stock. The varied parameters in this work include the promoter element,

reduction temperature, calcination temperature, particle size, reforming temperature,

steam-to-carbon (S/C) ratio and oxygen-to-carbon ratio (O/C). Besides the extended

period stability run that was conducted for 15 hours time on stream (TOS).

3.5.1 Experimental Setup

The utilized experimental setup was composed of several units as shown in Fig. 3.1.

Packed Bed Tubular Reactor (Inconel 0.5” ID, Homemde at UofR)

Electric Tubular Furnace (ZCP 386, Zesta Engineering Ltd.)

Two K-type thermocouples (1.6 mm x 18", Zesta Engineering Ltd.)

Thermal Mass Flow Controller (DFC26, Aalborg Instruments)

Mass Flow Meter (GFM 17, Aalborg Instruments)

Syringe Pump (100 ml Stainless Steel, KD Scientific)

Gas Chromatograph (GC-TCD, HP 6890, Agilent Technologies)

Pure Nitrogen Cylinder (UHP 5.0,Praxair)

Compressed Air Cylinder (Praxair)

Hydrogen-Nitrogen Cylinder (5% H2-balance N2, Praxair)

Condenser (Water-cooled, Homemade at UofR)

Condensate Collector (Ice-cooled, Homemade at UofR)

50

Figure 3.1 Schematic flow diagram of the experimental setup for synthetic CG autothermal

reforming in a PBTR

51

3.5.2 Catalytic Activity Evaluation

The main target of the experimental part was to optimize the various operating

variables and collect the required intrinsic data for the kinetic study. The components of

crude glycerol were added to the synthetic feed separately to investigate their individual

effects. Three promoter elements were tested (Gd, Mg and Ca). Three calcination

temperatures (550, 600, 650°C) and three reduction temperatures (500, 600, 700°C) were

also applied. Kinetic regions were established by varying particles size (0.55, 0.8, 1.09,

1.27 mm) and carbon flow rate (0.0019, 0.0026, 0.0033 mol C/min). Different catalyst

weights were used (0, 0.05, 0.1, 0.15, 0.2, 0.25 g) at four different reforming

temperatures (500, 550, 600, 650°C). All screening experiments were carried out at

atmospheric pressure.

In order to create a plug flow case and avoid heat and mass transfer limitations,

catalyst bed height (L) and particle diameter (dp) were selected with respect to reactor

diameter (d) so that d/dp≥10 and L/dp≥ 50. This selection was in line with the criteria

followed by [Froment & Bischof (1990); Idem & Bakhshi (1996); Ibrahim & Idem

(2006); Akpan et al. (2007)]. The effect mass diffusion from the bulk to the external

surface of the catalyst was tested by varying the flow rate inside the PBTR. The external

mass transport is strongly influenced by the thickness of the boundary layer that

surrounds the catalyst pellet. The thickness of this layer is inversely proportional to the

mass transfer coefficient and directly proportional to the diffusivity. A low flow velocity

might create a thick boundary layer and thus a small mass transfer coefficient. In this

case, the external mass transfer is the slowest step and the determination of intrinsic

reaction kinetics becomes impossible. On the other hand, catalyst particle diameter was

52

varied to make sure that the internal diffusion is not limiting the rates of reactions and

thus the slowest step is not the internal transfer inside the particles. The effects of heat

transport are totally analogous to those of mass transfer (Fogler, 2011). In the current

work, the ratios of bed length to particle diameter L/dp and reactor diameter to particle

diameter d/dp were maintained at 56.3 and 15.9, respectively after selecting the optimum

diameter for catalyst pellets.

In a typical experimental run, the reactor tube was first removed from the apparatus,

cleaned well with acetone and then installed back to the furnace assembly. The catalyst

was then mixed with 7.6 g of α-Al2O3 particles (0.8 mm) and loaded inside the reactor to

obtain a bed height of ~4.5 cm. The reactor was then heated to the desired reduction

temperature while an inert flow of 100 ml/min of nitrogen was maintained. The

temperature inside the catalyst bed was measured by a K-type thermocouple inserted

from the top of the reactor. The activation of catalyst was then performed by applying a

constant flow of 100ml/min of a reducing gas (5% H2; balance N2) for 2 hours. The

purpose of reduction was to get rid of the oxygen atoms in the metal oxides (NiO and

CeO) and obtain the free active metals Ni and Ce. Meanwhile, the synthetic crude

glycerol was prepared by weighing a specific amount of each component (error<1%).

The different components of synthetic CG plus the predetermined amount of water were

then mixed in a beaker and stirred for about 20 min to obtain a homogenous mixture of a

total volume ~ 120 ml. The syringe pump was then loaded with ~90 ml of the prepared

mixture. Next, the desired reaction temperature was adjusted from the control panel of

the tubular furnace around the reactor. The syringe pump was regulated at 0.15 ml/min

flow rate and turned on just after reaching the desired reforming temperature. At the same

53

time, the inlet flow of oxygen was started from the DFC software that connects to the gas

flow controller. The flow of air was adjusted between 19 and 126 ml/min based on the

oxygen-to-carbon ratio in the inlet feed. The product gas from the PBTR was passed

through the water-cooled condenser (16°C) and the condensate flowing out was collected

in an ice-cooled stainless steel collector. The outlet flow of gas from the liquid collector

was first measured by a digital flow meter, and then sent to a gas chromatograph

equipped with Hayesep Q and Molecular Sieve A columns to analyze for the different

gaseous species produced. The liquid samples were collected and stored in a fridge after

each run.

54

3.6 PERFORMANCE EVALUATION CRITRERIA

The main possibly involved reactions in the autothermal reforming of the synthetic

crude glycerol used in this study include the steam reforming and partial oxidations of

free glycerol, methanol and 2-propanol (present in the soap); besides the water gas shift

(WGS), methanation and methane dry reforming reactions. Reaction expressions with the

corresponding calculated change in the enthalpies (ΔHr) at a constant pressure are given

in the following:

Steam Reforming

C3H8O3 + 3H2O 7H2 + 3CO2 (3.1)

ΔHr (25C) = +122.5 kJ/mol

CH3OH + H2O 3H2 + CO2 (3.2)

ΔHr (25C) = +49.5 kJ/mol

C3H7OH + 5H2O 9H2 + 3CO2 (3.3)

ΔHr (25C) = +283.8 kJ/mol

Partial Oxidation

C3H8O3 + 1.5O2 4H2 + 3CO2 (3.4)

ΔHr (25C) = -603.5kJ/mol

CH3OH + 0.5O2 2H2 + CO2 (3.5)

55

ΔHr (25C) = -102.5KJ/mol

C3H7OH + 2.5O2 4H2 + 3CO2 (3.6)

ΔHr (25C) = -926.2 kJ/mol

Water Gas Shift

CO + H2O H2 + CO2 (3.7)

ΔHr (25C) = -41.8 kJ/mol

Methanation

CO + 3H2 CH4+ H2O (3.8)

ΔHr (25C) = -206.8 kJ/mol

Methane Dry Reforming

CO2 + CH4 2H2 + 2CO (3.9)

ΔHr (25C) = 246 kJ/mol

The combination of steam reforming and partial oxidation reactions of the main

reactants generate three autothermal reforming reactions as the following:

Glycerol Reforming

C3H8O3 + H2O + 0.5 O2 2CO2 + CO + 5H2 (3.10)

ΔHr (25C) = -78.3KJ/mol

56

Methanol Reforming

CH3OH +0.25 H2O+0.125O2 0.5CO2+ 0.5CO+2.25H2 (3.11)

ΔHr (25C) = 9.6 KJ/mol

Propanol Reforming

C3H7OH + 2H2O + O2 2CO2 + CO + 6H2 (3.12)

ΔHr (25C) = -159.0 KJ/mol

Integrating the different components of the synthetic crude glycerol into an

averaged molecular formula will yield an overall formula of C2.5H7O2. The different

stoichiometries for synthetic CG autothermal reaction (3.13) at different inlet ratios of

steam-to-carbon (S/C) and oxygen-to-carbon (O/C) are given in Table 3.3. Conversion

was calculated to determine the amount of reactants that has been converted into

products. However, conversion is a blind criterion if not combined with selectivity.

Selectivity shows the intensity of reactants conversion into target product which is

hydrogen in this case. Yield was determined as the combination of conversion and

selectivity to ease the optimization process. Turnover frequency was determined to select

the most active catalyst. It reveals the amount of product produced per catalytic cycle per

unit time, which in other words represents the activity of the catalyst. Rate of carbon

depositions was also measured in order to minimize coke formation and subsequently

attain a stable process for hydrogen production.

C2.5H7O2 + aH2O + bO2 2CO2 + 0.5CH4 + cH2 + dH2O (3.13)

57

Table 3.3 Different reaction Stoichiometries at different S/C and O/C ratios

S/C O/C a b c d

1.6 0.05 4 0.25 4 2.5

2.6 0.05 6.5 0.25 4 5

3.6 0.05 9 0.25 4 7.5

1.6 0.125 4 0.625 3.25 3.25

2.6 0.125 6.5 0.625 3.25 5.75

3.6 0.125 9 0.625 3.25 8.25

1.6 0.2 4 1 2.5 4

2.6 0.2 6.5 1 2.5 6.5

3.6 0.2 9 1 2.5 9

58

The performance criteria followed in this study are:

Crude glycerol conversion was defined as:

Hydrogen Selectivity was defined as:

RR: Fraction Ratio of H2 to CO2 based on the stoichiometric ratio of the reaction

Hydrogen Yield was defined as:

Rate of Coke Formation was defined as:

Turnover Frequency was defined as:

59

3.7 Design of Experiments (DOE)

The experimental phase of this work was divided into two parts: (1) catalyst

development, (2) optimization and intrinsic data collection. The individual effects of

synthetic CG components were early investigated by adding them separately to the feed

mixture. Promoter element, calcination and reduction temperatures were varied in order

to obtain an active and stable catalyst. The catalyst was selected based on the highest

exhibited activity. The operating parameters were then optimized to attain the best

conditions for a stable hydrogen yield. Then, the kinetic regions were established by

changing the catalyst pellets size and inlet flow rate. Kinetic experiments were designed

based on the factorial design methodology using two combinations of the four kinetically

controlled variables (1) reaction temperature, (2) steam-to-carbon ratio, (3) oxygen-to-

carbon ratio and (4) weight space time W/FA0. Three levels for S/C and O/C factors were

selected, and therefore 3x3 experiments were conducted. On the side of reaction T and

w/FA0 factors, 4 and 6 levels were selected for each factor, respectively, to result in 4x6

experiments. The regions selected for the different parameters were in line with the

thermodynamics and experimental studies discussed in chapter 2. These are given in

Tables 3.4 (a,b,c).

60

Table 3.4a Parametric experiments

Factors Levels

Synthetic CG

Components + methanol + Soap + FFA + Ashes

Promoter

Element Gadolinium Calcium Magnesium

Calcination

Temperature

(°C)

550 600 650

Reduction

Temperature

(°C)

500 600 700

Flow Rate

(ml/min) 0.15 0.2 0.25

Catalyst Pellets

Size (mm) 0.55 0.8 1.09 1.26

Extended TOS

Stability 16 hours TOS

Table 3.4b Kinetic experiments (variables: reactant concentrations)

O/C

S/C

1.6 2.6 3.6

0.05

0.125

0.2

Table 3.4c Kinetic experiments (variables: temperature and W/FA0)

Wcat/FA0

(gcat.min/ mol C)

Reaction Temperature (°C)

500 550 600 650

0.0

12.7

50.9

76.4

101.9

127.4

61

Chapter 4

RESULTS AND DISCUSSION

4.1 CATALYST CHARACTERIZATION

The developed catalysts were characterized for the sake of correlating the

observed catalytic activity to their chemical and structural properties. Investigating the

most affecting factors will help to reasonably select the proper catalyst and optimize the

process performance. In the current study, three promoter elements were checked for

their capabilities of enhancing the performance of the cerium-zircoinum supporting

material Ce-Zr-M, where M is the promoter element Mg, Ca or Gd. The prepared

supports were then impregnated with 5 wt% nickel to obtain an overall catalyst formula

designated by 5Ni/CeZrM.

4.1.1 Thermogravimetric Analysis (TGA)

Thermo-gravimetric analysis was conducted on the uncalcined supports and fresh

impregnated catalysts to investigate the optimum calcination temperature. The catalytic

material has gone through two calcination stages: (1) on the fresh supports just after

drying (2) on the fresh catalysts just after impregnation. The obtained TGA profiles are

shown in Figure 4.1. In these profiles, TGA% represents the weight change with respect

to the initial weight, and DrTGA represents the derivative value of weight change with

respect to time; in other words the time rate of weight change.

In the case of CeZrM support, the first peak of weight loss was observed between

20°C and 100°C where the rate of weight loss reached a maximum value of 0.3 mg/min.

62

This stage corresponds to the removal of moisture from the pores of the support and the

hydrated salts. The hydrated salts associated in this case are Ce(NO3)2.6H2O,

ZrO(NO3)2.xH2O and Ca(NO3)2.4H2O. The second stage of weight loss was identified by

a second peak between 180 and 400°C. This peak attributes to the removal of surfactant

(CTAB) and nitrates (NO3-) remained in the collected solid material (support) after

filtration and rinsing with hot deionized water.

In the case of 5Ni/CeZrM catalyst, the first stage of weight loss was also observed

between 20 and 100°C with a maximum weight loss rate of 0.15 mg/min. Along with the

moisture removed from the pores of the catalyst and the hydrated salts mentioned

previously, Ni(NO3)2.6H2O also contributes to the moisture peak in the catalyst profile.

The second peak of weight loss appears between 150 and 500°C. The wider peak in this

case is justified by the larger amount of nitrates present and their dispersion in the

internal pores of the catalyst. At the time of support preparation, most of the used nitrates

would have gone with the water through filtration and rinsing, unlike the impregnation of

the active material where all the nitrates present in the aqueous Nickel nitrate solution got

aggregated inside the catalyst. According to these results, an optimum calcination

temperature of 550°C was recommended.

63

Figure 4.1 TGA profiles of CeZrCa support and the corresponding 5wt% catalyst

64

4.1.2 N2 Physisorption and ICP-MS on Fresh Supports and Catalysts

N2 physisorption technique was applied to measure the specific surface area,

specific pore volume and pore size of the prepared supports and catalysts. The obtained

isotherms shown in Figure 4.2 manifest the volume change of nitrogen adsorbed at in a

range of relative pressure (P/P0) between 0 and 1.0, where P is the partial vapor pressure

of nitrogen at 77.4 K and P0 is the saturation pressure of nitrogen. The type of the

structure can be identified from the patterns of the isotherms. Obviously, the structure of

our catalyst can be categorized under “Type IV” group, which belongs to the mesoporous

material that has a strong adsorption affinity [Pierotti & Rouquerol, (1985); Schneider,

(1995)]. According to IUPAC Recommendations (1994), the adsorption hysteresis of Ca

tends to belong to H2 type where the interpretation of the loop hysteresis is complex, and

the role of the overall network effects must be considered for defining the distribution of

pore size and shape. On the other hand, H3 type loops corresponding to Gd and Mg

indicate the presence of plate-like particles that yield split-shaped pores where the

limiting adsorption tends to be null at high P/P0.

In this work, the collected measurements of BET surface area, average pore

volume, average pore diameter along with the actual mass fraction of nickel are given in

Table 4.1. The ranges of specific area and pore volume of the prepared catalytic material

were 115.3-187.3 m2/g and 0.23-0.45 cm

3/g, respectively. The average pore size was in

the range of 55.0-109.4 A° within the range of mesoporous material specifications (20-

500°A). Mesoporous materials are widely applied in the field of heterogeneous catalysis

due to the large surface area they provide. The high surface area of the catalyst increases

the rate of the reaction through providing a large number of active sites.

65

Table 4.1 BET and ICP-MS results of fresh supports and catalysts

Promoter

Element

BET Surface Area

(m2/g)

Pore Volume

(cm3/g)

Pore Size

(A0)

Pore V. / S. Area

(10-9

m)

Supports CeZrM, Surfactant/Metal = 0.5

Mg 187.3 0.39 67.3 2.08

Ca 124.4 0.25 63.1 2.01

Gd 131.4 0.45 109.4 3.42

Catalysts 5Ni/CeZrM, Surfactant/Metal= 0.5

Mg 178.4 0.32 55.0 1.78

Ca 115.3 0.23 62.8 1.99

Gd 122.7 0.36 103.5 2.93

Catalyst 5Ni/CeZrGd 5Ni/CeZrCa 5Ni/CeZrMg

Ni Mass Fraction

(ICP-MS) %

2.82 2.82 2.82

66

Figure 4.2 (a) N2 Isotherms of fresh CeZrM supports and corresponding 5Ni/CeZrM

catalysts

5Ni/CeZrGd CeZrGd

CeZrMg 5Ni/CeZrMg

CeZrCa 5Ni/CeZrCa

67

4.1.3 N2 Physisorption on the Spent Catalysts

The spent catalysts were characterized to investigate the effect of reforming temperature

on the structural properties of the catalyst. The reduction in specific area and pore volume with

reforming temperature can be observed from the results of N2 physisorption. The isotherm

profiles shown in Figure 4.2 correspond again to “Type IV” category, the mesoporous structure

material. The concavity of the desorption loop reduced with the reaction temperature to yield a

lower surface area and pore volume for the spent catalyst.

The promoted Cerium-Zirconium support loaded with 5 wt% nickel (5%Ni/CZM)

showed a stable perfromance in the autothermal reforming of CG. Sintering of metallic atoms and

coke depositions are the main causes of catalyst deactivation. The addition zirconium to cerium

oxide support was proved to enhance the thermal stability of the binary Ce-Zr metal oxides and

thus prevents sintering (Ozawa, 1998). According to Luo et al. (2010), cerium also suppresses

sintering and reduces the selectivity of methane in the product gas for Ni-Co based catalysts.

Previous study in our group by Sukonket et al. (2011) revealed the enhanced thermal stability of

Ni/CeZrMx catalysts. The 5Ni/CeZrCa catalyst retained the mesoporous property and no sintering

effects were observed under a very high calcinations temperature (900°C) (Khan et al. 2011).

In the case of synthetic CG ATR reactions, the possible coke formation over the

surface of Ni/CeZrCa catalyst could be related to the high conversion (C-C cleavage rate)

of synthetic crude glycerol at high temperatures, the reason that would have caused the

observed reduction in surface area and porosity of the catalyst. In order to study this

possibility, TPO runs were performed on the spent catalysts to quantify carbon deposits

in case any presents.

68

Table 4.2 BET results of spent catalysts at different reaction temperatures

Spent “5Ni/CeZrCa” Catalyst

(S/C= 2.6 O/C= 0.125)

Reaction

Temperature

(°C)

BET SA

(m2/g)

Pore Volume

(cm3/g)

Pore Size

(A0)

PV. / SA

(10-9

m)

500 96.57 0.202 69.53 2.08

550 80.46 0.196 80.46 2.43

600 70.35 0.187 106.49 2.66

650 48.79 0.183 133.38 3.54

69

Figure 4.2 (b) N2 Isotherms of spent 5Ni/CeZrCa catalysts at different reaction

temperatures

Reaction T= 500 °C Reaction T= 550 °C

Reaction T= 600 °C Reaction T= 650 °C

70

4.1.4 Temperature Programmed Oxidation (TPO)

TPO runs were performed for the sake of quantifying the carbon that have

deposited on the surface of the spent catalysts during the autothermal reforming of

synthetic CG at different operating variables. Under the flow of oxidizing gas, the

temperature of the spent catalyst was increased to burn off the carbon and measure the

weight change. The weight loss in the range of 360-600°C was considered to be the

amount of coke formed during the ATR of synthetic CG experiments. The peaks of CO

and CO2 that correspond to coke gasification appeared in similar temperature ranges

established by [Le Minh et al. (1997); Nagaoka et al. (2000); Guo et al. (2007)].

It is familiar that ternary oxide mixture (CeZrM) has a high capability of reducing

coking affinity. Indeed, the redox nature of Ce-Zr metal oxides facilitates the gasification

of superficial carbon and minimizes coke formation over the surface of Ni/CeZr catalyst

(Kašpar & Fornasiero, 2002). A very low carbon deposition (~ 8 μmol of C/mg of

catalyst) was observed over Ni/CeZr catalyst by Romero-Sarria et al. (2008) in ethanol

steam reforming study. Moreover, a low carbon propensity factor of <40 mg C/g cat.h

was obtained by Sengupta (2011) and Oluku (2013) for a promoted (CeZrM) support

utilized in the steam reforming of oxygenated hydrocarbons and water gas shift reactions.

The rate of carbon deposition depends on several factors such as feed composition,

catalyst used, reforming temperature, steam-to-carbon (S/C) and oxygen-to-carbon (O/C)

ratios in the inlet feed. TPO profiles obtained at different ratios of O/C, S/C and reaction

temperatures are represented in Figures 4.6(a,b,c). Coke formation over catalyst surface

was expectedly found to decrease with O/C and S/C ratios and increase with the

reforming temperature.

71

As given in figure 4.3, with the increase in oxygen-to-carbon (O/C) ratio, coke

deposition rate decreased from 19.41 mg C/gcat.h at low O/C of 0.05 to almost zero at

high O/C of 0.2. Indeed, this is the one of the major advantages of introducing partial

oxidations alongside with steam reforming reactions. The oxygen will burn off the carbon

and prevents catalyst deactivation due to coke formation. This was discussed in many

studies [Hardiman et al. (2006); Swami & Abraham (2006); Deleplanque at al. (2010);

Wang et al. (2010)].

The molar ratio of steam-to-carbon (S/C) in the inlet feed was varied from 1.6 to

3.6. Figure 4.4 illustrates the variation of the rate of coke formation with S/C ratio; the

highest rate obtained was 10.38 mg C/gcat.h at S/C ratio of 1.6. It decreased to 1.61 mg

C/gcat.h at S/C ratio of 3.6. Actually, the addition of steam suppresses the radical

polymerization reactions that lead to coke formation (Trimm, 1997). However, high

steam-to-carbon ratio might also slow down the kinetics of the involved reactions and

reduce the thermal efficiency of the process. A favorable ratio should be adjusted to keep

the process in a stable and pragmatic area.

Coke deposition rate was found to increase with an increase in reforming

temperature as given in figure 4.5. According to the results given in figure 4.5

recommends a reforming temperature between 550 and 600°C. The maximum rate (6.37

mg C/gcat.h) was reported at the highest temperature (650°C). It has been revealed by

Albright et al, (1983) that the high reaction temperature will cause carbonaceous

intermediates to deposit on the surface of catalyst. In a recent study about the catalytic

steam reforming of glycerol, Chiodo et al. (2010) recommended not to exceed 650°C

reforming temperature in order not to produce encapsulated carbon that consequently

72

affects the stability of the catalyst. The highest rate of carbon deposition was observed at

650°C (4.2 mg C/gcat.h).

Figure 4.3 Coke deposition rates at different O/C ratios

Figure 4.4 Coke deposition rates at different S/C ratios

0

5

10

15

20

25

0.05 0.125 0.2

Co

ke

Fo

rma

tio

n (

mg

C/g

cat.

h)

O/C

Oxygen-to-Carbon

0

2

4

6

8

10

12

1.6 2.6 3.6

Co

ke

Fo

rma

tio

n (

mg

C/g

cat.

h)

S/C

Steam-to-Carbon

5Ni/CeZrCa

S/C= 2.6

Reaction T= 600

°C

5Ni/CeZrCa

O/C= 0.125

Reaction T= 600 °C

73

Figure 4.5 Coke formation rates at different reaction temperatures

0

1

2

3

4

5

6

7

500 550 600 650

Ca

rbo

n D

epo

siti

on

(m

g C

/gca

t.h

)

Reforming T (C)

Temperature

5Ni/CeZrCa

O/C= 0.125

S/C= 2.6

74

Figure 4.6 (a) TPO profiles of spent catalysts at different oxygen-to-carbon ratios

75

Figure 4.6 (b) TPO profiles of spent catalysts at different steam-to-carbon ratios

76

Figure 4.6 (c) TPO profiles of spent catalysts at different reaction temperatures

77

4.1.5 Temperature Programmed Reduction (TPR)

The target of establishing H2-TPR profiles was to study the reducibility of the

prepared metal oxide supports (CeZrM) along with the 5 wt% nickel impregnated

catalysts. The reduction of the metal oxides was detected by a thermal conductivity

detector (TCD). The signal peaks represent the hydrogen consumption by the oxygen

atoms attached to the metallic phase. The corresponding profiles are given in Figure 4.7

and the optimum reduction temperature was determined accordingly.

The H2-TPR profile obtained for CeZrCa was very similar to the one presented by

Ishioma (2013); the main peak was observed in the temperature range of 600-700°C and

another small one around 800°C. In the case of 5Ni/CeZaCa catalyst, two consecutive

peaks were observed in the ranges of 420-530°C and 550-630°C, and a third small one in

the range of 850-900°C. In agreement with the results obtained by Khan & Simirniotis

(2008), the earlier peaks can be ascribed to the surface reduction of nickel and cerium

oxide molecules, while the ones appearing beyond 800°C represent the bulk reduction, in

which oxygen needs first to diffuse and reach the surface where it reacts with the

reducing gas. As expected, no peaks were observed for pristine zirconium oxide. Indeed,

it is well-known that the reduction of ZrO2 molecules takes place at high temperatures

(>1000°C). The reducibility was calculated based on 1/Tmax expression, where Tmax

corresponds to the biggest peak in the pattern. In the case of magnesium promoter

element, a shift to the right was observed with the major peak appeared between 600 and

700°C, again with a similar surface and bulk reduction behavior. The reducibility of

gadolinium promoted catalyst was very similar to calcium one. In both cases the highest

reduction peak appeared around 500°C.

78

Figure 4.7 TPR profiles of the fresh supports and catalysts

-1

4

9

14

19

24

0 100 200 300 400 500 600 700 800 900

Sig

nal

(m

V)

Temperature (C)

Ca

Support

Catalyst

-1

4

9

14

0 100 200 300 400 500 600 700 800 900

Sig

nal

(m

V)

Temperature (C)

Mg

Support

Catalyst

-1

4

9

14

19

24

29

34

0 100 200 300 400 500 600 700 800 900

Sig

nal

(m

V)

Temperature (C)

Gd

Support

Catalyst

CeZrCa

5Ni/CeZrCa

CeZrMg

5Ni/CeZrMg

CeZrGd

5Ni/CeZrGd

79

4.1.6 X-Ray Diffraction (XRD)

The crystallinity of the developed supports and catalysts was investigated through

powder XRD analyses. Figure 4.8 displays the XRD patterns obtained for the prepared

CeZrM supports together with the 5wt% impregnated nickel catalysts. The green

spectrum at the middle is for nickel oxide (NiO2). The reason for presenting nickel oxide

spectrum in between supports and catalysts ones is to be able to pinpoint the existence or

absence of a new crystalline phase; and subsequently this will ensure or deny the

uniformity of active metal dispersion over the support surface.

Same diffraction spectra were obtained for the modified Ce-Zr supports prepared

using different promoter elements, and this confirms the fact that these elements does not

really affect the orientational order of the supporting material phases. The diffraction

patterns shown in Figure 4.8 revealed a cubic fluorite structure having an exact match

with pristine ceria patterns for the prepared supports and catalysts. The diffraction spectra

of supports and catalysts are identical in the case of calcium promoter. No peaks for NiO

were detected in the catalyst patterns, and this supports the significance of using the

surfactant-assisted method for impregnation in order to create a monolayer of the active

metal over the surface of the support. The presence of NiO crystalline phase in case of

magnesium promoter can be identified from the figure. At the same time, it is very hard

to be identified in case of gadolinium promoter. The observed difference in active metal

segregation upon changing the promoter element can be attributed to the dissimilarity of

interactions between NiO and the promoting atoms.

80

Figure 4.8 XRD patterns of the fresh supports and catalysts

0

500

1,000

1,500

0 20 40 60 80 100

In

ten

sity

5Ni/CeZrMg

CeZrMg

5Ni/CeZrGd

5Ni/CeZrCa

CeZrGd

CeZrCa

Nickel Oxide NiO

81

4.2 CATALYTIC ACTIVITY

Several chemicals similar to the ones existing in the crude glycerol generated

from biodiesel production were added separately to the synthetic CG feed in order to

investigate their individual effects. The promoter element at the catalyst level was then

varied to select a genuine one for the synthetic CG ATR process.

4.2.1 Effects of Synthetic Crude Glycerol Components

The effects of the different components of crude glycerol were investigated in a

packed bed tubular reactor at steam-to-carbon ratio of 3.6 and oxygen-to-carbon ratio of

0.125 using 5Ni/CeZrGd catalyst. Nickel was impregnated with the help of the wet

impregnation technique at atmospheric pressure. Similar to the composition of crude

glycerol generated at biodiesel plants, the synthetic crude glycerol was prepared to

contain free glycerol, methanol, soap, free fatty acids and ashes. These components have

been widely reported in the literature and confirmed by the collected samples from

Milligan Biofuels Inc. that operates in Saskatchewan-Canada. The individual effect of

each component was experimentally inspected by adding them separately to the liquid

feed at the beginning of the screening tests. Gadolinium was used as the promoter

element. At this stage, the main concern was to check the capability of the prepared

catalyst for handling this mixture of different chemicals without being attacked and

deactivated.

In general, a stable performance was observed for the Gd catalyst with different

effects of each component on the overall activity. The activity of the catalyst was

evaluated based on synthetic CG conversion, hydrogen selectivity and hydrogen yield

82

results. The first data point for pure glycerol reforming was adopted from (Sabri, 2013).

The effect of methanol addition on the conversion of glycerol mixture was very slight,

and a steady performance can be observed from the main effects plot Figure 4.10(a). The

methanol in the feed was reformed to produce hydrogen too. Metallic nickel catalyzes the

decomposition of methanol into hydrogen and carbon monoxide at the proper reaction

conditions (Kobayashi et al., 1981). Coming to the effect of methanol addition on

hydrogen production results, one can clearly observe the decrease in these values from

78.2 to 75.5 mol% and from 1.32 to 1.26 mol H2/ mol Cin for hydrogen selectivity and

hydrogen yield, respectively. This can be attributed to the low moles of hydrogen that can

be produced from methanol reforming reactions compared to glycerol ones. On the other

hand, a sharp decrease in the conversion can be observed due to soap addition that

contains high amount of 2-propanol. This behavior can be ascribed to the soap residues

accumulation in the reaction media and coke formation over the surface of the catalyst

due to 2-propanol reforming reactions. A steam reforming experiment was conducted by

Mizuno et al. (2003) on 2-propanol (Isopropyl alcohol IPA) for 360 min, the amount of

carbon deposits reported over Rh/CeO2 catalyst was 280 mg C/ gcat, and that was the

minimum measured amount compared to the ones deposited over SiO2, ZrO2 and TiO2

supports. The calculated rate of coke formation is 46.67 mg C/gcat.h, which is still a very

high rate. On the side of hydrogen production, the yield of hydrogen production

decreased from 1.26 to 1.17 mol H2/ mol Cin. This can be attributed to the reduction in C-

C cleavage rate due to the active sites blockage by carbon deposits, leading more to liquid

phase products and/or un-reacted feed in the outlet stream. Scanning more the plots of the

main effects in Figure 4.10, one can observe the promoting effect of adding ashes (NaCl

83

and KCl) to the feed. The conversion of synthetic CG and the hydrogen selectivity in the

gas product increased respectively from 78.1 to 84.1 mol% and from 72.3 to 78.5 mol%.

Potassium and sodium chlorides supplied by the ashes enhances the catalytic activity by

stimulating the catalytic oxidation of carbon deposits and thus providing more tolerance

for the catalyst against coke formation. Kumamoto (1995) confirmed the potency of NaCl

and KCl salts in oxidizing Hg0 due to their efficacy of charge-transfer. This enhanced

performance is in line with the low deposition rates observed by [Zhang et al. (2005);

Ibrahim & Idem (2008)] in the presence of promoter elements. The last addition was for

the oleic acid that did not show a significant effect on the catalytic activity. The

conversion was found to slightly increase from 84.1 to 84.5 mol% with a proportional

increase in the hydrogen yield from 1.31 to 1.34 mol H2/ mol Cin. This behavior could be

attributed to the similar acidic nature of the gadolinium support (CeZrGd) and the oleic

acid.

84

Figure 4.9 Evaluation of the individual effects of the different components of CG on the

ATR reactions of synthetic CG at 600°C, 1.0 atm, S/C= 3.6 and O/C= 0.125 using

5Ni/CeZrGd catalyst in a PBTR. (GM: Glycerol + Methanol)

40

50

60

70

80

90

100

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol

%)

TOS (h)

Synthetic CG Conversion

GM

GM+Soap

GM+Soap+NaCl&KCl

GM+Soap+NaCl&KCl+Oleic acid

40

50

60

70

80

90

0 1 2 3 4 5 6

Hy

dro

gen

Sel

ecti

vit

y (

mo

l %

)

TOS (h)

Hydrogen Selectivity

GM

GM+Soap

GM+Soap+NaCl&KCl

GM+Soap+NaCl&KCl+Oleic acid

0

0.4

0.8

1.2

1.6

0 1 2 3 4 5 6

Hy

dro

gen

Yie

ld (

mo

l H

2/m

ol

Cin

)

TOS (h)

Hydrogen Yield

GM

GM+Soap

GM+Soap+NaCl&KCl

GM+Soap+NaCl&KCl+Oleic acid

85

GM +

Soa

p + N

aCl&

KCl +

Oleic

Acid

GM +

Soa

p + N

aCl&

KCl

GM +

Soa

p

Glyc

erol

+ M

etha

nol

Glyc

erol

85

84

83

82

81

80

79

78

GM: Glycerol + Methanol

Me

an

Main Effects Plot for ConversionData Means

GM + S

oap +

NaCl&KC

l + O

leic

Acid

GM +

Soa

p +

NaCl&K

Cl

GM +

Soa

p

Glycerol + M

etha

nol

Glyc

erol

80

79

78

77

76

75

74

73

72

GM: Glycerol + Methanol

Me

an

Main Effects Plot for SelectivityData Means

(a)

(b)

86

GM +

Soa

p + N

aCl&

KCl +

Oleic

Acid

GM +

Soa

p + N

aCl&

KCl

GM +

Soa

p

Glycer

ol +

Met

hano

l

Glycer

ol

1.35

1.30

1.25

1.20

1.15

GM: Glycerol + Methanol

Me

an

Main Effects Plot for YieldData Means

GM +

Soa

p + N

aCl&

KCl +

Oleic

Acid

GM +

Soa

p +

NaCl&

KCl

GM +

Soa

p

Glycer

ol +

Met

hano

l

Glyc

erol

0.31

0.30

0.29

0.28

0.27

0.26

0.25

0.24

GM: Glycerol + Methanol

Me

an

Main Effects Plot for TOFData Means

Figure 4.10 Individual effects of the different components of CG on the performance of

ATR reactions (a) Main effects plot for synthetic CG Conversion [mol %] (b) Main effects

plot for hydrogen selectivity [mol %] (c) Main effects plot for hydrogen yield [mol H2/ mol

C] (d) Main effects plot for turnover frequency [1/s].

(c)

(d)

87

4.2.2 CATALYSTS SCREENING- PROMOTER EFFECT

The activities of the prepared mixed oxide supports CeZrM (M= Ca, Mg, Gd)

loaded with 5 wt% nickel catalyst were tested for the authothermal reforming of synthetic

crude glycerol in a packed bed tubular reactor at atmospheric pressure and 600°C.

Throughout these experiments, the steam-to-carbon and oxygen-to-carbon ratios were

maintained at 3.6 and 0.125, respectively. The three tested supports were prepared

according to surfactant assisted technique followed by a wet impregnation of the same

amount of nickel (5 wt%). The catalysts were then calcined at 650°C and then palletized

to 0.8 mm sized particles. A catalyst weight time of 158.2 gcat.min/mol C was applied.

The screening runs of synthetic CG ATR are given as conversion, hydrogen selectivity

and hydrogen yield results versus time-on-stream (TOS) in Figure 4.12.

According to the main effects results shown in Figure 4.13, the performance of

calcium promoted catalyst was very similar to gadolinium one with a very slight

distinction. The conversion obtained was 84.5 mol% with 5Ni/CeZrCa while it was

84.3% with 5Ni/CeZrGd. Same slight difference was obtained in the hydrogen selectivity

results with 79.7 and 79.6 mol% for the Ca and Gd catalysts, respectively. A more

observable deviation was obtained with the magnesium promoted catalyst with 78.8 and

75.3 mol% for conversion and hydrogen selectivity, respectively. The order of catalytic

activity arranged based on the impact of promoter elements is as follows; Ca > Gd > Mg.

The non-catalytic run was carried out at an identical set of operating conditions to

differentiate between the thermal and catalytic effect on the process. The absence of

active sites and the accumulation of carbon deposits inside the reactor during the non-

catalytic experiment could have contributed to the observed poor and unstable

88

performance. The thermal effect (T= 600°C) stimulated the steam reforming reactions

due to their endothermic nature.

89

Figure 4.11 Performance evaluation of 5Ni/CeZrM catalysts for synthetic CG ATR at S/C=

3.6, O/C= 0.125 and reaction T=600°C in a PBTR.

20

40

60

80

100

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol

%)

TOS (hr)

Synthetic CG Conversion

5Ni/CZCa 5Ni/CZGd 5Ni/CZMg Non-catalytic

10

30

50

70

90

0 1 2 3 4 5 6

Hy

dro

gen

Sel

ecti

vit

y (

mo

l%)

TOS (hr)

Hydrogen Selectivity

5Ni/CZCa 5Ni/CZGd 5Ni/CZMg Non-Catalytic

0.2

0.6

1

1.4

0 1 2 3 4 5 6

Hy

dro

gen

Yie

ld(m

ol

H2/m

ol

Cin

)

TOS (hr)

Hydrogen Yield

5Ni/CZCa

5Ni/CZGd

5Ni/CZMg

Non-Catalytic

90

Non-Catalytic5Ni/CeZrMg5Ni/CeZrGd5Ni/CeZrCa

85

80

75

70

65

60

55

50

Catalyst

Me

an

Main Effects Plot for ConversionData Means

Non-Catalytic5Ni/CeZrMg5Ni/CeZrGd5Ni/CeZrCa

80

70

60

50

40

Catalyst

Me

an

Main Effects Plot for Hydrogen SelectivityData Means

Non-Catalytic5Ni/CeZrMg5Ni/CeZrGd5Ni/CeZrCa

1.4

1.3

1.2

1.1

1.0

0.9

0.8

0.7

0.6

0.5

Catalyst

Me

an

Main Effects Plot for Hydrogen YieldData Means

5Ni/CeZrMg5Ni/CeZrGd5Ni/CeZrCa

0.30

0.29

0.28

0.27

0.26

Catalyst

Me

an

Main Effects Plot for Turnover FrequencyData Means

Figure 4.12 Effects of the developed catalysts on the ATR of synthetic CG; (a) Main effects

plot for synthetic CG Conversion [mol %] (b) Main effects plot for hydrogen selectivity

[mol %] (c) Main effects plot for hydrogen yield [mol H2/ mol C] (d) Main effects plot for

Turnover Frequency [1/s].

(a) (b)

(c) (d)

91

4.2.3 Structure-Activity Relationship

In general, the overall performance of a catalytic process is the outcome of

complex interactions among the several chemical and structural properties of the utilized

catalyst, alongside with the applied operating conditions. The enhanced performance of

the 5Ni/CeZrCa catalyst in the ATR of CG can be attributed to the reducibility, nickel

dispersion, PV/SA and hammet basicity properties. Figure 4.13 shows the activity-

structure main effects plot based on conversion and hydrogen selectivity results.

According to TPR results given in section 4.1.4, Ca and Gd promoted catalysts were

reduced at 487 and 495°C, respectively; while the reduction of Mg promoted one was

achieved at a higher temperature of 624°C. Consequently, the status of Ca promoted

catalyst can stay more active during the reforming process due to the ease with which

reduction can occur. Moreover, the percentage of active metal dispersion reported by

Sabri (2013) was 11.46% for the 5Ni/CeZrCa catalyst; however this was 3.54 and 0.44

for Gd and Mg promoted ones, respectively. This indicates that more active sites will be

provided by Ca promoted catalyst compared to Gd and Mg ones. The high density of

active sites with the Ca catalyst could reasonably justify the observed trend of activity.

The ratio of pore volume to surface area (PV/SA) could have also contributed to the

observed catalytic performance. The two catalysts (Ca and Gd) that showed the highest

conversion and selectivity results possess the biggest ratios of pore volume to surface

area (PV/SA). Furthermore, the activity was correlated with the surface acid-base

properties of the catalyst. The investigation of surface basicity as a function of basic

strength (H-) was done by Hammet & Deyrup (1932). The values of Ca and Mg Hammett

basicity functions were reported in literature by idem et al. (2012). According to Di

92

Cosimo et al. (1998), some reactions involved in ethanol conversion such as

dehydrogenation, condensation and dehydration requires strong basic sites to initiate

surface ethoxide formation. The authors revealed that, the catalytic performance of Mg-

Al hydrotalcites, with high density and strength of basic sites, presented a higher activity

than the pure MgO.

In order to study the contribution of each factor to the observed catalytic

performance, Pearson correlation coefficients were calculated along with the

corresponding P-values. Pearson coefficient ranged between -1 and +1, measures the

extent of linearity between two variables. The direction and the strength of this

relationship are determined by the sign and the magnitude of the coefficient, respectively.

The degree of confidence in this linear correlation is represented by P-value. Table 4.3

lists the coefficients and their corresponding P-values according to synthetic CG

conversion and hydrogen selectivity results. The relationship between reducibility and

activity exhibits a perfect linearity with a coefficient of 1.0. The corresponding P-values

of 0.017 and 0.007 are both less than α-level (0.05); this confirms that the correlation

between reducibility and activity is different from zero. Nickel dispersion, with Pearson

coefficient of 0.75, comes after reducibility as the most affecting factor on catalytic

activity. In this case, the existence of linear correlation cannot be affirmed since the

corresponding P-value is greater than 0.05. According to Pearson correlation analysis, the

significance order of catalyst structural properties on the overall performance is as

follows; Reducibility> Nickel Dispersion> PV/SA> Surface Basicity.

93

Table 4.3 Pearson correlations coefficients for structure-activity relationship

Based on Conversion Based on H2 Selectivity

Catalyst Property

Pearson

Correlation

Coefficient

P-Value

Pearson

Correlation

Coefficient

P-Value

Reducibility 1.00 0.017 1.00 0.007

Nickel Dispersion 0.75 0.460 0.78 0.435

PV/SA 0.56 0.621 0.53 0.645

Surface Basicity 0.54 0.636 0.57 0.611

94

2.08E-032.04E-031.60E-03

80.0

79.5

79.0

78.5

78.0

11.463.540.44

2.91E-091.92E-091.77E-09

80.0

79.5

79.0

78.5

78.0

2827

Reducibility (1/C)

Me

an

Ni Dispersion (%)

PV/SA (m) Hammett Basicity (H-)

Main Effects Plot for Hydrogen selectivityData Means

2.08E-032.04E-031.60E-03

84.0

83.5

83.0

82.5

11.463.540.44

2.91E-091.92E-091.77E-09

84.0

83.5

83.0

82.5

2827

Reducibility (1/C)

Me

an

Ni Dispersion (%)

PV/SA (m) Hammett Basicity (H-)

Main Effects Plot for ConversionData Means

(a)

(b)

95

2.08E-032.04E-031.60E-03

0.300

0.295

0.290

0.285

11.463.540.44

2.91E-091.92E-091.77E-09

0.300

0.295

0.290

0.285

2827

Reducibility (1/C)

Mean

Ni Dispersion (%)

PV/SA (m) Hammett Basicity (H-)

Main Effects Plot for TOFData Means

Figure 4.13 Activity-structure correlations obtained for 5Ni/CeZrM catalysts where M=

Mg, Gd, Ca; (a) Main effects plot for synthetic CG conversion [mol %] (b) Main effects plot

for hydrogen selectivity [mol %] (c) Main effects plot for Turnover frequency [1/s].

(c)

96

4. 3 PARAMETRIC STUDY

Several operating variables were varied for the sake of optimizing and better

controlling the process. The experimental work was conducted in a PBTR at atmospheric

pressure. The performance was evaluated based on conversion, hydrogen selectivity and

hydrogen yield results. The investigated parameters include calcination temperature,

reduction temperature, particle size, reaction temperature, steam-to-carbon ratio and

oxygen-to-carbon ratio in the inlet feed. The obtained results are discussed in the

following section.

4.3.1 Effect of Reduction Temperature

The catalysts screened in the foregoing experiments for ATR of synthetic CG

were activated at a temperature of 700°C. According to the results obtained from TPR

analyses shown in Figure 4.7, a lower reduction temperature could be sufficient to

activate the 5Ni/CeZrCa catalyst and avoid any negative impacts of high temperature on

its structure. For this reason, the reduction temperature of 5Ni/CeZrCa catalyst was

varied between 500-700°C. 0.25 grams of the catalyst was used in each run. The

activation of the catalyst was achieved in a PBTR by flowing 100 ml/min of

5%H2/balance N2 gas for 2 hours at the desired reduction temperature prior to start

feeding the reactants. The reaction temperature, S/C ratio and O/C ratio were maintained

at 600°C, 3.6 and 0.125, respectively. The observed performances at different reduction

temperatures are illustrated in Figure 4.14. Evidently, a reduction temperature of 500°C

was not sufficient enough to completely reduce the metallic phase (NiO and CeO2) of the

catalyst. The initial conversion corresponding to the first two hours (TOS) was around

97

70% for the catalyst activated at 500°C; however this was around 85% for the catalysts

activated at 600 and 700°C. Almost the same behavior was observed with hydrogen

selectivity and yield results. This can be attributed to the reduction of CeO2 into active

cerium atoms that usually requires a temperature between 550-600°C. The second peak

(~590°C) of 5Ni/CeZrCa TPR profile in Figure 4.7 corresponds to hydrogen consumption

due to cerium oxide activation. The catalyst was then partially activated at 500°C and

tends to be fully active after 2 hours (TOS), because the ATR of synthetic CG was

carried out at 600°C with hydrogen being produced inside the reformer. The catalysts

reduced initially at different temperatures showed similar catalytic activities after

reaching steady state with conversion and hydrogen yield results of ~83% and 1.3 mol

H2/mol Cin, respectively. Taking into account the overall conversion, selectivity and yield

results presented in Figure 4.15, a reduction temperature of 600°C is recommended to

obtain the highest catalytic activity of 5Ni/CeZrCa.

98

Figure 4.14 Effect of reduction temperatures on the activity of 5Ni/CeZrCa catalyst at S/C=

3.6, O/C= 0.125 and reaction T=600°C in a PBTR.

50

55

60

65

70

75

80

85

90

95

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol%

)

TOS (h)

Synthetic CG Conversion

T reduction 700 C

T reduction 600 C

T reduction 500 C

50

55

60

65

70

75

80

85

90

95

0 1 2 3 4 5 6

Hy

dro

gen

Sel

ecti

vit

y(m

ol%

)

TOS (h)

Hydrogen Selectivity

T reduction 700 C

T reduction 600 C

T reduction 500 C

0.4

0.6

0.8

1

1.2

1.4

1.6

0 1 2 3 4 5 6

Hy

dro

gen

Yie

ld (

mo

l H

2/m

ol

Ci n

)

TOS (h)

Hydrogen Yield

T reduction 700 C

T reduction 600 C

T reduction 500 C

99

700600500

84

83

82

81

80

79

78

Reduction Temperature

Me

an

Main Effects Plot for ConversionData Means

700600500

81

80

79

78

77

Reduction Temperature

Me

an

Main Effects Plot for Hydrogen SelectivityData Means

700600500

1.34

1.32

1.30

1.28

1.26

1.24

1.22

1.20

Reduction Temperature

Me

an

Main Effects Plot for Hydrogen YieldData Means

700600500

0.295

0.290

0.285

0.280

0.275

0.270

0.265

Reduction Temperature

Me

an

Main Effects Plot for TOFData Means

Figure 4.15 Performance variations with reduction temperature (a) Main effects plot for

synthetic CG conversion [mol%] (b) Main effects plot for hydrogen selectivity [mol %] (c)

Main effects plot for hydrogen yield [mol H2/mol Cin) (d) Main effects plot for turnover

frequency [1/s].

(a) (b)

(c)

(a)

(c) (d)

100

4.3.2 Effect of Calcination Temperature

In the early stages of this work, the catalysts utilized for synthetic CG ATR were

calcined at 650°C. But the results obtained from TGA analyses shown in Figure 4.1

revealed that such a high temperature is unneeded as all the adsorbed moisture, CTAB

and nitrates can be removed at a lower temperature. Exposing the catalyst to a high

temperature is unfavorable since it might cause sintering and thus reduction in the

catalytic activity. The calcination temperature was then varied between 550 and 650°C to

investigate the impact of thermal treatment on the activity of the catalyst. The screening

experiments were conducted in a PBTR at a reforming temperature of 600°C, S/C ratio of

3.6 and O/C ratio of 0.125. According to the results shown in Figures 4.16 and 4.17, the

premier catalytic performance for 5Ni/CeZrCa was obtained at a calcination temperature

of 550 °C. The conversion increased from 82.8% to 87.98% as the calcination

temperature decreased from 650 to 550°C. In addition, the yield of hydrogen increased

from 1.32 to 1.42 mol H2/mol Cin with this decrease in calcination temperature. These

observations could be attributed to the possible deformation in the structure of the

catalyst due to thermal effect. According to Sukonket et al. (2011), the porosity and

surface area of these materials decreases as the temperature increases. Besides that, nickel

particles might agglomerate at high temperatures leading to a dramatic decrease in the

number of active sites.

101

Figure 4.16 Effect of calcination temperatures on the activity of 5Ni/CeZrCa catalyst at

S/C= 3.6, O/C= 0.125 and reaction T=600°C in a PBTR.

20

30

40

50

60

70

80

90

100

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol

%)

TOS (h)

Synthetic CG Conversion

Calcination T= 650 C

Calcination T= 600 C

Calcination T= 550 C

20

30

40

50

60

70

80

90

100

0 1 2 3 4 5 6

hy

dro

gen

Sel

ecti

vit

y (

mo

l %

)

TOS (h)

Hydrogen Selectivity

Calcination T= 650 C

Calcination T= 600 C

Calcination T= 550 C

0.4

0.6

0.8

1

1.2

1.4

1.6

0 1 2 3 4 5 6

H2 Y

ield

(m

ol

H2/m

ol

Cin

)

TOS (h)

Hydrogen Yield

Calcination T= 650 C

Calcination T= 600 C

Calcination T= 550 C

102

650600550

88

87

86

85

84

83

Calcination T (C)

Me

an

Main Effects Plot for ConversionData Means

650600550

81.0

80.5

80.0

79.5

79.0

Calcination T (C)

Me

an

Main Effects Plot for Hydrogen SelectivityData Means

650600550

1.42

1.40

1.38

1.36

1.34

1.32

Calcination T (C)

Me

an

Main Effects Plot for Hydrogen YieldData Means

650600550

0.310

0.305

0.300

0.295

0.290

Calcination T (C)

Me

an

Main Effects Plot for TOFData Means

Figure 4.17 Performance variations with calcination temperatures (a) Main effects plot for

synthetic CG conversion [mol %] (b) Main effects plot for hydrogen selectivity [mol %] (c)

Main effects plot for hydrogen yield [mol H2/ mol Cin] (d) Main effects plot for Turnover

frequency [1/s].

(a)

(b)

(c)

(d) (c)

(a)

103

4.3.3 Effect of Steam-to-Carbon Ratio

The effect of steam-to-carbon ratio on the ATR reactions of synthetic CG was

investigated in a PBTR with 5Ni/CeZrCa catalyst. Reaction temperature was maintained

at 600°C. The molar ratio of steam to carbon in the inlet feed was varied between 1.6 and

3.6 at three different oxygen-to-carbon molar ratios. An excess steam is usually utilized

to promote steam reforming reactions, suppress methanation ones and minimize coke

formation over the catalyst. Similar to the results obtained by Dauenhauer et al. (2006),

the addition of steam enhanced the production of H2 and CO2 gases as given in the

product distribution plot of Figure 4.18. As expected, the increase of S/C ratio also

reduced methanation. This behavior can be attributed to the chemical reaction given in

(4.1). According to Le Chatelier’s priniciple, at low steam concentrations inside the

reaction media, hydrogen will tend to react with carbon monoxide molecules to produce

methane and steam and shift the system towards equilibrium.

CO + 3H2 CH4 + H2O (4.1)

The conversion of synthetic CG was observed to decrease from 78.9 to 73.4

mol% with an increase in S/C molar ratio from 1.6 to 3.6. Nonetheless, hydrogen

selectivity increased from 78.7 to 89.2 mol%. The decrease in conversion with S/C ratio

can be ascribed to the formation of carbon based unsaturated liquid phase species at high

steam concentrations such as ethanal, acetic acid, acetol, 1,1-ethanediol and others; and

thus less carbon based gaseous species and consequently lower conversion results. Note

that, conversion in this study was just based on the outlet carbon in the gas phase and the

carbon in the inlet feed. The turnover frequency decreased significantly from 0.44 to 0.28

104

s-1

. Indeed, steam was provided by diluting the synthetic CG mixture with a

predetermined amount of water based on S/C molar ratio. For this reason, the

concentration of biomass in the feed dropped with the increase in the molar ratio of S/C

and eventually the potency of hydrogen production per active site decreased. This setup

limitation affected the results of TOF variation with S/C, but not conversion, selectivity

or yield ones. The increasing trend of selectivity with the increase in S/C ratio is due to

the promoted steam reforming reactions at high steam concentrations. Similar sets of S/C

ratios were also tested at O/C ratios of 0.125 and 0.2 and given in the next section.

105

Figure 4.18 Effect of S/C ratio on synthetic CG ATR reactions using 5Ni/CeZrCa catalyst at

O/C= 0.05 and reaction T=600°C in a PBTR.

20

30

40

50

60

70

80

90

100

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol

%)

TOS (h)

Synthetic CG Conversion

S/C= 3.6

S/C= 2.6

S/C= 1.6

40

50

60

70

80

90

100

0 1 2 3 4 5 6

Hy

dro

gen

Sel

ecti

vit

y (

mo

l %

)

TOS (h)

Hydrogen Selectivity

S/C= 3.6

S/C= 2.6

S/C= 1.6

0

10

20

30

40

50

60

70

1.5 2 2.5 3 3.5 4

Pro

du

ct d

istr

ibu

tio

n (

mo

l %

)

S/C

Product distribution variation with S/C

H2

CO2

CH4

106

Figure 4.19 Performance variations with steam-to-carbon (S/C) ratio at O/C= 0.05 and

reaction T=600°C in a PBTR.

73

74

75

76

77

78

79

80

1.2 2.2 3.2

Co

nv

ersi

on

(m

ol

%)

S/C

Synthetic CG Conversion

1.2

1.24

1.28

1.32

1.2 2.2 3.2

H2

Yie

ld (

mo

l H

2/M

ol

Cin

)

S/C

Hydrogen Yield

78

80

82

84

86

88

90

1.2 2.2 3.2

H2

Sel

ecti

vit

y (

mo

l %

)

S/C

Hydrogen Selectivity

0.2

0.25

0.3

0.35

0.4

0.45

0.5

1.2 2.2 3.2

TO

F (

1/s

)

S/C

TOF

107

4.3.4 Effect of Oxygen-to-Carbon Ratio

The addition of oxygen to the process of synthetic CG ATR has three principal

effects: (1) improves the energy efficiency by supplying heat via oxidation reactions; (2)

enhances the stability by reducing carbon deposits over the catalyst; (3) moves the

process into a practical level by reducing the fixed and operating cost of the reformer.

However, O/C ratio is a very critical operating parameter that needs attentive adjustment,

since the presence of oxygen at high concentrations depletes the production of hydrogen.

The speedy kinetics of partial oxidations suppresses the reactions of steam reforming

resulting in a small unreacted fraction of the initial synthetic CG in the feed ready to react

with steam and generate hydrogen. In order to investigate the effect of oxygen-to-carbon

ratio on the catalytic ATR of synthetic CG, the ratio of O/C was varied in the range of

0.05-0.2 in a PBTR under atmospheric pressure. The performance was screened at

different S/C ratios with five O/C ratios tested at S/C= 2.6 as given in Figure 4.20. The

overall results of activity variations with O/C ratio at different steam concentrations are

presented in Figure 4.21. Taking an example of S/C at 3.6, the conversion of synthetic

CG increased from 73.5 to 88.4 mol% with an increase of O/C from 0.05 to 0.2; however,

hydrogen selectivity decreased from 89.2 to 54.3 mol%. This can be explained by the

increased flow of CO2 in the effluent gas stream generated via partial oxidation reactions,

concurrently with generating water due to hydrogen consumption by oxygen molecules.

In other words, increasing the ratio of O/C in the inlet feed increased the concentration of

CO2 in the outlet stream and thus the conversion; but at the same time, it reduced the

molar percentage of H2 molecules in the effluent gas and consequently lessened the yield

of production and the turnover frequency.

108

Figure 4.20 Effect of O/C ratio on synthetic CG ATR reactions using 5Ni/CeZrCa catalyst

at S/C= 2.6 and reaction T=600°C in a PBTR.

50

60

70

80

90

100

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol%

)

TOS (h)

Synthetic CG Conversion

O/C= 0.2 O/C= 0.16 O/C= 0.125 O/C= 0.1 O/C= 0.05

40

50

60

70

80

90

0 1 2 3 4 5 6

H2 S

elec

tiv

ity

(m

ol%

)

TOS (h)

Hydrogen Selectivity

O/C= 0.05

O/C= 0.1

O/C= 0.125

O/C= 0.16

O/C= 0.2

0

10

20

30

40

50

60

70

0.025 0.075 0.125 0.175 0.225

Pro

du

ct D

istr

ibu

tio

n (

mo

l %

)

O/C

Product distribution variation with O/C

H2

CO2

CH4

109

Figure 4.21 Performance variations with oxygen-to-carbon (O/C) ratio at 600°C and

W/FA0=127.4 gcat.min/ mol C using 5Ni/CeZrCa catalyst.

40

50

60

70

80

90

100

0.025 0.075 0.125 0.175 0.225

Hy

dro

gen

Sel

ecti

vit

y (

mo

l %

)

O/C

Hydrogen Selectivity

S/C= 1.6

S/C= 2.6

S/C= 3.6

60

70

80

90

100

0.025 0.075 0.125 0.175 0.225

Co

nv

ersi

on

(m

ol

%)

O/C

Synthetic CG Conversion

S/C= 1.6

S/C= 2.6

S/C= 3.6

0.9

1

1.1

1.2

1.3

1.4

0.025 0.075 0.125 0.175 0.225

H2 Y

ield

(m

ol

H2 /m

ol

Cin

)

O/C

Hydrogen Yield

S/C=1.6

S/C=2.6

S/C=3.6

0

0.1

0.2

0.3

0.4

0.5

0.025 0.075 0.125 0.175 0.225

TO

F (

1/s

)

O/C

TOF

S/C=1.6

S/C=2.6

S/C=3.6

110

4.3.5 Effect of Reaction Temperature

Several chemical reactions are anticipated to be involved in the ATR process of

synthetic CG. Some of these reactions are endothermic such as steam reforming and

methane dry reforming; and others are exothermic such as partial oxidation, methanation

and water gas shift reactions. Hence, operating temperature is a very critical parameter

that can significantly affect the process and drives the production toward the planned

target. According to the experimental results presented in Figure 4.22, the performance of

ATR process of synthetic CG was enhanced by increasing reaction temperature up to a

certain point (550-600°C); beyond this temperature, a steady behavior was observed. The

production of hydrogen increased with increasing the operating temperature due to the

endothermic nature of steam reforming reactions that posses the ability of generating

more H2 than the partial oxidation ones. Turnover frequency also increased from 0.079 to

about 0.17 s-1

with the same raise in temperature. The concentration of CO2 decreased

with increasing the operating temperature; however H2 percentage in the product gas

mixture was practically invariant beyond 575°C. This can be attributed to the promoted

reactions of dry methane reforming at high temperatures. In addition, methane formation

was suppressed by increasing the operating temperature up to ~575°C; beyond this value,

CH4 percentage in the outlet gas remained constant. This can be explained by the

exothermic nature of methanation reactions. The endothermicity of reaction (4.2)

explains the negative impact of operating temperature on the rate of coke formation

presented in Figure 4.5.

CH4 2H2 + C ΔH298k= 74.85 kJ/mol (4.2)

111

Figure 4.22 Performance variations with reaction temperature at S/C= 2.6, O/C= 0.125,

W/FA0 =127.4 gcat.min/ mol C using 5Ni/CeZrCa catalyst.

50

60

70

80

90

100

450 500 550 600 650

Co

nv

ersi

on

(m

ol

%)

Reaction T (°C)

Synthetic CG Conversion

40

50

60

70

80

450 500 550 600 650

Hy

dro

gen

Sel

ecti

vit

y (

mo

l %

)

Reaction T (°C)

Hydrogen Selectivity

0

0.05

0.1

0.15

0.2

450 500 550 600 650

TO

F (

1/s

)

Reaction T (°C)

TOF

0

0.3

0.6

0.9

1.2

1.5

450 500 550 600 650

H2 Y

ield

(m

ol

H2

/mo

l C

in)

Reaction T (°C)

Hydrogen Yield

0

20

40

60

450 500 550 600 650

Pro

du

ct d

istr

ibu

tio

n (

mo

l%)

Reaction T (°C)

Product distribution variation with reaction temperature

H2

CO2

CH4

112

4.4 STATISTICAL ANALYSIS

In order to investigate the causality and quantify the significance of each factor on

the overall performance, a statistical analysis was performed on the experimental data

collected from the ATR of synthetic CG in a PBTR setup. Minitab software was used to

develop main effects, interactions and Pareto charts. Regressions and model development

were achieved with the help of NLREG software.

4.4.1 Main Effects and Interactions

A statistical analysis was conducted to discover the effects of calcination

temperature, reduction temperature, particle size, promoter element, reaction temperature,

S/C and O/C ratios on the catalytic performance of 5Ni/CeZrCa for the ATR of Synthetic

CG. Figure 4.23 shows the main effects plots of these factors based on synthetic CG

conversion, hydrogen selectivity and turnover frequency criteria. The performance of the

process at different levels of several operating factors was shown in one plot with the

corresponding mean line. The presence of a main effect is determined by the relative

position of the response trend to the overall mean line. The larger the verticality of the

plotted response line, the larger the effect of this operating variable on the process

performance. Accordingly, conversion of synthetic CG is mostly affected by reaction

temperature and O/C ratio as shown in Figure 4.23(a). Hydrogen selectivity and turnover

frequency are strong functions of S/C and O/C ratios, with a considerable effect of

reaction and calination temperatures as shown in Figures 4.23(b) and (c), respectively.

These plots are beneficial to divide the parameters into two subcategories, namely, (1)

main factors, and (2) minor factors. Therefore, reaction temperature along with O/C and

113

S/C ratios are the most significant parameters in the ATR of synthetic CG process. The

horizontal trends for reduction temperature, particle size and promoter elements indicate

the minor impacts of these factors on the overall performance. The possible reasons for

these individual trends are presented in the parametric study. Similar statistical analysis

was conducted by Saha et al. (2014).

In general, the overall performance of a chemical process is a result of complex

interactions among the various operating parameters in the system. In order to assess the

extent of significance of these interrelations in the process of synthetic CG ATR,

interaction plots given in Figure 4.24 were established. Again, the degree of intersection

between the plotted lines demonstrates the strength of interaction between the two

corresponding factors. The more is the vertical crossing, the stronger is the interaction.

According to Figure 4.24(a), conversion of synthetic CG is strongly affected by the

interactions among reaction temperature, O/C and S/C ratios. This can be ascribed to the

controlling role of these variables over the kinetics of ATR reactions, and thus possessing

the ability of driving them into different pathways. The interactions of these main factors

with reduction temperature, calcinations temperature, particle size and promoter elements

are negligible. According to Figure 4.24(b), the variations in hydrogen selectivity could

be attributed to the high interactions between S/C and reaction temperature from one side,

and between these two factors and O/C from the other side. Turnover frequency is mostly

affected by the interactions between S/C and reaction temperature as shown in Figure

4.24 (c). In order to appropriately optimize, control and design the different units in this

process, it is very critical to understand these interactions and their resulting effects on

the overall performance.

114

In an attempt to evaluate the magnitude of both main and interaction effects on

the performance of synthetic CG ATR, Pareto charts were established based on

conversion, hydrogen selectivity and turnover frequency results. As illustrated in Figure

4.25(a), reaction temperature is the most affecting parameter on the conversion of

synthetic CG, followed by O/C, S/C and calcinations temperature. Besides that, S/C ratio

was found to be the most affecting factor on the hydrogen selectivity and O/C ratio on the

turnover frequency as presented in Figures 4.25(b) and (c), respectively. The parameters

with standardized effects less than the mean value of 1.341 are not significant

contributors to the overall performance. Pareto charts confirm also the minor effects of

reduction temperature, catalyst pellet size and promoter element on the overall

performance.

115

650600550

90

75

60

700600500 1.261.090.800.55

CaGdMg

90

75

60

650600550500 3.62 .61 .6

0 .2000.1250.050

90

75

60

Calcination T (C)

Me

an

Reduction T (C) Particle size (mm)

Promoter Element Reaction T (C) S/C

O/C

Main Effects Plot for ConversionData Means

650600550

80

65

50

700600500 1.261.090.800.55

CaGdMg

80

65

50

650600550500 3.62 .61 .6

0 .2000.1250.050

80

65

50

Calcination T (C)

Me

an

Reduction T (C) Particle size (mm)

Promoter Element Reaction T (C) S/C

O/C

Main Effects Plot for SelectivityData Means

650600550

0.3

0 .2

0 .1

700600500 1.261.090.800.55

CaGdMg

0.3

0 .2

0 .1

650600550500 3.62 .61 .6

0 .2000.1250.050

0.3

0 .2

0 .1

Calcination T (C)

Me

an

Reduction T (C) Particle size (mm)

Promoter Element Reaction T (C) S/C

O/C

Main Effects Plot for TOF (1/S)Data Means

Figure 4.23 Main effects plots (a) Synthetic CG Conversion (b) Hydrogen Selectivity (c)

Turnover Frequency.

(a)

(c)

(b)

116

Figure 4.24 Interactions plots (a) Synthetic CG Conversion (b) Hydrogen Selectivity (c)

Turnover Frequency.

700

600

500

1.26

1.09

0.80

0.55

0.00

208

0.00

204

0.00

160

650

600

550

500

3.6

2.6

1.6

0.20

0

0.12

5

0.05

0

90

75

6090

75

6090

75

6090

75

6090

75

6090

75

60

Calcination T (C)

Reduction T (C)

Particle size (mm)

Reducibility (1/C)

Reaction T (C)

S/C

O/C

550

600

650

T (C)

Calcination

500

600

700

T (C)

Reduction

0.55

0.80

1.09

1.26

size (mm)

Particle

0.00160

0.00204

0.00208

(1/C)

Reducibility

500

550

600

650

T (C)

Reaction

1.6

2.6

3.6

S/C

Interaction Plot for ConversionData Means

700

600

500

1.26

1.09

0.80

0.55

0.00

208

0.00

204

0.00

160

650

600

550

500

3.6

2.6

1.6

0.20

0

0.12

5

0.05

0

80

65

50

80

65

50

80

65

50

80

65

50

80

65

50

80

65

50

Calcination T (C)

Reduction T (C)

Particle size (mm)

Reducibility (1/C)

Reaction T (C)

S/C

O/C

550

600

650

T (C)

Calcination

500

600

700

T (C)

Reduction

0.55

0.80

1.09

1.26

size (mm)

Particle

0.00160

0.00204

0.00208

(1/C)

Reducibility

500

550

600

650

T (C)

Reaction

1.6

2.6

3.6

S/C

Interaction Plot for SelectivityData Means

700

600

500

1.26

1.09

0.80

0.55

0.00

208

0.00

204

0.00

160

650

600

550

500

3.6

2.6

1.6

0.20

0

0.12

5

0.05

0

0.3

0.2

0.1

0.3

0.2

0.1

0.3

0.2

0.1

0.3

0.2

0.1

0.3

0.2

0.1

0.3

0.2

0.1

Calcination T (C)

Reduction T (C)

Particle size (mm)

Reducibility (1/C)

Reaction T (C)

S/C

O/C

550

600

650

T (C)

Calcination

500

600

700

T (C)

Reduction

0.55

0.80

1.09

1.26

size (mm)

Particle

0.00160

0.00204

0.00208

(1/C)

Reducibility

500

550

600

650

T (C)

Reaction

1.6

2.6

3.6

S/C

Interaction Plot for TOFData Means

(a) (b)

(c)

117

Figure 4.25 Pareto Charts (a) Synthetic CG Conversion (b) Hydrogen Selectivity (c)

Turnover Frequency.

B

D

C

F

E

A

G

2.01.51.00.50.0

Te

rm

Standardized Effect

1.341

A C alcination T (C )

B Reduction T (C )

C Particle size (mm)

D Promoter Element

E Reaction T (C )

F S/C

G O /C

Factor Name

Pareto Chart of the Standardized Effects(response is TOF)

C

B

A

D

E

G

F

876543210

Te

rm

Standardized Effect

1.341

A C alcination T (C )

B Reduction T (C )

C Particle size (mm)

D Promoter Element

E Reaction T (C )

F S/C

G O /C

Factor Name

Pareto Chart of the Standardized Effects(response is Selectivity)

C

B

D

A

F

G

E

876543210

Te

rm

Standardized Effect

1.341

A C alcination T (C )

B Reduction T (C )

C Particle size (mm)

D Promoter Element

E Reaction T (C )

F S/C

G O /C

Factor Name

Pareto Chart of the Standardized Effects(response is Conversion)

(a) (b)

(c)

118

4.4.2 Model Development

A regression analysis was performed to correlate the catalytic activity with the

operating parameters for the ATR of synthetic CG. Reducibility and Pore volume to

surface area ratio were also included in as significant textural property in this case. A

non-linear model was developed using NLREG software. The experimental data used to

develop the model are given in Table 4.4 along with the experimental and predicted

conversions. The model expression is given in (4.3).

0.2 0.54× 4.1+0.55× 4.2+60.39× 0.3 0.75× 2+63.34× 0.3

where;

a: Calcination temperature in [°C]

b: Reduction temperature in [°C]

c: Pore volume-to-Surface area ratio in [cm3/m

2]

d: Reducibility in [°C -1

]

e: Reaction temperature in [°C]

f: Steam-to-Carbon molar ratio

g: Oxygen-to-Carbon molar ratio

(4.3)

119

The average absolute deviation (AAD) was then determined based on equation

(4.4). An excellent agreement between the experimental and predicted conversions of

synthetic CG was obtained with an AAD of 3.12%. The parity chart presented in Figure

4.26 shows the plot of predicted versus experimental data.

120

Table 4.4a Experimental data for model development

Run Calcination

T (°C)

Reduction

T (°C)

PV/SA

(cm3/m

2)

Reducibility

(C-1

)

Reaction

T (°C)

S/C O/C

1 650 500 0.0019 0.00208 600 3.6 0.125

2 650 600 0.0019 0.00208 600 3.6 0.125

3 650 700 0.0019 0.00208 600 3.6 0.125

4 650 600 0.0019 0.00208 600 3.6 0.125

5 650 600 0.0019 0.00208 600 3.6 0.125

6 650 600 0.0019 0.00208 600 3.6 0.125

7 650 600 0.0019 0.00208 600 3.6 0.125

8 550 600 0.0019 0.00208 600 3.6 0.125

9 600 600 0.0019 0.00208 600 3.6 0.125

10 650 600 0.0019 0.00208 600 3.6 0.125

11 650 600 0.0017 0.00160 600 3.6 0.125

12 650 600 0.0019 0.00208 600 3.6 0.125

13 650 600 0.0029 0.00204 600 3.6 0.125

14 650 600 0.0019 0.00208 500 2.6 0.125

15 650 600 0.0019 0.00208 550 2.6 0.125

16 650 600 0.0019 0.00208 600 2.6 0.125

17 650 600 0.0019 0.00208 600 2.6 0.125

18 650 600 0.0019 0.00208 600 1.6 0.125

19 650 600 0.0019 0.00208 600 2.6 0.125

20 650 600 0.0019 0.00208 600 3.6 0.125

21 650 600 0.0019 0.00208 600 2.6 0.05

22 650 600 0.0019 0.00208 650 2.6 0.125

23 650 600 0.0019 0.00208 600 2.6 0.25

121

Table 4.4b Experimental and predicted conversions

Run

Experimental

Conversion

(mol%)

Predicted

Conversion

(mol%)

AAD

(%)

1 78.6 79.5 1.06

2 83.4 81.8 1.91

3 82.9 83.8 1.04

4 81.3 81.7 0.64

5 84.3 81.7 2.96

6 83.7 81.8 2.31

7 83.6 81.7 2.15

8 87.9 86.9 1.16

9 85.9 84.5 1.64

10 82.8 81.7 1.22

11 79.6 81.7 2.79

12 84.1 81.7 2.77

13 83.8 81.8 2.49

14 60.4 64.5 6.89

15 81.6 75.8 7.05

16 90.5 86.4 4.44

17 90.5 86.4 4.52

18 92.9 89.5 3.59

19 90.5 86.4 4.52

20 86.9 81.7 5.93

21 76.4 78.2 2.41

22 91.1 96.4 5.90

23 92.3 94.3 2.12

Average 3.12

122

Figure 4.26 Parity plot of experimental versus predicted conversion

50

60

70

80

90

100

50 60 70 80 90 100

Pre

dic

ted

Con

ver

sion

(m

ol

%)

Experimental Conversion (mol %)

AAD= 3.12 %

123

4.5 KINETICS

A kinetic study was conducted in order to investigate the behaviour of the

corresponding ATR reactions of synthetic CG at both macroscopic and microscopic

levels. The developed rate expression demonstrates the speed of change of synthetic CG

concentration in terms of the main kinetic factors applied; reaction temperature and

concentrations of reactants.

4.5.1 Heat and Mass Transport Limitations

Intrinsic kinetic data can only be collected after eliminating the effects of heat and

mass transfer resistances on the rate of the studied reactions. For this reason, a kinetic

region free of heat and mass transfer limitations was experimentally established prior to

kinetic data collection.

4.5.1.1 Effect of Pellets Size

The size of catalyst pellets were varied in the range of 0.55-1.26 mm. The

experiments were performed in a PBTR at atmospheric pressure using 5Ni/CeZrCa

catalyst. An identical set of operating conditions were applied: Treaction=600°C; S/C= 3.6;

O/C= 0.125 and W/FA0= 15.82 gcat.min/mol C. In order to ensure a plug flow behavior

inside the PBTR, the ratios of bed length to particle diameter (l/dp) and reactor diameter

to particle diameter (d/dp) were maintained at 56.3 and 15.9, respectively, after selecting

the optimum particle diameter. The variation of process performance with the size of

pellets is shown in Figure 4.27. It can be observed from the figure that synthetic CG

conversion, hydrogen selectivity and hydrogen yield results are steady within the tested

range. A conversion of around 83% and a hydrogen selectivity of around 79% were

124

obtained with invariant behavior with the change in particle size. Therefore, any particle

diameter between 0.55 and 1.26 mm is appropriate to avoid heat and mass transport

effects and carry out the kinetic study on synthetic CG ATR process. These criteria:

d/dp≥10 and l/dp≥50 must be satisfied in a PBTR to have a plug flow conditions and

avoid back mixing and channeling [Froment & Bischof (1990); Idem & Bakhshi (1996)].

Accordingly, a particle diameter of 0.8 mm was selected.

125

1.261.090.800.55

100

80

60

40

20

0

Particle size (mm)

Co

nv

ersio

n (

mo

l %

)

Synthetic CG Conversion

1.261.090.800.55

100

80

60

40

20

0

Particle size (mm)

Hy

dro

ge

n S

ele

ctiv

ity

(m

ol %

)

Hydrogen Selectivity (mol %)

1.261.090.800.55

2.0

1.5

1.0

0.5

0.0

Particle size (mm)

Yie

ld (

mo

l H

2/

mo

l C

)

Hydrogen Yield

Figure 4.27 Effects of catalyst pellets size on the activity (a) Synthetic CG Conversion (b)

Hydrogen Selectivity (c) Hydrogen Yield

(a) (b)

(c)

126

4.5.1.2 Effect of Flow Rate

Besides the size of catalyst particles, the inlet flow rate was also varied in the

range of 0.15-0.25 ml/min in order to ensure the absence of heat and mass transfer

limitations on the rates of synthetic CG ATR reactions. The reactions were carried out at

a lower S/C ratio than the 3.6 ratio applied with the particle size experiments. The reason

was to provide higher carbon flow rates and make sure the rate of reaction is not affected

by the dilute stream. The reactions were carried out at 600°C temperature and

atmospheric pressure in the same PBTR. The ratios of S/C and O/C were maintained at

2.6 and 0.125, respectively. A fixed amount of 0.25 grams of 5Ni/CeZrCa (dp= 0.8 mm)

catalyst was employed in each run. The overall performance of the synthetic CG ATR

reactions is presented in Figure 4.28. A conversion of around 90% and a hydrogen

selectivity of around 70% were obtained within the tested range of flow rates.

Consequently, the rate of synthetic CG ATR reactions is invariant to feed flow rate in the

range of 0.15-0.25 ml/min, and any rate within this region is adequate to be utilized for

intrinsic kinetic data collection. Tsipouriari & Verykios (2001) followed a similar

experimental procedure to conduct their kinetic study on CO2 reforming of methane using

Ni/La2O3 catalyst.

127

0.003260.002610.00196

100

80

60

40

20

0

Inlet Flow (mol C/min)

Co

nv

ersio

n (

mo

l %

) Synthetic CG Conversion

0.003260.002610.00196

100

80

60

40

20

0

Inlet Flow (mol C/min)

Hy

dro

ge

n S

ele

ctiv

ity

(m

ol %

)

Hydrogen Selectivity

0.003260.002610.00196

2.0

1.5

1.0

0.5

0.0

Inlet Flow (mol C/min)

Yie

ld (

mo

l H

2/

mo

l C

)

Hydrogen Yield

Figure 4.28 Effect of feed molar flow rate on the activity (a) Synthetic CG Conversion (b)

Hydrogen Selectivity (c) Hydrogen Yield

(a) (b)

(c)

128

4.5.2 Extended TOS Stability

An extended period experiment was performed for the ATR of synthetic CG to

check the stability of 5Ni/CeZrCa catalyst. The results of conversion, hydrogen

selectivity and turnover frequency are given Figure 4.29. A stable performance was

observed with an average conversion of ~ 85 mol%; hydrogen selectivity of ~ 73% and

turnover frequency of ~0.14 s-1

under the prevailing conditions. The activity of the

catalyst was observed to ascend with time due probably to the accumulation of NaCl and

KCl salts in the reaction media, thus suppressing coke formation over the surface of

catalyst. This is clear from the increasing trends of conversion, hydrogen selectivity and

turnover frequency presented in Figures 4.29. I-control charts were then established to

assess whether the process is under control or not. The center line represents the overall

average of the process along with two control limit . As seen from Figure 4.30(a)

and (b), the first point of conversion results was the only observation that failed to pass

the control test, since the corresponding value was less than . This was expected

since it takes more than an hour for the system to reach steady state. Besides this point,

the last point from turnover frequency results was also beyond the control limits, most

likely due to the promoted performance of catalyst and/or the refill of the syringe pump

after 10 hours. Unlike the conventional catalysts and their deactivation with time, I-charts

illustrate the enhanced activity of 5Ni/CeZrCa catalyst for the ATR of synthetic CG

process.

129

Figure 4.29 Long term stability run for synthetic CG ATR over 5Ni/CeZrCa catalyst at

S/C= 3.6; O/C= 0.125; reaction T= 600°C and W/FA0= 2.64 gcat. h/mol C.

0

20

40

60

80

100

0 2 4 6 8 10 12 14 16

Co

nv

ersi

on

(m

ol%

)

TOS (h)

Synthetic CG Conversion

0

20

40

60

80

100

0 2 4 6 8 10 12 14 16

H2 S

elec

tiv

ity

(m

ol%

)

TOS (h)

Hydrogen Selectivity

0

0.05

0.1

0.15

0.2

0 2 4 6 8 10 12 14 16

TO

F (

1/s

)

TOS (h)

Turnover Frequency

130

1413121110987654321

95

90

85

80

75

70

Observation

Ind

ivid

ua

l V

alu

e _X=85.28

UCL=93.66

LCL=76.91

1

I Chart of Conversion

1413121110987654321

0.16

0.15

0.14

0.13

0.12

0.11

Observation

Ind

ivid

ua

l V

alu

e

_X=0.13923

UCL=0.15159

LCL=0.12688

1

1

I Chart of Turover Frequency

Figure 4.30 I-Control charts of conversion (a) and turnover frequency (b).

131

4.5.3 Mass Distribution

A material balance was conducted around the reformer in order to investigate the

mass distribution in the system. The law of mass conservation allows us to quantify the

carbon deposited over the walls of the PBTR and study the compatibility of the

experimental measurements with the predicted ones. The analyses of carbon in the

system presented in Figure 4.31 were performed as follows:

• Carbon in the gas phase was analyzed in ITC labs using an online GC/TCD (Agilent

6890) equipped with Hayesep Q and Molecular Sieve A columns.

• Carbon deposited over the catalyst was analyzed in ITC labs using TGA-50

(Shimadzu) instrument via TPO analysis.

• Carbon in the liquid phase was analyzed by Loring Laboratories LTD. (Calgary, AB)

according to ASTM D5373 standard method.

The law of mass conservation states that:

(4.5)

Therefore;

(4.6)

The inlet flow of carbon was determined based on the carbon content of the various

chemicals used in synthetic CG preparation.

132

The difference between the measured carbon in the outlet streams and the inlet carbon in

the feed is the carbon deposited over walls of the reactor, besides the possible errors in

measurements.

The distribution of carbon in the different streams was analyzed as mentioned before;

Therefore;

0.041

The carbon in the liquid phase was measured by a total carbon analysis (TOC)

given in Appendix D. The carbon deposited over the walls of the reactor was found as the

difference between inlet carbon and the carbon analyzed in the liquid, gas and over the

catalyst surface. As can be seen from Table 4.5, the majority of carbon goes into the gas

phase as carbon dioxide and methane gases; however the amount of carbon deposited

over the catalyst was very low. This justifies the stable and enhanced performance of

5Ni/CeZrCa catalyst.

133

Table 4.5 Carbon distribution in the system (basis: 1 hour of operations)

Stream Mass of Carbon (g) Carbon Percentage (wt%)

Effluent Gas 1.32 94.06

Liquid Condensate 0.042 2.99

Catalyst surface 0.00024 0.017

Reactor walls 0.041 2.93

134

Figure 4.31 Carbon balance around the reformer

135

4.5.4 Results of Kinetics Experiments

The reaction kinetics of synthetic CG autothermal reforming over 5Ni/CeZrCa

catalyst was investigated in a PBTR at atmospheric pressure. Four kinetic controlled

factors were varied, (1) reaction temperature, (2) weight space velocity W/FA0, (3) steam-

to-carbon ratio, and (4) oxygen-to-carbon ratio. Four operating temperatures of 500, 550,

600 and 650°C were applied. Weight space time (W/FA0) was varied between 0-127.4

gcat.min/ mol C at six different levels. This was achieved by employing different catalyst

weights of 0, 0.05, 0.1, 0.15, 0.2 and 0.25 grams. Steam-to-carbon ratio was varied

between 1.6-3.6 at three different O/C ratios of 0.05, 0.125 and 0.2. The variations of

conversion results at different S/C and O/C were presented above in Figure 4.21. The

variations of conversion with respect to weight space time (W/FA0) at different reaction

temperatures are given in Figure 4.32.

It is clear that the conversion of synthetic CG increased with increasing O/C ratio,

reaction temperature and weight space time (W/FA0). However, it decreased with

increasing the S/C ratio in the inlet feed. Figure 4.33 shows the variations of experimental

conversion versus W/FA0 at different reaction temperatures. The integral method of

kinetic analysis discussed by [Froment et al. (1990); Fogler (2011)] was then applied to

develop a rate expression for the ATR reactions of synthetic CG.

136

Figure 4.32 (a) Conversion variations with W/FA0 at different reaction temperatures

(500 and 550°C)

0

20

40

60

80

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol

%)

TOS (h)

T= 500 °C

W/FA0=127.4 gcat.min/mol C

W/FA0=101.9 gcat.min/mol C

W/FA0=76.4 gcat.min/mol C

W/FA0=50.9 gcat.min/mol C

W/FA0=12.7 gcat.min/mol C

W/FA0=0.0 gcat.min/mol C

0

20

40

60

80

100

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol%

)

TOS (h)

T= 550 °C

W/FA0= 127.4 gcat.min/mol C

W/FA0= 101.9 gcat.min/mol C

W/FA0= 76.4 gcat.min/mol C

W/FA0= 50.9 gcat.min/mol C

W/FA0= 12.7 gcat.min/mol C

W/FA0= 0.0 gcat.min/mol C

137

Figure 4.32 (b) Conversion variations with W/FA0 at different reaction temperatures

(600 and 650°C)

0

20

40

60

80

100

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol%

)

TOS (h)

T= 600 °C

W/FA0= 127.4 gcat.min/mol C

W/FA0= 101.9 gcat.min/mol C

W/FA0= 76.4 gcat.min/mol C

W/FA0= 50.9 gcat.min/mol C

W/FA0= 12.7 gcat.min/mol C

W/FA0= 0.0 gcat.min/mol C

0

20

40

60

80

100

0 1 2 3 4 5 6

Co

nv

ersi

on

(m

ol

%)

TOS (h)

T= 650 °C

W/FA0= 127.4 gcat.min/mol C

W/FA0= 101.9 gcat.min/mol C

W/FA0= 76.4 gcat.min/mol C

W/FA0= 50.9 gcat.min/mol C

W/FA0= 12.7 gcat.min/mol C

W/FA0= 0.0 gcat.min/mol C

138

Figure 4.32 (c) Conversion of synthetic CG versus W/FA0

y = -0.0014x2 + 0.343x + 70.669

R² = 0.9989

y = -0.0014x2 + 0.3966x + 62.114

R² = 0.9917

y = -0.0017x2 + 0.4704x + 48.85

R² = 0.9967

y = -0.0005x2 + 0.2135x + 41.352

R² = 0.9862

30

40

50

60

70

80

90

100

0 20 40 60 80 100 120 140

Co

nv

ersi

on

(m

ol%

)

W/FA0 (gcat.min/mol C)

T= 650 C

T= 600 C

T= 550 C

T= 500 C

139

4.5.5 Rate Model

A power law model was proposed for the rate of synthetic CG ATR reactions.

This kind of rate expressions is widely applied in the field of reaction engineering and

can accurately predict the rate values. Nonetheless, generating a mechanistic model was

always preferred due to its accountability for the pathways followed by the corresponding

chemical reactions. Discovering the true reaction mechanism gives a deep understanding

of the process, and thus improves the efficiency of designing and controlling tasks.

However, it is incredibly hard to determine a reaction mechanism in the case of crude

hydrocarbon reforming, because according to Parmar et al. (2010), such systems include

thousands of different reactions of hydrogenation/dehydrogenation, cracking,

isomerization, hydrogenolysis and hydrocyclisation. In our study about the ATR of

synthetic CG, several chemical species were involved, namely, (1) free glycerol, (2)

methanol, (3) 2-propanol, (4) oxygen, and (5) steam. These compounds and others such

as NaCl, KCl, oleic acid and soap, were all present in the inlet feed to the reformer.

Accordingly, an overall reaction given in (4.7) was suggested for the ATR of synthetic

CG and a power law expression was developed. In this case of the synthetic feed, coming

up with a mechanistic model for reaction kinetics was very hard due to the complexity of

reactions and large number of reactants. For the sake of simplicity, a power law model

was then proposed.

(4.7)

140

Where; the stoichiometric coefficients (a,b,c & d) are strong functions of operating

conditions. Tables 4.6 lists the corresponding values of a,b,c and d for the kinetic runs

based on the experimental observations at different conditions.

141

Table 4.6 Reaction stoichiometries at different operating conditions

Run S/C O/C T (K)

W/FA0

(gcat.min/

mol C)

a b c d

1 2.6 0.125 773 12.71 6.5 0.625 3.25 5.75

2 2.6 0.125 773 50.97 6.5 0.625 3.25 5.75

3 2.6 0.125 773 76.47 6.5 0.625 3.25 5.75

4 2.6 0.125 773 101.94 6.5 0.625 3.25 5.75

5 2.6 0.125 823 12.71 6.5 0.625 3.25 5.75

6 2.6 0.125 823 50.97 6.5 0.625 3.25 5.75

7 2.6 0.125 823 76.47 6.5 0.625 3.25 5.75

8 2.6 0.125 823 101.94 6.5 0.625 3.25 5.75

9 2.6 0.125 873 12.71 6.5 0.625 3.25 5.75

10 2.6 0.125 873 50.97 6.5 0.625 3.25 5.75

11 2.6 0.125 873 76.47 6.5 0.625 3.25 5.75

12 2.6 0.125 873 101.94 6.5 0.625 3.25 5.75

13 2.6 0.125 923 12.71 6.5 0.625 3.25 5.75

14 2.6 0.125 923 50.97 6.5 0.625 3.25 5.75

15 2.6 0.125 923 76.47 6.5 0.625 3.25 5.75

16 2.6 0.125 923 101.94 6.5 0.625 3.25 5.75

17 1.6 0.05 873 96.86 4 0.25 4 2.5

18 2.6 0.05 873 127.42 6.5 0.25 4 5

19 3.6 0.05 873 158.23 9 0.25 4 7.5

20 1.6 0.125 873 96.86 4 0.625 3.25 3.25

21 2.6 0.125 873 127.42 6.5 0.625 3.25 5.75

22 3.6 0.125 873 158.23 9 0.625 3.25 8.25

23 1.6 0.2 873 96.86 4 1 2.5 4

24 2.6 0.2 873 127.42 6.5 1 2.5 6.5

25 3.6 0.2 873 158.23 9 1 2.5 9

142

An empirical, reversible power law model can then be written as:

(4.8)

Here, A= synthetic CG; B= steam; C= oxygen;

rA= rate of reaction with respect to synthetic CG, mol C gcat-1

min-1

k0: frequency or collision factor, mol C gcat-1

min-1

atm-(x+y+z)

EA: activation energy, kJ mol-1

T: reaction temperature, K

R: molar gas constant, 8.314 J mol-1

K-1

PA: partial pressure of Synthetic CG, atm

PB: partial pressure of steam, atm

PC: partial pressure of oxygen, atm

x: reaction order with respect to A

y: reaction order with respect to B

z: reaction order with respect to C

143

4.5.6 Estimation and Validation of Model Parameters

Experimental conversion data were used to estimate the parameters of the power

law model. In line with a study by Parmar et al. (2010), an integral analysis was

performed to obtain the outlet flow of reactants based on the minimization of squared

sum of residuals (SSE). The change of synthetic crude glycerol flow along the height of

PBTR can be written as:

(4.9)

Where;

FA: flow rate of synthetic CG, mol min-1

W: weight of catalyst, g

ρ: density of catalyst, g ml-1

A´: cross sectional area, cm2

z: axial direction inside the PBTR

Therefore;

(4.10)

Where;

: outlet flow of synthetic CG for the ith

run, mol min-1

: inlet flow of synthetic CG for the ith

run, mol min-1

144

The regression analysis was done using PLKA software developed by our group.

A modified genetic algorithm was applied to get to the best estimates of parameters based

on the least deviation between experimental and predicted outlet flow of synthetic CG.

Then, the predicted conversion for each run was obtained based on this formula:

Where; (i) always corresponds to the number of the specific run.

After that, the individual error of each experiment was calculated as follows:

The average absolute deviation is given by:

The results of regression are given in Tables 4.7 and 4.8. An excellent agreement

between the experimental and predicted conversions was obtained with 5.2% average

absolute deviation (AAD).

145

Table 4.7 Estimation results of model parameters

Parameter Value

K0 2.09 x 1011

[mol C gcat-1

min-1

atm-3.46

]

EA 9.37 x 104

[kJ mol-1

]

x 1.05

y 0.56

z 1.85

AAD 5.2%

Figure 4.33 Parity plot of predicted data versus experimental data

146

Table 4.8 Experimental Kinetic data with the predicted conversion results

Run T

(K)

W/FA0

(gcat.min/

mol C)

Psynthetic

CG *102

(atm)

P

steam

(atm)

Poxygen

*102

(atm)

Pnitrogen

(atm)

Experimental

Conversion

(mol%)

Predicted

Conversion

(mol%)

AAD

(%)

1 773 12.71 9.48 0.61 5.87 0.23 45.25 38.00 16.11

2 773 50.97 9.48 0.61 5.87 0.23 50.15 60.12 19.90

3 773 76.47 9.48 0.61 5.87 0.23 55.34 65.71 18.75

4 773 101.94 9.48 0.61 5.87 0.23 56.99 69.34 21.66

5 823 12.71 9.48 0.61 5.87 0.23 55.65 52.53 5.60

6 823 50.97 9.48 0.61 5.87 0.23 68.83 71.61 4.03

7 823 76.47 9.48 0.61 5.87 0.23 74.09 75.96 2.53

8 823 101.94 9.48 0.61 5.87 0.23 79.38 78.70 0.87

9 873 12.71 9.48 0.61 5.87 0.23 65.03 64.08 1.40

10 873 50.97 9.48 0.61 5.87 0.23 79.15 79.51 0.40

11 873 76.47 9.48 0.61 5.87 0.23 84.77 82.83 2.32

12 873 101.94 9.48 0.61 5.87 0.23 88.46 84.48 4.54

13 923 12.71 9.48 0.61 5.87 0.23 74.85 72.70 2.93

14 923 50.97 9.48 0.61 5.87 0.23 84.29 84.95 0.78

15 923 76.47 9.48 0.61 5.87 0.23 88.82 100 12.61

16 923 101.94 9.48 0.61 5.87 0.23 90.16 91.23 1.14

17 873 96.86 16.11 0.64 3.99 0.16 78.93 78.63 0.34

18 873 127.42 11.52 0.74 2.85 0.11 76.43 76.21 0.24

19 873 158.23 8.95 0.79 2.22 0.08 73.45 73.42 0.10

20 873 96.86 12.39 0.49 7.68 0.30 92.45 88.72 4.08

21 873 127.42 9.48 0.61 5.87 0.23 90.54 86.50 4.42

22 873 158.23 7.67 0.68 4.59 0.19 86.94 85.24 1.91

23 873 96.86 10.07 0.39 9.99 0.39 94.56 94.39 0.22

24 873 127.42 8.07 0.52 7.99 0.32 92.32 89.84 2.66

25 873 158.23 6.72 0.60 6.66 0.26 88.42 88.82 0.48

Average 5.2

147

Finally, it is clear that the predicted conversions match extremely well with the

experimental values. The developed rate expression is valid at atmospheric pressure and

in a temperature range of 773-923 K; S/C range of 1.6-3.6; O/C range of 0.05-0.2; and

W/FA0 range of 0-158.23 gcat.min/ mol C. The final rate model can be written as:

148

CHAPTER 5

CONCLUSIONS AND RECOMMENDATIONS

5.1 CONCLUCIONS

A set of ternary metal oxide catalysts having a nominal composition of

5Ni/CeZrM (where M= Ca, Gd, Mg) was prepared, characterized and tested for the

autothermal reforming of synthetic CG. The catalysts were prepared using the surfactant

assisted method and nickel was impregnated according to the wet impregnation

technique. The catalyst containing calcium promoter element exhibited the highest

activity with a conversion of 84.5% and hydrogen selectivity of 79.7%. Synthetic CG

composed of free glycerol, methanol, soap, FFA, water and ashes was utilized in this

study. The individual effects of these components were investigated by adding them

separately to the feed mixture. Ashes composed of NaCl and KCl, showed a positive

effect on the catalytic performance.

In order to investigate the relationship between the observed activity and the

structure of the catalyst, several characterization techniques were applied on the prepared

supports and catalysts. These techniques include TGA, TPR, TPO, N2 physisorption,

XRD and ICP-MS analyses. The results of structure activity relationship (SAR) revealed

that the catalytic activity in the case of synthetic CG ATR process is mostly affected by

the reducibility and Ni dispersion properties of the catalyst.

Several variables at the level of catalyst preparation and operating conditions

were varied in order to optimize the process. Calcination temperature was varied in a

range of 550-650°C, reduction temperature was varied in the range of 500-700°C,

149

catalyst pellets size was varied in the range of 0.55-1.26 mm, volumetric flow rate was

varied in the range of 0.15-0.25 ml/min, reaction temperature was varied in the range of

500-650°C and, S/C and O/C ratios were varied in the ranges of 1.6-3.6 and 0.05-2.0,

respectively. All screening experiments were conducted in a lab scale PBTR at

atmospheric pressure. An increase in catalytic activity was observed with a decrease in

calcination temperature from 650 to 550°C. A temperature of 600 °C was enough to

activate the catalyst prior to screen its performance. Thus calcination and reduction

temperatures of 550 and 600°C were selected for further parametric studies, respectively.

In addition, S/C ratio was revealed to have a significant effect on process

performance. At a constant O/C ratio of 0.05, hydrogen selectivity increased from 78.7 to

89.2 mol% with an increase in S/C from 1.6 to 3.6; however synthetic CG conversion

decreased from 78.9 to 73.4 mol%. On the other hand, increasing O/C ratio from 0.05 to

0.2 showed a negative effect on hydrogen selectivity and turnover frequency that

decreased from 89.2 to 54.3 mol% and from 0.29 to 0.1 s-1

, respectively, at a fixed S/C of

3.6. Coke formation over catalyst surface was observed to become negligible at S/C and

O/C ratios greater than 2.6 and 0.125, respectively. Again, hydrogen selectivity, turnover

frequency and synthetic CG conversion increased with an increase in reaction

temperature from 550 to 650°C, but this increase was completely insignificant with a

considerable amount of carbon depositions at reaction temperatures beyond 600°C.

Therefore, the optimum reaction temperature, S/C and O/C ratios selected based on

conversion, hydrogen selectivity, turnover frequency and coke formation criteria were

550°C, 2.6 and 0.125, respectively.

150

A statistical analysis was then conducted to investigate the main effects and

interactions among the different operating parameters. The results revealed that reaction

temperature is the most affecting parameter on the conversion of synthetic CG, followed

by O/C, S/C and calcinations temperature. S/C and O/C ratios were the most effective in

terms of hydrogen selectivity. O/C and calcination temperature were the most effective

based on turnover frequency results. Overall, the results from analysis of variances

(ANOVA) showed that reaction temperature, S/C, O/C and calcination temperatures are

the most significant parameters in the process. A nonlinear mathematical model was then

developed to express the conversion of synthetic CG in terms of the different operating

parameters.

In order to collect intrinsic data for the kinetic study, an operating region free of

heat and mass transfer limitations was experimentally established. The stability of

5Ni/CeZrCa catalyst was then confirmed via an extended TOS run for a period of 15

hours. The mass distribution in the process was also studied by a carbon material balance

around the reactor. The carbon content of outlet streams was experimentally measured

using different techniques. The majority of carbon in the inlet feed was found to move

into the gas phase with a negligible deposition over the catalyst surface. A power law

model was then suggested for the rate expression. An integral analysis for reaction

kinetics was applied to estimate the parameters of the model. The activation energy was

found to be 93.7 kJ/mol; along with reaction orders of 1, 0.5 and 2 with respect to

synthetic crude glycerol, steam and oxygen, respectively. The predicted conversion

values were in an excellent agreement the experimental conversion results with an AAD

of 5.2%.

151

5.2 RECOMMENDATIONS

Based on the scope of this research work, it is recommended that:

A crude glycerol feed collected directly from biodiesel operations should be

tested at the optimum conditions obtained from this study.

A steam-to-carbon ratio of more than 2.6 would be more desirable in case of real

crude glycerol that contains more impurities and might yield more depositions.

Reactor modeling and simulations should be performed using software such as

Aspen or COMSOL Multiphysics.

The process should be tested at a pilot scale for further confirmation on its

performance and efficiency.

152

REFERENCES

1. Adhikari, S., Fernando, S. D., & Haryanto, A. (2009). Hydrogen production from

glycerol: An update. Energy Conversion and Management, 50(10), 2600-2604.

2. Adhikari, S., Fernando, S. D., & Haryanto, A. (2009). Kinetics and reactor

modeling of hydrogen production from glycerol via steam reforming process over

Ni/CeO2 catalysts. Chemical engineering & technology, 32(4), 541-547.

3. Adhikari, S., Fernando, S. D., To, S. F., Bricka, R. M., Steele, P. H., & Haryanto,

A. (2008). Conversion of glycerol to hydrogen via a steam reforming process over

nickel catalysts. Energy & Fuels, 22(2), 1220-122

4. Adhikari, S., Fernando, S., Gwaltney, S. R., Filip To, S. D., Mark Bricka, R.,

Steele, P. H., & Haryanto, A. (2007). A thermodynamic analysis of hydrogen

production by steam reforming of glycerol. International Journal of Hydrogen

Energy, 32(14), 2875-2880.

5. Ahmed, S., & Krumpelt, M. (2001). Hydrogen from hydrocarbon fuels for fuel

cells. International Journal of Hydrogen Energy, 26(4), 291-301.

6. Akpan, E., Akande, A., Aboudheir, A., Ibrahim, H., & Idem, R. (2007).

Experimental, kinetic and 2-D reactor modeling for simulation of the production

of hydrogen by the catalytic reforming of concentrated crude ethanol (CRCCE)

over a Ni-based commercial catalyst in a packed-bed tubular reactor. Chemical

engineering science, 62(12), 3112-3126.

7. Akpan, E., Sun, Y., Kumar, P., Ibrahim, H., Aboudheir, A., & Idem, R. (2007).

Kinetics, experimental and reactor modeling studies of the carbon dioxide

reforming of methane (CDRM) over a new Ni/CeO2–ZrO2 catalyst in a packed

bed tubular reactor. Chemical engineering science, 62(15), 4012-4024.

8. Albright, L. F., Crynes, B. L., & Corcoran, W. H. (1983). Pyrolysis, theory and

industrial practice.

9. Authayanun, S., Arpornwichanop, A., Paengjuntuek, W., & Assabumrungrat, S.

(2010). Thermodynamic study of hydrogen production from crude glycerol

autothermal reforming for fuel cell applications. International Journal of

Hydrogen Energy, 35(13), 6617-6623.

153

10. Ayoub, M., & Abdullah, A. Z. (2012). Critical review on the current scenario and

significance of crude glycerol resulting from biodiesel industry towards more

sustainable renewable energy industry. Renewable and Sustainable Energy

Reviews, 16(5), 2671-2686.

11. Bartels, J. R., Pate, M. B., & Olson, N. K. (2010). An economic survey of

hydrogen production from conventional and alternative energy sources.

International journal of hydrogen energy, 35(16), 8371-8384.

12. Bournay, L., Casanave, D., Delfort, B., Hillion, G., & Chodorge, J. A. (2005).

New heterogeneous process for biodiesel production: A way to improve the

quality and the value of the crude glycerin produced by biodiesel plants. Catalysis

Today, 106(1), 190-192.

13. Buffoni, I. N., Pompeo, F., Santori, G. F., & Nichio, N. N. (2009). Nickel

catalysts applied in steam reforming of glycerol for hydrogen production.

Catalysis Communications, 10(13), 1656-1660.

14. Cavendish, H. (1766). Three Papers, Containing Experiments on Factitious Air,

by the Hon. Henry Cavendish, FRS. Royal Society of London Philosophical

Transactions Series I, 56, 141-184.

15. Chen, H., Ding, Y., Cong, N. T., Dou, B., Dupont, V., Ghadiri, M., & Williams,

P. T. (2011). A comparative study on hydrogen production from steam-glycerol

reforming: thermodynamics and experimental. Renewable Energy, 36(2), 779-

788.

16. Cheng, C. K., Foo, S. Y., & Adesina, A. A. (2010). Glycerol Steam Reforming

over Bimetallic Co− Ni/Al2O3. Industrial & Engineering Chemistry Research,

49(21), 10804-10817.

17. Chhetri, A. B., Watts, K. C., & Islam, M. R. (2008). Waste cooking oil as an

alternate feedstock for biodiesel production. Energies, 1(1), 3-18.

18. Chi, Z., Pyle, D., Wen, Z., Frear, C., & Chen, S. (2007). A laboratory study of

producing docosahexaenoic acid from biodiesel-waste glycerol by microalgal

fermentation. Process Biochemistry, 42(11), 1537-1545.

154

19. Chiodo, V., Freni, S., Galvagno, A., Mondello, N., & Frusteri, F. (2010).

Catalytic features of Rh and Ni supported catalysts in the steam reforming of

glycerol to produce hydrogen. Applied Catalysis A: General, 381(1), 1-7.

20. Chisti, Y. (2007). Biodiesel from microalgae. Biotechnology advances, 25(3),

294-306.

21. Czernik, S., French, R., Feik, C., & Chornet, E. (2002). Hydrogen by catalytic

steam reforming of liquid byproducts from biomass thermoconversion processes.

Industrial & Engineering Chemistry Research, 41(17), 4209-4215.

22. Dauenhauer, P. J., Salge, J. R., & Schmidt, L. D. (2006). Renewable hydrogen by

autothermal steam reforming of volatile carbohydrates. Journal of Catalysis,

244(2), 238-247.

23. Davda, R. R., Shabaker, J. W., Huber, G. W., Cortright, R. D., & Dumesic, J. A.

(2005). A review of catalytic issues and process conditions for renewable

hydrogen and alkanes by aqueous-phase reforming of oxygenated hydrocarbons

over supported metal catalysts. Applied Catalysis B: Environmental, 56(1), 171-

186.

24. Deleplanque, J., Dubois, J. L., Devaux, J. F., & Ueda, W. (2010). Production of

acrolein and acrylic acid through dehydration and oxydehydration of glycerol

with mixed oxide catalysts. Catalysis Today, 157(1), 351-358.

25. Di Cosimo, J. I., Dıez, V. K., Xu, M., Iglesia, E., & Apesteguıa, C. R. (1998).

Structure and surface and catalytic properties of Mg-Al basic oxides. Journal of

Catalysis, 178(2), 499-510.

26. Dou, B., Dupont, V., Williams, P. T., Chen, H., & Ding, Y. (2009).

Thermogravimetric kinetics of crude glycerol. Bioresource technology, 100(9),

2613-2620.

27. Douette, A. M., Turn, S. Q., Wang, W., & Keffer, V. I. (2007). Experimental

investigation of hydrogen production from glycerin reforming. Energy & Fuels,

21(6), 3499-3504.

28. Dyer, J. A., Vergé, X. P. C., Desjardins, R. L., Worth, D. E., & McConkey, B. G.

(2010). The impact of increased biodiesel production on the greenhouse gas

155

emissions from field crops in Canada. Energy for Sustainable Development,

14(2), 73-82.

29. Encyclopædia Britannica. 2007. Encyclopædia Britannica Online. 24 July 2007.

30. Fogler, H. S. (2010). Essentials of chemical reaction engineering. Pearson

Education (Ch.7).

31. Froment, G. F., & Bischoff, K. B. (1990). Chemical Reactor Analysis and Design.

2nd ed. New York: Wiley.

32. Glycerin Market Analysis, ABG, Inc. for the U.S. Soybean Export Council

(2007). Available at: http://www.glycosbio.com/partners/biodiesel-producers/

[accessed on 10 April 2014]

33. International Panel on Climate Change (IPCC), 2014.

34. Guo, J., Lou, H., & Zheng, X. (2007). The deposition of coke from methane on a

Ni/MgAl< sub> 2</sub> O< sub> 4</sub> catalyst. Carbon, 45(6), 1314-1321.

35. Gutiérrez Ortiz, F. J., Ollero, P., & Serrera, A. (2011). Thermodynamic analysis

of the autothermal reforming of glycerol using supercritical water. International

Journal of Hydrogen Energy, 36(19), 12186-12199.

36. Gutiérrez Ortiz, F. J., Ollero, P., Serrera, A., & Galera, S. (2012). Process

integration and exergy analysis of the autothermal reforming of glycerol using

supercritical water. Energy, 42(1), 192-203.

37. Hammett, L. P., & Deyrup, A. J. (1932). A series of simple basic indicators. I.

The acidity functions of mixtures of sulfuric and perchloric acids with water1.

Journal of the American Chemical Society, 54(7), 2721-2739.

38. Hansen, C. F.; Hernandez, A.; Mullan, B. P.; Moore, K.; Trezona-Murray, M.;

King, R. H.; Pluske, J. R. A Chemical Analysis of Samples of Crude Glycerol

from the Production of Biodiesel in Australia, and the Effects of Feeding Crude

Glycerol to Growing- Finishing Pigs on Performance, Plasma Metabolites and

Meat Quality at Slaughter. Anim. Product. Sci. 2009, 49, 154−161

39. Hardiman, K. M., Cooper, C. G., Adesina, A. A., & Lange, R. (2006). Post-

mortem characterization of coke-induced deactivated alumina-supported Co–Ni

catalysts. Chemical engineering science, 61(8), 2565-2573.

156

40. Hill, J., Nelson, E., Tilman, D., Polasky, S., & Tiffany, D. (2006). Environmental,

economic, and energetic costs and benefits of biodiesel and ethanol biofuels.

Proceedings of the National Academy of Sciences, 103(30), 11206-11210.

41. Hirai, T., Ikenaga, N. O., Miyake, T., & Suzuki, T. (2005). Production of

hydrogen by steam reforming of glycerin on ruthenium catalyst. Energy & Fuels,

19(4), 1761-1762

42. Hoffmann, P. (2012). Tomorrow's energy: hydrogen, fuel cells, and the prospects

for a cleaner planet. MIT press.

43. Holladay, J. D., Hu, J., King, D. L., & Wang, Y. (2009). An overview of

hydrogen production technologies. Catalysis Today, 139(4), 244-260.

44. Hu, S., Luo, X., Wan, C., & Li, Y. (2012). Characterization of crude glycerol

from biodiesel plants. Journal of agricultural and food chemistry, 60(23), 5915-

5921

45. Ibrahim, H. H., & Idem, R. O. (2006). Kinetic studies of the partial oxidation of

isooctane for hydrogen production over a nickel–alumina catalyst. Chemical

engineering science, 61(17), 5912-5918.

46. Ibrahim, H. H., & Idem, R. O. (2008). Single and mixed oxide-supported nickel

catalysts for the catalytic partial oxidation reforming of gasoline. Energy & Fuels,

22(2), 878-891.

47. Idem, R. O., & Bakhshi, N. N. (1996). Kinetic modeling of the production of

hydrogen from the methanol-steam reforming process over Mn-promoted

coprecipitated Cu-Al catalyst. Chemical Engineering Science, 51(14), 3697-3708.

48. Idem, R., Khan, A.M., Saha, B. (2012) Sulphur Tolerant Catalysts for Hydrogen

Production by carbon dioxide reforming of methane rich gas, WIPO 2012/167351

A1.

49. Iriondo, A., Barrio, V. L., Cambra, J. F., Arias, P. L., Guemez, M. B., Sanchez-

Sanchez, M. C., ... & Fierro, J. L. G. (2010). Glycerol steam reforming over Ni

catalysts supported on ceria and ceria-promoted alumina. International Journal of

Hydrogen Energy, 35(20), 11622-11633.

50. Ishioma J.O. (2013). Hydrogen Production by Water Gas Shift Reaction

Following Dry Reforming of Biogas Methane in Membrane and Packedbed

157

Tubular Reactors (Master’s Thesis). University of Regina, Canada.

51. Ito, T., Nakashimada, Y., Senba, K., Matsui, T., & Nishio, N. (2005). Hydrogen

and ethanol production from glycerol-containing wastes discharged after biodiesel

manufacturing process. Journal of bioscience and bioengineering, 100(3), 260-

265.

52. IUPAC Recommendations, (1994). Pure Appl. Chem. 66, 1739.

53. Johnson, D. T.; Taconi, K. A. The Glycerin Glut: Options for the Value-Added

Conversion of Crude Glycerol Resulting from Biodiesel Production. Envi. Pro.

2007, 26, 338−348.

54. Kale, Ganesh R., and Bhaskar D. Kulkarni. "Thermodynamic analysis of dry

autothermal reforming of glycerol." Fuel Processing Technology 91.5 (2010):

520-530.

55. Kamonsuangkasem, K., Therdthianwong, S., & Therdthianwong, A. (2011).

Hydrogen Production via Oxidative Steam Reforming of Biodiesel By-products

over Ni/CeO2-ZrO2/Al2O3 Catalyst. In Proceedings of International Conference

on Chemistry and Chemical Process (ICCCP 2011).

56. Kapdan, I. K., & Kargi, F. (2006). Bio-hydrogen production from waste materials.

Enzyme and microbial technology, 38(5), 569-582.

57. Kašpar, J., & Fornasiero, P. A. O. L. O. (2002). Structural properties and thermal

stability of ceria-zirconia and related materials. Catalysis by Ceria and Related

Materials"//A. Trovarelli (Ed.), 217-241.

58. Khan, A., & Smirniotis, P. G. (2008). Relationship between temperature-

programmed reduction profile and activity of modified ferrite-based catalysts for

WGS reaction. Journal of Molecular Catalysis A: Chemical, 280(1), 43-51.

59. Khan, A., Sukonket, T., Saha, B., & Idem, R. (2011). Catalytic Activity of

Various 5 wt% Ni/Ce0. 5Zr0. 33M0. 17O2− δ Catalysts for the CO2 Reforming

of CH4 in the Presence and Absence of Steam. Energy & Fuels, 26(1), 365-379.

60. King, D. L., Zhang, L., Xia, G., Karim, A. M., Heldebrant, D. J., Wang, X., ... &

Wang, Y. (2010). Aqueous phase reforming of glycerol for hydrogen production

over Pt–Re supported on carbon. Applied Catalysis B: Environmental, 99(1), 206-

213.

158

61. Kulkarni, M. G., & Dalai, A. K. (2006). Waste cooking oil an economical source

for biodiesel: a review. Industrial & engineering chemistry research, 45(9), 2901-

2913.

62. Kunkes, E. L., Simonetti, D. A., Dumesic, J. A., Pyrz, W. D., Murillo, L. E.,

Chen, J. G., & Buttrey, D. J. (2008). The role of rhenium in the conversion of

glycerol to synthesis gas over carbon supported platinum–rhenium catalysts.

Journal of Catalysis, 260(1), 164-177.

63. Le Minh, C., Jones, R. A., Craven, I. E., & Brown, T. C. (1997). Temperature-

programmed oxidation of coke deposited on cracking catalysts: combustion

mechanism dependence. Energy & fuels, 11(2), 463-469.

64. Lehnert, K., & Claus, P. (2008). Influence of Pt particle size and support type on

the aqueous-phase reforming of glycerol. Catalysis Communications, 9(15), 2543-

2546.

65. Levenspiel, O., (1999), Chemical Reaction Engineering (3rdedition), Wiley, New

York.

66. Li, C. E., & Brown, T. C. (1999). Temperature-programmed oxidation of coke

deposited by 1-octene on cracking catalysts. Energy & fuels, 13(4), 888-894.

67. Li, C. E., & Brown, T. C. (2001). Carbon oxidation kinetics from evolved carbon

oxide analysis during temperature-programmed oxidation. Carbon, 39(5), 725-

732.

68. Luo, N., Fu, X., Cao, F., Xiao, T., & Edwards, P. P. (2008). Glycerol aqueous

phase reforming for hydrogen generation over Pt catalyst–effect of catalyst

composition and reaction conditions. Fuel, 87(17), 3483-3489.

69. Luo, N., Ouyang, K., Cao, F., & Xiao, T. (2010). Hydrogen generation from

liquid reforming of glycerin over Ni–Co bimetallic catalyst. Biomass and

bioenergy, 34(4), 489-495.

70. Luo, N., Zhao, X., Cao, F., Xiao, T., & Fang, D. (2007). Thermodynamic study

on hydrogen generation from different glycerol reforming processes. Energy &

Fuels, 21(6), 3505-3512.

71. Ma, F. R.; Hanna, M. A. Biodiesel Production: A Review. Bioresource

Technology 1999, 70, 1−15

159

72. McMichael, A. J. (Ed.). (2003). Climate change and human health: risks and

responses. World Health Organization.

73. McNaught, A. D., Wilkinson, A. (1997). Compendium of chemical terminology

(Vol. 1669). Oxford: Blackwell Science.

74. Meher Kotay, S., & Das, D. (2008). Biohydrogen as a renewable energy resource

prospects and potentials. International Journal of Hydrogen Energy, 33(1), 258-

263.

75. Menezes, A. O., Rodrigues, M. T., Zimmaro, A., Borges, L. E., & Fraga, M. A.

(2011). Production of renewable hydrogen from aqueous-phase reforming of

glycerol over Pt catalysts supported on different oxides. Renewable Energy,

36(2), 595-599.

76. Milligan Biofuels Inc. (2013, February 27). Stripped Glycerin. [Material Safety

Data Sheet]. Obtained directly from the company.

77. Mills, R. L., Ray, P., Dhandapani, B., Nansteel, M., Chen, X., & He, J. (2002).

New power source from fractional quantum energy levels of atomic hydrogen that

surpasses internal combustion. Journal of molecular structure, 643(1), 43-54.

78. Mizuno, T., Matsumura, Y., Nakajima, T., & Mishima, S. (2003). Effect of

support on catalytic properties of Rh catalysts for steam reforming of 2-propanol.

International journal of hydrogen energy, 28(12), 1393-1399.

79. Nagaoka, K., Seshan, K., Aika, K. I., & Lercher, J. A. (2001). Carbon Deposition

during Carbon Dioxide Reforming of Methane—Comparison between Pt/Al<

sub> 2</sub> O< sub> 3</sub> and Pt/ZrO< sub> 2</sub>. Journal of Catalysis,

197(1), 34-42.

80. Ozawa, M. (1998). Role of cerium–zirconium mixed oxides as catalysts for car

pollution: A short review. Journal of alloys and compounds, 275, 886-890.

81. Parmar, R. D., Kundu, A., Thurgood, C., Peppley, B. A., & Karan, K. (2010).

Kinetic studies of the autothermal reforming of tetradecane over Pt/Al< sub>

2</sub> O< sub> 3</sub> catalyst in a fixed-bed reactor. Fuel, 89(6), 1212-1220.

82. Pierotti, R. A., & Rouquerol, J. (1985). Reporting physisorption data for gas/solid

systems with special reference to the determination of surface area and porosity.

Pure Appl Chem, 57(4), 603-619.

160

83. Pompeo, F., Santori, G., & Nichio, N. N. (2010). Hydrogen and/or syngas from

steam reforming of glycerol. Study of platinum catalysts. International Journal of

Hydrogen Energy, 35(17), 8912-8920.

84. Rand, D. A. J., & Dell, R. (2008). Hydrogen energy: challenges and prospects

(Vol. 1). Royal Society of Chemistry.

85. Rennard, D. C., Kruger, J. S., & Schmidt, L. D. (2009). Autothermal Catalytic

Partial Oxidation of Glycerol to Syngas and to Non‐Equilibrium Products.

ChemSusChem, 2(1), 89-98.

86. Romero-Sarria, F., Vargas, J. C., Roger, A. C., & Kiennemann, A. (2008).

Hydrogen production by steam reforming of ethanol: Study of mixed oxide

catalysts Ce2Zr1. 5Me0. 5O8: Comparison of Ni/Co and effect of Rh. Catalysis

Today, 133, 149-153.

87. Sabri F. (2013). Catalysts for Hyrdrogen Production by the Auto-thermal

Reforming of Glycerol. (Master’s Thesis). University of Regina, Canada.

88. Saha, B., Khan, A., Ibrahim, H., & Idem, R. (2014). Evaluating the performance

of non-precious metal based catalysts for sulfur-tolerance during the dry

reforming of biogas. Fuel, 120, 202-217.

89. Schneider, P. (1995). Adsorption isotherms of microporous-mesoporous solids

revisited. Applied Catalysis A: General, 129(2), 157-165.

90. Sengupta, P. (2011). Catalysts for hydrogen production by the steam reforming of

mixtures of oxygenated hydrocarbons (Master’s Thesis). University of Regina,

Canada.

91. Shabaker, J. W., Huber, G. W., & Dumesic, J. A. (2004). Aqueous-phase

reforming of oxygenated hydrocarbons over Sn-modified Ni catalysts. Journal of

Catalysis, 222(1), 180-191.

92. Simon, Jonathan. "Authority and authorship in the method of chemical

nomenclature." Ambix 49.3 (2002): 206-226.

93. Slinn, M., Kendall, K., Mallon, C., & Andrews, J. (2008). Steam reforming of

biodiesel by-product to make renewable hydrogen. Bioresource technology,

99(13), 5851-5858.

161

94. Smith, E. G., Janzen, H. H., & Newlands, N. K. (2007). Energy balances of

biodiesel production from soybean and canola in Canada. Canadian Journal of

Plant Science, 87(4), 793-801.

95. Sørensen, B. (2012). Hydrogen and fuel cells: Emerging technologies and

applications. Academic Press.

96. Sukonket, T., Khan, A., Saha, B., Ibrahim, H., Tantanayon, S., Kumar, P., Idem,

R. (2011). Influence of the catalyst preparation method, surfactant amount, and

steam on CO2 reforming of CH4 over 5Ni/CeZrO catalysts. Energy Fuel, 25, (864

– 877).

97. Sutar, P. N., Vaidya, P. D., & Rodrigues, A. E. (2010). Glycerol‐Reforming

Kinetics Using a Pt/C Catalyst. Chemical Engineering & Technology, 33(10),

1645-1649.

98. Swami, S. M., & Abraham, M. A. (2006). Integrated catalytic process for

conversion of biomass to hydrogen. Energy & fuels, 20(6), 2616-2622.

99. Takehira, K., Shishido, T., & Kondo, M. (2002). Partial Oxidation of CH< sub>

4</sub> over Ni/SrTiO< sub> 3</sub> Catalysts Prepared by a Solid-Phase

Crystallization Method. Journal of Catalysis, 207(2), 307-316.

100. Thompson, J. C.; He, B. B. Characterization of Crude Glycerol from

Biodiesel Production from Multiple Feedstocks. Appl. Eng. Agric. 2006, 22,

261−265.

101. Trimm, D. L. (1997). Coke formation and minimisation during steam

reforming reactions. Catalysis Today, 37(3), 233-238.

102. Tsipouriari, V. A., & Verykios, X. E. (2001). Kinetic study of the catalytic

reforming of methane with carbon dioxide to synthesis gas over Ni/La< sub>

2</sub> O< sub> 3</sub> catalyst. Catalysis Today, 64(1), 83-90.

103. US President's Committee of Advisors on Science and Technology.

Report to the President on Federal Energy Research and Development for the

Challenges of the Twenty-First Century, Panel on Energy Research and

Development (November 1997)

162

104. Vaidya, P. D., & Rodrigues, A. E. (2009). Glycerol reforming for

hydrogen production: a review. Chemical engineering & technology, 32(10),

1463-1469.

105. Varrone, C., Giussani, B., Izzo, G., Massini, G., Marone, A., Signorini, A.,

& Wang, A. (2012). Statistical optimization of biohydrogen and ethanol

production from crude glycerol by microbial mixed culture. International Journal

of Hydrogen Energy, 37(21), 16479-16488.

106. Wang, F., Dubois, J. L., & Ueda, W. (2010). Catalytic performance of

vanadium pyrophosphate oxides (VPO) in the oxidative dehydration of glycerol.

Applied Catalysis A: General, 376(1), 25-32.

107. Wang, H., Wang, X., Li, M., Li, S., Wang, S., & Ma, X. (2009).

Thermodynamic analysis of hydrogen production from glycerol autothermal

reforming. International Journal of Hydrogen Energy, 34(14), 5683-5690.

108. Wang, W. (2010). Thermodynamic analysis of glycerol partial oxidation

for hydrogen production. Fuel Processing Technology, 91(11), 1401-1408.

109. Wang, X., Li, M., Wang, M., Wang, H., Li, S., Wang, S., & Ma, X.

(2009). Thermodynamic analysis of glycerol dry reforming for hydrogen and

synthesis gas production. Fuel, 88(11), 2148-2153.

110. Wawrzetz, A., Peng, B., Hrabar, A., Jentys, A., Lemonidou, A. A., &

Lercher, J. A. (2010). Towards understanding the bifunctional

hydrodeoxygenation and aqueous phase reforming of glycerol. Journal of

Catalysis, 269(2), 411-420.

111. Wen, G., Xu, Y., Ma, H., Xu, Z., & Tian, Z. (2008). Production of

hydrogen by aqueous-phase reforming of glycerol. international journal of

hydrogen energy, 33(22), 6657-6666.

112. Wijesekara, R. G., Nomura, N., Sato, S., & Matsumura, M. (2008).

Pre‐treatment and utilization of raw glycerol from sunflower oil biodiesel for

growth and 1, 3‐propanediol production by Clostridium butyricum. Journal of

chemical technology and biotechnology, 83(7), 1072-1080.

163

113. Yang, G., Yu, H., Peng, F., Wang, H., Yang, J., & Xie, D. (2011).

Thermodynamic analysis of hydrogen generation via oxidative steam reforming

of glycerol. Renewable Energy, 36(8), 2120-2127.

114. Yazdani, S. S., & Gonzalez, R. (2007). Anaerobic fermentation of

glycerol: a path to economic viability for the biofuels industry. Current opinion in

biotechnology, 18(3), 213-219.

115. Zhang, B., Tang, X., Li, Y., Xu, Y., & Shen, W. (2007). Hydrogen

production from steam reforming of ethanol and glycerol over ceria-supported

metal catalysts. International Journal of Hydrogen Energy, 32(13), 2367-2373.

116. Zhang, J., Xu, H., Jin, X., Ge, Q., & Li, W. (2005). Characterizations and

activities of the nano-sized Ni/Al< sub> 2</sub> O< sub> 3</sub> and Ni/La–

Al< sub> 2</sub> O< sub> 3</sub> catalysts for NH< sub> 3</sub>

decomposition. Applied Catalysis A: General, 290(1), 87-96.

117. Zhang, Y., Dube, M. A., McLean, D., & Kates, M. (2003). Biodiesel

production from waste cooking oil: 1. Process design and technological

assessment. Bioresource technology, 89(1), 1-16.

APPENDICES

164

APPENDIX A: Representative GC data sheets

Reaction T= 500 °C; Wcat= 0.1 g; S/C= 2.6; O/C= 0.125

165

Reaction T= 550 °C; Wcat= 0.25 g; S/C= 2.6; O/C= 0.125

166

Reaction T= 550 °C; Wcat= 0.15 g; S/C= 2.6; O/C= 0.125

167

Reaction T= 550 °C; Wcat= 0.1 g; S/C= 2.6; O/C= 0.125

168

Reaction T= 600 °C; Wcat= 0.15 g; S/C= 2.6; O/C= 0.125

169

Reaction T= 600 °C; Wcat= 0.1 g; S/C= 2.6; O/C= 0.125

170

Reaction T= 600 °C; Wcat= 0.05 g; S/C= 2.6; O/C= 0.1

171

Reaction T= 600 °C; Wcat= 0.25g; S/C= 2.6; O/C= 0.2

172

Reaction T= 650 °C; Wcat= 0.025 g; S/C= 2.6; O/C= 0.125

173

Reaction T= 500 °C; Wcat= 0.15 g; S/C= 2.6; O/C= 0.125

174

Reaction T= 500 °C; Wcat= 0.25 g; S/C= 2.6; O/C= 0.125

175

Reaction T= 550 °C; Wcat= 0.15 g; S/C= 2.6; O/C= 0.125

176

Reaction T= 600 °C; Wcat= 0 g; S/C= 2.6; O/C= 0.125

177

Reaction T= 600 °C; Particle size= 1.26 mm; S/C= 3.6; O/C= 0.125

178

Reaction T= 650 °C; Wcat= 0.25 g; S/C= 2.6; O/C= 0.125

179

Reaction T= 650 °C; Wcat= 0.2 g; S/C= 2.6; O/C= 0.125

180

Reaction T= 650 °C; Wcat= 0.15 g; S/C= 2.6; O/C= 0.125

181

Reaction T= 650 °C; Wcat= 0.1 g; S/C= 2.6; O/C= 0.125

182

Reaction T= 600 °C; Wcat= 0.25 g; S/C= 2.6; O/C= 0.125; L= 0.25 ml/min

183

Reaction T= 600 °C; Wcat= 0.25 g; S/C= 2.6; O/C= 0.125; L= 0.2 ml/min

184

APPENDIX B: Regression results from NLREG software for conversion model

---- Final Results ----

NLREG version 6.3

Copyright (c) 1992-2005 Phillip H. Sherrod. All rights reserved.

This is a registered copy of NLREG that may not be redistributed.

Number of observations = 23

Maximum allowed number of iterations = 500

Convergence tolerance factor = 1.000000E-010

Stopped due to: Function did not converge before iteration limit reached.

Number of iterations performed = 500

Final sum of squared deviations = 1.8474232E+002

Final sum of deviations = -8.6097533E-002

Standard error of estimate = 4.09814

Average deviation = 2.2959

Maximum deviation for any observation = 6.44201

Proportion of variance explained (R^2) = 0.8214 (82.14%)

Adjusted coefficient of multiple determination (Ra^2) = 0.6428 (64.28%)

Durbin-Watson test for autocorrelation = 1.463

Analysis completed 1-Jun-2014 16:55. Runtime = 0.16 seconds.

185

---- Descriptive Statistics for Variables ----

Variable Minimum value Maximum value Mean value Standard dev.

------------------ -------------- -------------- -------------- --------------

calcT 550 650 643.4783 22.88483

reducT 500 700 600 30.15113

PVtoSA 0.0017708 0.0029128 0.001958778 0.0002103501

Reducibility 0.0016 0.00208 0.002057391 0.0001000553

reactionT 500 650 595.6522 25.73044

StoC 1.6 3.6 3.165217 0.5897678

OtoC 0.05 0.25 0.1271739 0.03099949

con 60.3665 92.925 84.14085 6.85719

---- Calculated Parameter Values ----

Parameter Initial guess Final estimate Standard error t Prob(t)

------------------ ------------- ---------------- -------------- --------- -------

a 1 -393.165709 0 1.0E+030 0.00001

b 1 -1.34168725E-005 0 1.0E+030 0.00001

c 1 12.9106537 0 1.0E+030 0.00001

d 1 0.539416108 0 1.0E+030 0.00001

g 1 -0.751241511 0 1.0E+030 0.00001

h 1 63.3448397 0 1.0E+030 0.00001

f 1 60.3999101 0 1.0E+030 0.00001

k 1 0.547904849 0 1.0E+030 0.00001

q 1 4.18480948 0 1.0E+030 0.00001

x 1 2.17347658 0 1.0E+030 0.00001

186

y 1 0.233845151 0 1.0E+030 0.00001

z 1 4.05743302 0 1.0E+030 0.00001

---- Analysis of Variance ----

Source DF Sum of Squares Mean Square F value Prob(F)

---------- ---- -------------- -------------- --------- -------

Regression 11 849.7209 77.24736 4.60 0.00890

Error 11 184.7423 16.79476

Total 22 1034.463

187

APPENDIX C: Regression results for reaction kinetics using PLKA software

188

APPENDIX D: TOC Carbon Analysis results from Loring Laboratories LTD.

189

APPENDIX E: Synthetic CG average molecular formula

Average “CHO” molecular formula calculations for synthetic CG

For the sake of simplicity, Potassium and Sodium atoms were not included in the overall

formula

Note: the very small amounts of ethanol and phenolphthalein come from the soap solution

Chemical Formula Amount

(mol)

C H O Component

fraction

C H O

Glycerol C3H8O3 0.25518 3 8 3 0.4534252 1.36028 3.6274 1.36028

Methanol CH3OH 0.17478 1 4 1 0.3105638 0.31056 1.24226 0.31056

2-propanol C3H8O 0.1284 3 8 1 0.2281519 0.68446 1.82522 0.22815

Potassium

palmitate

C16H31KO2 0.00123 16 31 2 0.0021856 0.03497 0.06775 0.00437

Oleic acid C18H34O2 0.0023 18 34 2 0.0040868 0.07356 0.13895 0.00817

Ethanol C2H5OH 0.00051 2 6 1 0.0009009 0.0018 0.00541 0.0009

Phenolphthalein C20H14O4 0.00039 20 14 4 0.0006859 0.01372 0.0096 0.00274

Total 0.56278 SUM 1 2.47935 6.91658 1.91518