Sie sind auf Seite 1von 278

The effects of chemical and physical properties

of chars derived from inertinite-rich, high ash


coals on gasification reaction kinetics


Gregory Nworah Okolo
B.Eng (Chem. Eng.) (ESUT, Enugu, Nigeria)


Dissertation submitted in partial fulfilment of the requirements for the
degree Master of Engineering in Chemical Engineering at the
Potchefstroom Campus of North-West University, South Africa.

Supervisor: Prof. R. C. Everson
Co-supervisor: Prof. H. W. J. P Neomagus

November 2010
Potchefstroom







i










Any fact facing us is not as important as our attitude towards it,
for that determines our success or failure

-Norman Vincent Peale-













ii

DEDICATION



This dissertation is gratefully dedicated to the loving memory of my late sister,
Miss Juliana Okolo, who passed on to glory on the 13
th
of April, 2008.
May her gentle soul rest in perfect peace!
Eternal rest grant unto her Oh Lord, And let perpetual light shine upon her, May
she rest in peace. Amen


iii

DECLARATION


I, Gregory Nworah Okolo, do hereby declare that the dissertation with the title: The
effects of chemical and physical properties of chars derived from inertinite-rich,
high ash coals on gasification reaction kinetics, submitted in partial fulfilment of
the requirements for the degree of Master of Engineering (Chemical Engineering) is
my work and has not been submitted at any other university either in part or as a
whole.


Signed at Potchefstroom on the................day of ....................................................2010.




..............................
Gregory N. Okolo


iv

ACKNOWLEDGEMENT

The author expresses his appreciation and gratefully acknowledges the following
people for their help, contribution and assistance during the course of this research:
The Almighty God and our Mother Mary, for the spiritual support, guidance,
courage and wisdom to persevere to the end.
Prof. Ray Everson and Prof. Hein Neomagus for their excellent foresight,
guidance and assistance, invaluable suggestions, criticisms and magnanimous
supervisorship without which this investigation would not have been
successful.
Prof. Harold Schobert of Penn State University for his advice, suggestions and
discussions on coal characterisation and carbon crystallite analysis.
Prof. Frans Waanders and Prof. John Bunt for their priceless, suggestions and
discussions and for editing the original draft of the dissertation.
Dr. Sabine Verryn (XRD Analytical and Consulting cc), for carbon crystallite
analyses; and Mrs Vivien du Cann (Petrograhic SA), for petrographic analyses
of the samples and interpretation of the results.
My Mum and Dad, brother and sisters: Ejiyke, Juliet, Vero, Agatha and
Paschal; for their prayers, love, morale support, perseverance, and patience.
My former boss Engr. Charles Chidebelu and Engr. Patrick ThankGod for
their motivation and support.
Mr. Jan Kroeze and Mr. Adrian Brock for keeping the experimental apparatus
in excellent and safe condition.
The Coal Research Group for their co-operation and lively arguments during
the weekly presentations.
All the personnel of the School of Chemical and Minerals Engineering.
My big friends in Joburg: Ezebuilo Hyginus A., Alum Gerald O., Nweke
Humpery N., Eluwa Nnamdi S., Okereke Chidozie P., Duru Nnamdi K.,
Ogbonna Kelechi Micheal, Anyaoku Kingsley C., Ifeanyi Blessing
Emmanuel.. Jide ka unu ji !!!!!!!!!


v

ABSTRACT
With growing energy demand across the world, where the good quality coals are
gradually going into extinction creating great opportunity for low grade coals, a better
understanding of the important properties of these coals and the subsequent chars, that
control their behaviour in various utilisation processes, becomes pertinent. An
investigation was therefore undertaken, to study the effects of chemical and physical
properties imparted on chars during pyrolysis on the subsequent gasification reaction
kinetics of typical South African inertinite-rich, high ash Highveld coals. Attempt was
made at following these changes in the transition from coals to chars by a detailed
characterisation of both the parent coals and the respective chars. These changes were
determined using various conventional and advanced techniques, which included
among others, carbon crystallite analysis using XRD and char carbon forms analysis
using petrography.
Three of the four original coals were characterised as Bituminous Medium rank C (B,
C and C2), while coal D2 was found to be slightly lower in rank (Bituminous Medium
rank D). The coals were rich in inertinites (> 54 vol. %, mmb with coal C2 having as
high as 79 vol. %, mmb) and high in ash contents (> 26.7 wt. %, db) and cabominerite
and minerite contents (26 - 39 vol. %, mmb). Inertinite-vitrinite ratios of the coals
were found to range from 1.93 to 26.3.
Characterization results show that both volatile matter and inherent moisture contents
decreased, while ash, fixed carbon and elemental carbon contents increased from
coals to chars. A good indication that the pyrolysis process was efficient. Elemental
hydrogen, oxygen and nitrogen contents decreased, whereas total sulphur content
increased from coals to chars. This reveals that the total sulphur contained in the char
samples were associated with the char carbon matrix and the minerals. Hydrogen-
carbon and oxygen-carbon ratios decreased considerably from coals to chars showing
that the chars are more aromatic and denser products than the original coals. Despite
the fact that mineral matter increased from coals to chars, the relative abundance of
the different mineral phases and ash components did not exhibit significant variation
amongst the samples. However, the alkali index was found to vary considerably
among the subsequent chars. Petrographic analysis on the coals and char carbon forms

vi

analysis on the chars reveal that total reactive components (TRC) decrease while the
total inert components (TIC) increase from coals to chars. The 0% gain in TIC
observed in char C2 was attributed to its relatively high content of partially reacted
maceral char carbon form. Total maceral reflectance shifted to higher values in the
chars (4.43 - 5.28 Rsc %) with respect to the coals (1.15 - 1.63 Rsc %); which
suggests a higher structural ordering in the chars. Carbon crystallite analyses revealed
that the chars were condensed (smaller in size) relative to the parent coals. Lattice
parameters: inter-layer spacing, d
002
, increased, while the average crystallite height,
L
c
, crystallite diameter, L
a
, and number of aromatic layers per crystallite, N
ave
,
decreased from coals to chars. Carbon aromaticity generally increased whereas the
fraction of amorphous carbon and the degree of disorder index decreased from parent
coals to the respective chars. Both micropore surface area and microporosity were
observed to increase while the average micropore diameter decreased from coals to
chars. This shows that blind and closed micropores were opened up during the
charring process.
It was generally observed that, although the original coal samples did not show much
variations in their properties (except for their maceral content), the subsequent chars
exhibited substantial differences both amongst themselves and from the parent coals.
The increasing orders of magnitude of micropore surface area, microporosity, fraction
of amorphous carbon and structural disorderliness were found to change in the
transition, a good indication that the chars properties varies from that of the respective
precursor coals.
Isothermal CO
2
gasification experiments were conducted on the chars in a Thermax
500 thermogravimetric analyser in the temperature range of 900 - 950 C with varying
concentrations of CO
2
(25 - 100 mol. %) in the CO
2
-N
2
reaction gas mixture at
ambient pressure (0.875 bar in Potchefstroom). The effects of temperature and CO
2

concentration were observed to be in conformity to established trends. The initial
reactivity of the chars were found to increase in the order: chars C2 < C < B < D2,
with char D2 reactivity greater than the reactivity of the other chars by a factor > 4.
Gasification reactivity results were correlated with the parent coals and chars
properties. Except for the rank parameter (the vitrinite reflectance), no significant
trend was observed with any other coal petrographic property. Correlations with char

vii

properties gave more significant and systematic trends. Major factors affecting the
gasification reactivity of the chars as it pertains to this investigation are: parent coal
vitrinite reflectance, and: aromaticity, fraction of amorphous carbon, degree of
disorder and alkali indices, micropore surface area, microporosity and average
micropore diameter of the chars.
The random pore model (chemical reaction controlling) was found to adequately
describe the gasification reaction experimental data (both conversions and conversion
rates). The determined activation energy ranged from 163.3 kJmol
-1
for char D2 to
235.7 kJmol
-1
for char B; while the order of reaction with respect to CO
2
concentration was 0.52 to 0.67 for the four chars. The lower activation energy of char
D2 was possibly due to its lower rank, lower coal vitrinite reflectance and higher
alkali index. The estimated kinetic parameters of the chars in this study correspond
very well with published results in open literature. It was possible to express the
intrinsic reactivity, r
s
, of the chars (rate of carbon conversion per unit total surface
area) using kinetic results, in empirical Arrhenius forms.

Keywords: Inertinite-rich coal, Coal and char properties, Char carbon forms, Carbon
crystallite analysis, Carbon dioxide gasification, kinetic modelling.




viii

OPSOMMING
Die toeneemlike gebruik van lae graad steenkool word al hoe meer n realiteit, as
gevolg van ho kwaliteit steenkole wat verkwis word om te voorsien in die behoeftes
van die groeiende energiekwessie dwarsoor die wreld. Dit is dus van uiterste belang
dat n goeie begrip rondom die eienskappe en gedrag van lae graad steenkole en hulle
gevolglike sintels gevorm word om sodanig gebruik te word in verskeie kommersile
prosesse. n Studie was dus onderneem om die effek van chemiese- en fisiese
eienskappe van die gevormde sintels van verskeie lae graad steenkole op die
gevolglike vergassings reaksie kinetika van tipiese Suid-Afrikaanse inertiniet ryk, ho
as steenkole te ondersoek. n Poging was aangewend om die verandering in die
karakteristieke eienskappe van steenkool tot die vorming van sintels te monitor, deur
gebruik te maak van n gedetaileerde karakteriserings ondersoek op beide die rou
steenkole en hulle gevormde sintels. Konvensionele- en gevorderde metodes soos
koolstof kristallografie met behulp van XRD en sintels koolstof vorm analise met
behulp van petrografie is ingespan.
Drie van die vier steenkole wat gebruik is, is gekarakteriseer as Bitumineuse
Gemiddelde Rang C (steenkole B, C en C2) steenkole, terwyl dit gevind is dat
steenkool D2 egter n effense laer rang gehad het (Bitumineuse Gemiddelde Rang D).
Al vier steenkole het ho inhoude van inertiniet (> 54 vol. %, mineraal basis met
steenkool C2 wat die hoogste inhoud gehad het met 79 vol. % mineraal basis), as (>
26.7 wt. %, dro basis), karbomineriet en mineriet (26-39 vol. %, mineraal basis)
bevat. Die inertiniet-vitriniet verhoudings van die vier steenkole het gewissel tussen
ongeveer 1.93 en 26.3.
Vanuit die karakteriserings resultate was dit duidelik dat beide die inherente
waterinhoude en vlugtige stofinhoude afneem, terwyl die aswaardes, vaste
koolstofinhoude en elementre koolstof inhoude dienooreenkomstig toeneem vanaf
steenkool tot sintels. Hieruit kon dit afgelei word dat die pirolise proses, vir die
generering van sintels, effektief was. Hiermee saam het die elementre waterstof-,
suurstof- en stikstof inhoude ook afgeneem, terwyl die swawel inhoud toegeneem het.
Die verhoogde swawel vlakke in die sintels toon aan dat dit hoogs waarskynlik meer
geassosieer is met die koolstof/sintels koolstofmatriks as met die anorganiese

ix

minerale. Die waterstof-koolstof- en suurstof-koolstof verhoudings het ook n
noemenswaardige afname getoon van steenkool tot sintel, wat aandui dat die
gevormde sintels meer aromaties en digter is as die oorspronklike steenkole. Ten
spyte van die feit dat minerale inhoud toeneem van steenkool na sintels kon daar geen
noemenswaardige variasie tussen die relatiewe voorkoms van die verskeie
mineraalfases en askomponente onderskei word nie. Dit is egter gevind dat die alkali
indeks noemenswaardig varieer tussen die verskillende sintels. Petrografiese analise
op die steenkole en sintels koolstof vormanalise op die sintels het getoon dat die totale
reaktiewe komponente (TRC) afneem terwyl die totale inerte komponente (TIC)
toeneem van steenkool na sintels. Die 0% toename in TIC vir sintels C2 kan
toegeskryf word aan sy relatiewe ho inhoud van parsiel gereageerde maserale
sintels koolstofvorm. Totale maserale reflektiwiteit was aansienlik hor vir die sintel
(4.43 5.28 Rsc %) as vir die oorspronklike steenkole (1.15 1.63 Rsc %), wat op n
hor strukturele geordenheid van die sintels dui. Vanuit koolstof kristallografie analise
was dit duidelik dat die sintels meer gekondenseer (kleiner in grootte) is as die
oorspronklike steenkole. Kristalstruktuur parameters soos die inter-laag spasiering
(d
002
) het toegeneem, terwyl die gemiddelde kristalhoogte (L
c
), kristaldiameter (L
a
) en
aantal aromatiese lae per kristal afgeneem het van steenkool na sintels. Koolstof
aromatisiteit het gevolglik toegeneem, terwyl die fraksie van amorfe koolstof en graad
van wanordelikheidsindeks afgeneem het van steenkool na sintels. Vanuit n fisiese
perspektief het beide die mikroporieuse oppervlakarea en mikroporositeit toegeneem
terwyl die gemiddelde mikroporieuse diameter afgeneem het vanaf steenkool tot
sintels. Hieruit kan afgelei word dat blinde en geslote mikroporie geopen het
gedurende die pirolise proses.
Algeheel was daar geen duidelike verskil tussen karakteristieke eienskappe van die
oorpsronklike steenkole nie, behalwe vir die verskil in maserale inhoud. In kontras het
daar groot karakteristieke verskille bestaan tussen die verskillende sintels asook van
hulle oorpronklike steenkole. Die toenemende orde van grootte van mikroporieuse
oppervlakarea, fraksie van amorfe koolstof en strukturele ongeordenheid dui
grotendeels daarop dat die sintels grootliks verskillend is van hulle oorspronklike
steenkole.

x

Isotermiese CO
2
vergassing is gedoen om die reaktiwiteit van die gevormde sintels te
toets. Vir hierdie doeleinde is gebruik gemaak van n Thermax 500
termogravimetriese analiseerder. Temperature tussen 900 en 950 C en CO
2

konsentrasies van 25 to 100 mol. % (CO
2
-N
2
reaksie gasmengsel) by atmosferiese
druk is gebruik (0.875 bar in Potchefstroom) om die sintelsreaktiwiteit te assesseer.
Die effek van temperatuur en CO
2
konsentrasie op die reaktiwiteit van die sintels het
ooreengestem met wat gevind is in literatuur. Hiermee saam het die aanvanklike
(inisile) reaktiwiteit van die sintels afgeneem in die volgende orde: C2 < C < B < D2
met sintel D2 wat se reaktiwiteit n faktor 4 groter was as die ander sintels.
Vergassings reaktiwiteitsresultate is verder gekorreleer met die karakteristieke
eienskappe van die oorspronklike steenkole en die sintels. Geen ander
noemenswaardige korrelasie is verkry tussen die petrografiese eienskappe van die
steenkole nie, behalwe vir die vitriniet reflektiwiteit. Korrelasies met die sintels-
eienskappe het meer sistematiese verduidelikings gelewer. Vir die betrokke studie was
die belangrikste faktore wat n rol gespeel het in vergassingsreaktiwiteit:
oorspronklike steenkool vitriniet reflektiwiteit; aromatisiteit-, fraksie amorfe koolstof-
, graad van wanordelikheidsindeks-, mikroporieuse oppervlakte-, mikroporositeit- en
gemiddelde mikroporie diameter van die sintels.
Die eksperimentele vergassingsresultate (beide omsetting en reaksietempo) kon
akkuraat beskryf word deur die willekeurige poriemodel (chemiese reaksie beherend).
Die bepaalde aktiveringsenergie het gewissel tussen 163.3 kJmol
-1
vir sintel D2 tot
235.7 kJmol
-1
vir sintels B, terwyl die reaksie-orde met betrekking tot CO
2

konsentrasie gewissel het tussen 0.52 en 0.67 vir die vier sintels. Die lae
aktiveringsenergie van sintels D2 kan heel waarskynlik toegeskryf word aan die
steenkool se lae vitriniet reflektiwiteit en hor alkali indeks. Die beraamde kinetiese
parameters van die sintels toon goeie ooreenstemming met wat bevind is in literatuur.
Dit was verder ook moontlik om die intrinsieke reaktiwiteit, van die sintels (tempo
van koolstofomsetting per eenheidsoppervlakarea) uit te druk deur gebruik te maak
van kinetiese resultate in empiriese Arrhenius vorms.



xi

TABLE OF CONTENTS
Dedication ..................................................................................................................... ii
Declaration ...................................................................................................................iii
Acknowledgement ....................................................................................................... iv
Abstract ......................................................................................................................... v
Opsomming ................................................................................................................viii
Table of contents .......................................................................................................... xi
List of figures ............................................................................................................. xvi
List of tables ............................................................................................................... xix
Nomenclature ............................................................................................................ xxi
Greek symbols ......................................................................................................... xxiv
Abbreviations ........................................................................................................... xxv
Conference Presentation Resulting from this Investigation ..............................xxviii

Chapter 1: GENERAL INTRODUCTION ......................................................... 1
1.1 Introduction ..................................................................................................... 1
1.2 Background Information and Motivation ........................................................ 2
1.2.1 Clean Coal Technologies .......................................................................... 4
1.3. Hypotheses of the Study .................................................................................. 6
1.4 The Objectives of the Study ............................................................................ 7
1.5 Scope of the Research Work ........................................................................... 9

Chapter 2: LITERATURE REVIEW ................................................................ 11
2.1 Introduction ................................................................................................... 11
2.2 Coal Gasification ........................................................................................... 12
2.2.1 The Industry Context of Coal Gasification ............................................. 13
2.2.2 History of Coal Gasification ................................................................... 14
2.2.3 Modern Coal Gasification ....................................................................... 15
2.3 The Coal Gasification Process ....................................................................... 17
2.4 Chemical and Physical Structure of Coal and Char ....................................... 18
2.4.1 Chemical and Physical Structure of Coal ............................................... 19
2.4.2 Chemical and Physical Structure of Coal Char ....................................... 22
2.4.3 The Crystallite Structure of the Carbon Basic Structural Unit (BSU) ..... 27
2.5 Coal, Char and Gasification Reactivity ......................................................... 30
2.6 Factors Influencing Gasification Reactivity .................................................. 31
2.6.1 Properties of the Parent Coal .................................................................. 32

xii

2.6.1.1 Volatile Matter Content .................................................................... 32
2.6.1.2 Fixed Carbon Content ...................................................................... 32
2.6.1.3 Petrographic Properties of Coal ...................................................... 33
2.6.2 Pyrolysis Conditions and Heat Treatment .............................................. 34
2.6.3 Chemical Structure and Composition of Coal and Char .......................... 38
2.6.4 Changes in Carbon Crystallite Properties ............................................... 39
2.6.5 Catalysis by Mineral Matter .................................................................... 42
2.6.6 Physical Structural Properties of Chars .................................................. 45
2.6.6.1 Total Surface Area ........................................................................... 45
2.6.6.2 Active Surface Area .......................................................................... 46
2.6.6.3 Surface Complex Concentration during Reaction ............................ 47
2.7 Methods of Measuring Gasification Reactivity ............................................. 48
2.7.1 Thermogravimetric Analysers ................................................................ 48
2.8 Char-CO
2
Gasification Reactions .................................................................. 50
2.8.1 Char-CO
2
Reaction Mechanism .............................................................. 50
2.9 Heterogeneous Char-Gas Kinetics ................................................................. 51
2.9.1 Reaction Rate Models ............................................................................. 51
2.9.2 Overall CO
2
Gasification Kinetics .......................................................... 54
2.10 Homogeneous Gas-Phase Reactions .......................................................... 55
2.11 Structural Kinetic Models .......................................................................... 56
2.11.1 The Volume Reaction Model .............................................................. 57
2.11.2 Shrinking Core Model ......................................................................... 58
2.11.3 Random Pore Model ............................................................................ 59

Chapter 3: COAL AND CHAR CHARACTERISATION ............................... 60
3.1 Introduction ................................................................................................... 60
3.2 Origin of Coal Samples ................................................................................. 61
3.3 Sample Preparation ........................................................................................ 61
3.4 Char Preparation at 900 C ............................................................................ 62
3.4.1 Charring Apparatus and Procedure ......................................................... 63
3.5 Coal and Char Characterisation Analyses ..................................................... 65
3.6 Coal and Char Characterisation Equipment and Techniques ......................... 66
3.6.1 Chemical Analyses .................................................................................. 66
3.6.2 X-ray Diffraction (XRD) Mineral Analysis ............................................ 67
3.6.3 Ash Analysis (XRF) ................................................................................ 68
3.6.4 X-ray Diffraction (XRD) Carbon Crystallite Analysis ............................ 69
3.6.5 Petrographic Analysis ............................................................................. 73
3.6.6 Structural Analysis .................................................................................. 74
3.6.6.1 CO
2
Adsorption Analysis .................................................................. 75
3.6.6.2 Helium Pycnometry .......................................................................... 76
3.7 Characterisation Results and Discussion ....................................................... 77
3.7.1 Chemical Analyses .................................................................................. 77
3.7.2 XRD Mineral Analyses ........................................................................... 80

xiii

3.7.3 Ash Analysis (XRF) ................................................................................ 83
3.7.4 X-ray Diffraction (XRD) Carbon Crystallite Analysis ............................ 85
3.7.4.1 Determination of Aromaticity of Coal and Char Samples ................ 91
3.7.4.2 Determination of Fraction of Amorphous Carbon of the Coal and
Char Samples ....................................................................................... 92
3.7.5 Petrographic Analyses ............................................................................ 98
3.7.5.1 Reflectance Properties ..................................................................... 98
3.7.5.2 Maceral Analysis of Parent Coals ................................................. 101
3.7.5.3 Microlithotype Analysis of Parent Coals ........................................ 103
3.7.5.4 Carbominerite and Minerite Analysis of Parent Coals ................... 105
3.7.5.5 General Condition of Coal Samples .............................................. 106
3.7.6 Char Carbon Forms Analysis ................................................................ 107
3.7.7 Physical Structural Analysis: Coal and Char Samples .......................... 118
3.8 Summary ...................................................................................................... 124

Chapter 4: EXPERIMENTAL: CHAR GASIFICATION WITH CARBON
DIOXIDE .................................................................................................................. 128
4.1 Introduction ................................................................................................. 128
4.2 Materials Used ............................................................................................. 129
4.2.1 Coals and Subsequent Chars ................................................................. 129
4.2.2 Reactant Gases ..................................................................................... 129
4.3 Reactivity Equipment: Thermogravimetry ................................................. 130
4.3.1 Thermax 500 Thermogravimetric Analyser (TGA) .............................. 130
4.3.2 Gas Supply ............................................................................................. 134
4.3.3 Data Acquisition Interface .................................................................... 134
4.4 Experimental Procedure ............................................................................. 135
4.5 TGA Experimental Programme ................................................................... 136

Chapter 5: GASIFICATION WITH CARBON DIOXIDE: RESULTS AND
DISCUSSION ........................................................................................................... 138
5.1 Introduction ................................................................................................. 138
5.2 Normalisation of the Experimental Results ................................................. 139
5.3 Reproducibility of the Experimental Results ............................................... 141
5.4 Effect of Operating Conditions on Char-CO
2
Gasification Reactivity ......... 141
5.4.1 Effect of Isothermal Temperature of Reaction ..................................... 141
5.4.2 Effect of CO
2
Concentration in the Reaction Gas ................................ 143
5.5 Determination of the CO
2
Reactivity of the Chars ....................................... 144
5.6 Effect of Coal and Char Properties on CO
2
Reactivity of the Chars ............ 147
5.6.1 Effect of Parent Coals Petrographic Properties ..................................... 147
5.6.1.1 Effect of Maceral Index and Modified Reactive Maceral Index of the
Parent Coals ...................................................................................... 149
5.6.1.2 Effect of Rank Parameter of the Parent Coals ................................ 152

xiv

5.6.2 Influence of Char Properties on Char-CO
2
Reactivity ........................... 153
5.6.2.1 Influence of Char Petrography (Char- TRC and TIC) .................... 153
5.6.2.2 Influence of Char Carbon Crystallite (Chemical Structural)
Properties ........................................................................................... 155
5.6.2.2.1 Influence of Char Aromaticity .................................................... 155
5.6.2.2.2 Influence of Fraction of Amorphous Carbon in Char ................ 156
5.6.2.2.3 Influence of Degree of Disorder Index of chars ......................... 157
5.6.2.3 Inherent Catalytic Effects of Ash Components of Chars ................. 158
5.6.2.4 Effect of Physical Structural Properties of Chars ......................... 159
5.6.2.4.1 Effect of Micropore Surface Area of Chars ................................ 159
5.6.2.4.2 Influence of Average Micropore Diameter of Chars .................. 160
5.6.2.4.3 Influence of Char Porosity .......................................................... 160
5.7 Comparison of the CO
2
Reactivity of the four Chars ................................... 161
5.8 Summary ...................................................................................................... 164

Chapter 6: CHAR GASIFICATION WITH CARBON DIOXIDE: KINETIC
MODELLING AND PARAMETERS EVALUATION ......................................... 166
6.1 Introduction ................................................................................................. 166
6.2 The Random Pore Model ............................................................................. 167
6.3 The Random Pore Model Equation ............................................................. 168
6.4 Validation Procedure ................................................................................... 171
6.5 Evaluation of Kinetic Parameters ............................................................... 173
6.5.1 Evaluation of the Structural Factor, ................................................ 174
6.5.2 Determination of the Time Factor, t
f
..................................................... 176
6.5.3 Determination of Activation Energy, E
a
............................................... 178
6.5.4 Determination of Order of Reaction, m ................................................ 181
6.5.5 Determination of Lumped Pre-exponential Factor,
'
so
k ........................ 183
6.6 Validation of Kinetic Model and Associated Parameters ............................. 184
6.7 Summary ...................................................................................................... 189

Chapter 7: CONCLUSION AND RECOMMENDATIONS .......................... 191
7.1 Introduction ................................................................................................. 191
7.2 General Conclusions .................................................................................... 192
7.3 Contributions to Knowledge Base of Coal Science and Technology ....... 195
7.4 Recommendations for Future Studies ...................................................... 196

REFERENCES ......................................................................................................... 197

APPENDICES .......................................................................................................... 218
APPENDIX A ........................................................................................................... 219
Coal and Char Characterisation and Results .................................................... 219

xv

A-1 Description of Standard Methods Used for Characterisation .................... 219
A-2 Vitrinite Reflectance Scan Histograms of Coal Samples .......................... 220
A-3 Total Maceral Reflectance Scan Histograms of Coal and Char samples .. 221
A-4 Outline of Classification System for Char Carbon Forms. ....................... 223
APPENDIX B ........................................................................................................... 224
Char-CO
2
Gasification Reactivity Results .......................................................... 224
B-1 Reproducibility of Experimental Results and Reactivity of the Chars ..... 224
B-2 Determination of CO
2
Reactivity of the Chars .......................................... 226
B-3 Effect of Isothermal Temperature of Reaction on the Char Reactivity ..... 228
B-4 Effect of CO
2
Concentration in the Reaction Gas on Char Reactivity ...... 230
B-5 Comparison of CO
2
Reactivity of the Four Chars ..................................... 232
APPENDIX C ........................................................................................................... 234
Evaluation of Kinetic Parameters and Gasification Modelling ........................ 234
C-1 Summary of Structural Parameter, Time Factor and Initial Reactivity of the
Chars .............................................................................................................. 234
C-2 Dimensionless Plots for Chars C, C2 and D2. ............................................... 237
C-3 Comparison of Experimental and Model Gasification Results for Chars C, C2
and D2 ............................................................................................................ 238
APPENDIX D ........................................................................................................... 240
Model Validation: Random Pore Model (RPM) ................................................ 240
D-1 RPM Fitting to the Experimental Data of Chars B, C, C2 and D2. ............... 240
D-2 RPM Fitting of Char Conversion Rate to Experimental Results for the Chars ...
................................................................................................................... 242























xvi

LIST OF FIGURES

Figure 1.1: Total world energy supply and generation by fuel respectively in 2006 .... 2
Figure 1.2: Scope of the research work. ........................................................................ 9
Figure 2.1: Coal gasification products ......................................................................... 15
Figure 2.2: Molecular model for the inertinite-rich Highveld coal. ............................ 22
Figure 2.3: Molecular model for the vitrinite-rich Waterberg coal ............................. 22
Figure 2.4: A schematic representation of the structural changes that occurs upon
heating of coal. ............................................................................................................. 24
Figure 2.5: Geometry optimised structural conformations of average coal and char
molecules and intermediates, in the coal to char pyrolysis reaction. ........................... 26
Figure 2.6: Schematic representation of a crystallite of graphite ................................ 27
Figure 3.1: Experimental setup for char preparation ................................................... 63
Figure 3.2: Raw diffractograms of coal and char samples. .......................................... 87
Figure 3.3: Corrected and smoothened diffractograms of coal and char samples. ...... 88
Figure 3.4: Comparison of coal and char diffractograms for samples B and C. .......... 90
Figure 3.5: Determination of area under d
002
and - band using HighScore Plus for
coal B and char C2. ...................................................................................................... 92
Figure 3.6: Determination of amorphous fraction of carbon, X
A
, from (002) profile of
coal C2 and char C. ...................................................................................................... 93
Figure 3.7: Relationship between aromaticity and fraction of amorphous carbon and
the atomic ratios of hydrogen and oxygen to carbon in coal samples. ........................ 96
Figure 3.8: Relationship between various crystallite parameters of char samples. ..... 97
Figure 3.9: Photomicrographs of different categories of char carbon forms. ............ 111
Figure 3.10: Photomicrographs of different categories of char carbon forms. .......... 112
Figure 3.11: Photomicrographs of different categories of char carbon forms. .......... 113
Figure 3.12: Comparison of parent coals macerals and their specific char carbon forms
in the chars. ................................................................................................................ 117
Figure 3.13: Comparison of the total inert and reactive components in the parent coals
and the resultant chars. ............................................................................................... 118
Figure 3.14: Skeletal density of coal and char samples. ............................................ 118
Figure 3.15 Micropore surface area of coal and char samples. ................................. 119
Figure 3.16: CO
2
adsorption isotherm plots for coal and char samples. .................... 122
Figure 3.17: Comparison of the CO
2
adsorption isotherm plots for coals and chars: B,
C, C2 and D2. ............................................................................................................. 123
Figure 4.1: Schematic representation of Thermax 500 TGA showing the essential
parts and gas flow system. ......................................................................................... 131
Figure 4.2: Photograph of Thermax 500 TGA showing the essential parts. .............. 132
Figure 5.1: Typical mass loss curve for char C2 at 900 C, 100% CO
2
, 0.875 bar. .. 139
Figure 5.2: Conversion-time plot for char C2 at 900 C, 100% CO
2
, 0.875 bar. ...... 140
Figure 5.3: Effect of temperature on the CO
2
reactivity of the chars at different
constant CO
2
concentrations, 0.875 bar. ................................................................... 142

xvii

Figure 5.4: Effect of CO
2
concentration on the char reactivity at various constant
isothermal temperatures, 0.875 bar. .......................................................................... 143
Figure 5.5: Rate of reaction versus fractional conversion for the chars at 25% CO
2

concentration, 0.875 bar. ........................................................................................... 145
Figure 5.6: Relationship between the initial reactivity of the chars and the
petrographic properties of the parent coals at 100% and 75% CO
2
, 0.875 bar. ........ 148
Figure 5.7: Relationship between initial reactivity of the chars and the MI and RMI*
of the parent coals at 100% and 75% CO
2
, 0.875 bar. .............................................. 152
Figure 5.8: Relationship between the initial reactivity of the chars and the vitrinite
reflectance (Rr %) of the parent coals at 100% and 75% CO
2
, 0.875 bar. ................ 153
Figure 5.9: Relationship between the initial reactivity of the chars with the char TRC
at 100% and 75% CO
2
, 0.875 bar. ............................................................................. 154
Figure 5.10: Relationship between the initial reactivity of the chars and the
aromaticity of the char samples at 50% and 25% CO
2
, 0.875 bar. ............................ 155
Figure 5.11: Relationship between the initial reactivity of the chars and the fraction of
amorphous carbon in chars at 50% and 25% CO
2
, 0.875 bar. ................................... 156
Figure 5.12: Relationship between the initial reactivity of chars and the degree of
disorder index, DOI, at 50% and 25% CO
2
, 0.875 bar. ............................................. 157
Figure 5.13: Influence of the alkali index on the initial reactivity of the chars at 50%
and 25% CO
2
, 0.875 bar. ........................................................................................... 158
Figure 5.14: Influence of the D-R micropore surface area of chars on their initial
reactivity at 100% and 75% CO
2
, 0.875 bar. ............................................................. 159
Figure 5.15: Influence of the average micropore diameter of chars on the initial
reactivity at 100% and 75% CO
2
, 0.875 bar. ............................................................. 160
Figure 5.16: Influence of porosity of char on the initial reactivity at 100% and 75%
CO
2
, 0.875 bar. .......................................................................................................... 161
Figure 5.17: Comparison of CO
2
reactivity of the chars at various temperatures and
CO
2
concentrations, 0.875 bar. .................................................................................. 162
Figure 6.1: Comparison of the gasification experimental and model results for char B,
0.875 bar. ................................................................................................................... 176
Figure 6.2: Arrhenius plots of the char-CO
2
gasification reaction at 100, 75, 50 and
25% CO
2
concentrations, 0.875 bar. ......................................................................... 180
Figure 6.3: Determination of char-CO
2
gasification reaction order for chars B and C at
various temperature and constant CO
2
concentrations, 0.875 bar. ............................ 181
Figure 6.4: Determination of char-CO
2
gasification reaction order for chars C2 and D2
at various temperature and constant CO
2
concentrations, 0.875 bar. ........................ 182
Figure 6.5: Parity plot of predicted versus actual t
f
values for all four char samples.184
Figure 6.6: Comparison between experimental and model gasification results for ... 186
char B. ........................................................................................................................ 186
Figure 6.7: RPM fitting of char conversion at different CO
2
concentrations. ........... 187
Figure 6.8: RPM fitting of char conversion rate for chars B, C, C2 and D2 at 25% CO
2

concentration and different temperatures. .................................................................. 188
Figure A-1: Vitrinite reflectance scan histogram of the coal samples. ..................... 220

xviii

Figure A-2: Total maceral reflectance scan histogram for coals and chars: B and C.
.................................................................................................................................... 221
Figure A-3: Total maceral reflectance scan histogram for coals and chars: C2 and D2.
.................................................................................................................................... 222
Figure B-1: Reproducibility results for chars B, C, C2 and D2 at different
experimental conditions, 0.875 bar. .......................................................................... 224
Figure B-2: Rate of reaction versus fractional conversion for chars B and C at
different experimental conditions, 0.875 bar. ............................................................ 226
Figure B-3: Rate of reaction versus fractional conversion for chars C2 and D2 at
different experimental conditions, 0.875 bar. ............................................................ 227
Figure B-4: Effect of temperature on the CO
2
reactivity of chars B and C, 0.875 bar.
.................................................................................................................................... 228
Figure B-5: Effect of temperature on the CO
2
reactivity of chars C2 and D2, 0.875
bar. ............................................................................................................................. 229
Figure B-6: Effect of CO
2
concentration on the reactivity of chars B and C, 0.875 bar.
.................................................................................................................................... 230
Figure B-7: Effect of CO
2
concentration on the reactivity of chars C2 and D2, 0.875
bar. ............................................................................................................................. 231
Figure B-8: Comparison of the CO
2
reactivity of chars at 100 and 75% CO
2

concentration, 0.875 bar. ........................................................................................... 232
Figure B-9: Comparison of the CO
2
reactivity of chars at 50 and 25% CO
2

concentration, 0.875 bar. ........................................................................................... 233
Figure C-1: Dimensionless plots of conversion versus reduced time for chars C and
C2. .............................................................................................................................. 237
Figure C-2: Dimensionless plot of conversion versus reduced time for char D2. ..... 238
Figure C-3: Comparison between the experimental and model gasification results of
char C. ........................................................................................................................ 238
Figure C-4: Comparison between the experimental and the model gasification results
of chars C2 and D2. ................................................................................................... 239
Figure D-1: RPM fitting of the experimental results of chars B and C at 0.875 bar. 240
Figure D-2: RPM fitting of the experimental results of chars C, C2 and D2 at 0.875
bar. ............................................................................................................................. 241
Figure D-3: RPM fitting of the experimental results of char D2 at 0.875 bar. ......... 242
Figure D-4: RPM fitting of the char conversion rates for char B at 25% and 50% CO
2

concentration and different temperatures. .................................................................. 242
Figure D-5: RPM fitting of the char conversion rates for chars B, C, C2 at different
experimental conditions, 0875 bar. ........................................................................... 243
Figure D-6: RPM fitting of the char conversion rates for chars C and D2 at different
experimental conditions, 0.875 bar. .......................................................................... 244




xix

LIST OF TABLES

Table 2.1: Gasification based power generating plants. .............................................. 16
Table: 2.2: Some of the Plants using FBDB

as Sasol

. ............................................ 17
Table 2.3: Basic structures and functional groups in coal ........................................... 38
Table 3.1: Size requirements for coal and char characterisation analyses. .................. 62
Table 3.2: Char production conditions. ........................................................................ 63
Table 3.3: Char yield after production. ........................................................................ 64
Table 3.4: Characterisation analyses conducted on the coal and char samples. .......... 65
Table 3.5: Analytical methods used for chemical and mineralogical analysis. ........... 66
Table 3.6: Analysis parameters and settings on the XRD system for mineral analysis.
...................................................................................................................................... 67
Table 3.7: Analysis parameters and settings on the XRD system for carbon crystallite
analysis. ........................................................................................................................ 70
Table 3.8: Result of proximate and chemical analyses of coal and char samples. ...... 79
Table 3.9: Pecentage of graphite and total crystalline mineral phases of the coal and
char samples from XRD results. .................................................................................. 81
Table 3.10: Mineral abundance of coals and chars (graphite free bases, (wt. %, gfb)).
...................................................................................................................................... 82
Table 3.11: Char sample ash chemistry on LOI and sulphur free basis (wt. %, lfb and
sfb ). .............................................................................................................................. 84
Table 3.12: Proximate analysis of raw and demineralised coal and char samples (wt.
%, db). .......................................................................................................................... 86
Table 3.13: Comparison of aromaticity results from HighScore Plus and Origin 6.1. 91
Table 3.14: Determination of amorphous fraction of carbon for coal C2 and char C. 94
Table 3.15: Result on carbon crystallite analysis using XRD. .................................... 94
Table 3.16: Reflectance properties of coal and char samples. ................................... 100
Table 3.17: Maceral component summary of the coal samples (vol. %, mmb). ........ 101
Table 3.18: Maceral compositions of coal samples (vol. %, mmb). .......................... 102
Table 3.19: Microlithotype analysis of coal sample (vol. %,mmb). .......................... 104
Table 3.20: Carbominerite and minerite results as percentage of total carbominerite
and minerite (vol. %, mmb). ....................................................................................... 106
Table 3.21: Char carbon forms analysis result (vol. %, mmb). .................................. 109
Table 3.22: Total reactive and inert components of coal and char samples (vol. %,
mmb) .......................................................................................................................... 116
Table 3.23: Physical structural properties of coal and char samples. ........................ 120
Table 4.1 Specifications of gaseous reagents. ........................................................... 129
Table 4.2: Thermax 500 TGA specifications. ............................................................ 133
Table 4.3: Reaction conditions for char-CO
2
gasification experiments. ................... 137
Table 5.1: Determined initial gasification reactivity, R of the char at various operating
conditions, 0.875 bar. ................................................................................................ 146

xx

Table 5.2: Results used to evaluate the maceral index (MI) and the modified reactive
maceral index (RMI*) of the parent coals on mineral matter free basis (mmfb). ....... 151
Table 6.1: Dimensionless structural parameters for the char pores. .......................... 174
Table 6.2: Summary of the structural parameter, time factors and initial reactivity for
char B, 0.875 bar. ...................................................................................................... 177
Table 6.3: Details of results of activation energy for the char samples at different CO
2
concentrations in the reaction gas, 0.875 bar. ........................................................... 179
Table 6.4: Details of result for reaction order with respect to CO
2
concentration at
different temperatures and constant CO
2
gas composition at 0.875 bar. ................... 182
Table 6.5: Determination of lumped pre-exponential factor for the four char samples,
0.875 bar. ................................................................................................................... 183
Table 6.6: Summary of structural and kinetic parameters for chars B, C, C2 and D2
.................................................................................................................................... 184
Table 6.7: Models used, structural parameter, and kinetic parameters obtained for
char-CO
2
gasification reaction by other investigators. .............................................. 185
Table B-1: Analysis of reproducibility and experimental error. ................................ 225
Table C-1: Summary of the structural parameter, time factor and initial reactivity for
char C, 0.875 bar. ...................................................................................................... 234
Table C-2: Summary of the structural parameter, time factor and initial reactivity for
char C2, 0.875 bar. .................................................................................................... 235
Table C-3: Summary of the structural parameter, time factor and initial reactivity for
char D2, 0.875 bar. .................................................................................................... 236


















xxi

NOMENCLATURE


Symbol Description Unit
A Breadth of X-ray beam mm
A() Absorption factor -
A, k
so
Pre-exponential factor min
-1
bar
-m

A
002
Area under the (002) peak
2

A
f
Final ash content of coal or char after demineralisation wt. %
A
i
Original ash content of coal or char before demineralisation wt. %
AI Alkali index -
A
x
Breadth of X-ray cm
A Area under the gamma side band of (002) peak
2

C
g
Concentration of gaseous reactant mole m
-3

CV Calorific value MJ kg
-1

d
002
Inter-layer spacing for a group of N
ave
parallel layers
d
p
Average diameter of coal or char particles m, mm
D
p
Pore diameter / average pore diameter
E Activation Energy kJ mol
-1

E
d
Effectiveness of demineralisation %
f(X) Structural factor m
-1

f
a
Carbon aromaticity -
GCV Gross calorific value MJ kg
-1

H
a
Hydrogen aromaticity -
I X-ray reduced intensity / X-ray intensity Atomic units / counts
I
002
Reduced intensity due to (002) reflection atomic units
I
am
X-ray reduced intensity due to amorphous carbon atomic units
I
cr
X-ray reduced intensity due to crystalline carbon atomic units
I
max
Maximum reduced intensity of (002) peak atomic units
K Constant depending on X-ray refection plane -
K Absolute temperature scale K
k, k
1,
k
2
, k
3
Reaction rate constant min
-1

k
so

Lumped pre-exponential factor min
-1
bar
-m


xxii

Symbol Description Unit
k
so
Pre-exponential factor m min
-1
bar
-m

k
v
Intrinsic rate constant of the volume reaction model. min
-1

K
1

X-ray radiation from Cobalt due to
1
K
counts
K
2

X-ray radiation from Cobalt due to
2
K
counts
L
a
Crystallite diameter
L
c
Crystallite height
L
o
Total pore length per unit volume m m
-3

M Molarity of acid M
m Order of reaction with respect to CO2 concentration -
m
ash
mass of ash mg
MI Maceral index -
m
o
Initial mass of char mg
m
t
Mass of char at time, t mg
N
ave
Average number of aromatic layers per carbon crystallite -
P
n
Fraction of aromatic carbon contained within the d
002
peak -
R Ideal gas constant J K mol
-1

R Initial reactivity of the chars min
-1

r
1
, r
2
, r
3

Reaction rates min
-1

RMI Reactive maceral index -
RMI* Modified reactive maceral index -
Rr Mean random vitrinite reflectance %
r
s
reaction rate m min
-1

Rsc Mean random maceral reflectance %
s 2sin/
-1

S
max
Value of s (2sin/) at which I
max
occurs
-1

S
o
Initial surface area m
2
m
-3

T Temperature C or K
t Time min
t
0.5
Time for fractional carbon conversion of 50% min
t
0.9
Time for fractional carbon conversion of 90% min
t
f
Time factor min
-1


xxiii

Symbol Description Unit
V Volume per unit mass m
3
g
-1

X Fractional conversion of carbon -
X
A
Fraction of amorphous carbon -
CO
y

Molar fraction / partial pressure of CO - / bar
2
CO
y
Molar fraction / partial pressure of CO
2
- / bar



























xxiv

GREEK SYMBOLS

Symbol Description Unit
Wavelength of incident X-ray

Full width at half maximum of the corresponding peak or band degrees ()
o

Initial porosity of char samples %
Peak position / XRD angle of scan degrees ()
002

Peak position of (002) peak degrees ()
10

Peak position of (10) peak degrees ()
11

Peak position of (11) Peak degrees ()
/ Absorption coefficient for Cobalt-K radiation -
Skeletal density / density of coal or char samples kg m
-3

',
c
Bulk density of coal or char sample kg m
-3

Standard deviation various unit
Dimensionless time -
a
E
s
Slope of ) ln(
f
t against T
-1
at constant CO
2
concentration
K
-1

t
s

Slope of the plot of real time t , against 1 ) 1 ln( 1 X ,
min
9 . 0

Dimensionless time at 90% conversion -

Dimensionless structural parameter for char pores -












xxv

ABBREVIATIONS
Acronym Meaning
% Ave. Dev. Percent average deviation
A.S.M.E. American Society of Mechanical Engineers
ACT Advanced Coal Technology, Pretoria
adb Air dry basis
afb Ash free basis
Afrox African Oxygen
AI Alkali index
ASA Active surface area
ASAP Accelerated surface area and porosimetry
ASTM American Society for Testing Materials
Ave. Average value
Ave. Dev. Average deviation
BET Brunauer-Emmett-Teller Method
BFBC Bubbling fluidised bed combustion
Bit. Med. Bituminous Medium Rank
BSU Basic structural unit
Cat. Category
CCT Clean Coal Technology
CDM Clean Development Mechanism
CFBC Circulating fluidised bed combustion
daf Dry ash free basis
db Dry basis
Demin Demineralised coal or char sample
DME Department of Minerals and Energy
dmmb Dry mineral matter basis
DOI Degree of disorder index
D-R Dubinin-Radushkevich method
DTF Drop tube furnace
EFR Entrained flow reactor

xxvi

Acronym Meaning
ESKOM South African Electricity Supply Commission
ESS Error sum of squares
FBC Fluidised bed combustion
FBDB Fixed bed dry Bottom gasifier
FBG Fluidised bed gasification
FC Fixed carbon
Fig. Figure
FWHM Full width at half maximum
gfb Graphite (carbon) free basis
H/C Hydrogen-carbon atomic ratio
HCL Hydrochloric acid
HF Hydrofluoric acid
H-K Horvath-Kawazoe method
HP Helium pycnometry
HPTGA High pressure thermogravimetric analyser
HRTEM High resolution transmission electron microscopy
HTR Horizontal tube reactor
ID Identity
IGCC Integrated Gasification Combined Cycle
IR Infra-red
ISO International Standard Organisation
lfb LOI free basis
LMO Local molecular orientation
LOI Loss on ignition
LTB Lithium tetraborate
MIP Mercury intrusion porosimetry
mmb Visible mineral matter basis
mmfb Visible mineral matter free basis
MOD Molecular orientation domain
NMR Nuclear magnetic resonance
NO
X
Oxides of nitrogen
NWU North-West University

xxvii

Acronym Meaning
O/C Oxygen-carbon atomic ratio
PBBR Packed bed balance reactor
PCC Pulverised coal combustion
PCI Pulverised coal injection
PDTF Pressurised drop tube furnace
PF Pulverised fuel
PFB Pressurised fluidised bed
Pp Page number / pages
PSD Position sensitive detectors
rpm Revolution per minute
RPM Random pore model
SA South Africa
SABS South African Bureau of Standards
SCM Shrinking core model
SO
X
Oxides of sulphur
sfb Sulphur free basis
TGA Thermogravimetric Analyser
TIC Total inert components
TPD Temperature programmed desorption
TRC Total reactive components
TSA Total surface area
UCG Underground Coal Gasification
UNFCCC The United Nations Framework Convention on Climate Change
VM Volatile matter content
vol. % Volume percent
VRM Volumetric reaction model
VTR Vertical tube reactor
WCI World Coal Institute
wt. % Weight percent
XRD X-ray diffraction
XRF X-ray fluorescence
The nomenclatures for the petrographic analyses are provided in the relevant sections
of Chapter 3 and Appendix A.

xxviii

Conference Presentation Resulting from this Investigation

Okolo, G.N., Everson, R.C. and Neomagus, H.W.J.P. (2010). The effects of
chemical and physical properties of chars derived from inertinite-rich, high ash
coals on CO
2
gasification reaction kinetics. Presented at the Fossil Fuel
Foundation of Africa 15
th
Southern African Conference on Clean Coal Energy,
17-18
th
November, 2010, Johannesburg, South Africa.


1











CHAPTER CHAPTER CHAPTER CHAPTER 1 11 1





1.0 GENERAL INTRODUCTION


1.1 Introduction
A brief introduction of the research project is given in this chapter. The basic
management question of what, why, and how are concisely treated here. Sections 1.1
- 1.2.1 give the background information and motivation for engaging in this study.
The working hypotheses and objectives of the research work are presented in Sections
1.3 - 1.4, while the scope of the research is laid down in Section 1.5.
Chapter 1:

1.2 Background Information
Coal is the most abundant fossil fuel and it will be available long after petroleum and
natural gas wells are dry (Podolski
fuel energy sources and resources. Currently
source for power generation as well as for industrial processes and this will remain
unchanged at least till 2030 (Cloke
other fossil fuels. It is safe to transp
achieving a diverse and balanced energy mix. It is also the major source of energy for
the developed and the developing economies; it provides 26% of world primary energy
needs and 41% of worlds electricity is generated from coal (WCI Coal Statistics,
2007). This is illustrated in Fig

Figure 1.1: Total world e


South Africa generated 95% of its electricity from coal (E
2008) and is ranked 5
th
in
of South Africas primary energy needs (DME: Digest of South African Energy
Statistics, 2006). ESKOM fact Sheet
General Introduction
2
Background Information and Motivation
ndant fossil fuel and it will be available long after petroleum and
natural gas wells are dry (Podolski et al., 2008). It is a major contribution of the fossil
fuel energy sources and resources. Currently, it is the most important primary energy
power generation as well as for industrial processes and this will remain
unchanged at least till 2030 (Cloke et al., 2003). Coal is comparatively cheaper than
other fossil fuels. It is safe to transport and storage is easier. Thus, it remains vital in
hieving a diverse and balanced energy mix. It is also the major source of energy for
the developed and the developing economies; it provides 26% of world primary energy
needs and 41% of worlds electricity is generated from coal (WCI Coal Statistics,
in Figure 1.1.
energy supply and generation by fuel respectively in 2006
Coal Statistics, 2007)
South Africa generated 95% of its electricity from coal (ESKOM
in the world coal export market. Coal also accounted for 77%
South Africas primary energy needs (DME: Digest of South African Energy
Statistics, 2006). ESKOM fact Sheet, (2007) also noted that this trend is not likely to
General Introduction

ndant fossil fuel and it will be available long after petroleum and
., 2008). It is a major contribution of the fossil
it is the most important primary energy
power generation as well as for industrial processes and this will remain
Coal is comparatively cheaper than
easier. Thus, it remains vital in
hieving a diverse and balanced energy mix. It is also the major source of energy for
the developed and the developing economies; it provides 26% of world primary energy
needs and 41% of worlds electricity is generated from coal (WCI Coal Statistics,
uel respectively in 2006 (WCI
Annual report,
Coal also accounted for 77%
South Africas primary energy needs (DME: Digest of South African Energy
also noted that this trend is not likely to
Chapter 1: General Introduction

3

change significantly in the next decade due to the relative lack of suitable alternatives
to coal as an energy source.
An average of 224 million tonnes of coal is produced annually in South Africa of
which 25% is exported and the remainder used locally by various coal utilisation
industries (DME Coal Statistics, 2006). Among these users are: Eskom, 53% for
electricity generation; Sasol, 33% for transport fuel and petrochemical production;
12% was consumed by the metallurgical industries and the remaining 2% was used
for domestic cooking and heating. These show the energy intensity level of South
Africa and its heavy dependency on coal, more so as it produces 45% of total
electricity in the African continent (ESKOM Fact Sheet, 2007). The Eskom Annual
Report (2008) also estimated coal reserves in South Africa at 53 billion tonnes and
with the present production rate there should be almost 200 years of coal supply left.
South Africas extractable coals are located in widely separated coal provinces
stretching interruptedly from the border with Botswana in the North-West, through
the Limpopo and Mpumalanga provinces and into Kwazulu-Natal in the east (Keaton
Energy, 2009). These coal provinces are themselves divided into distinct coalfields in
which most of the commercially mineable resources are contained in the Permian-
aged Vryheid formations of the Ecca Group (Highveld and Witbank coalfields)
(Snyman, 1989; Falcon, 1989; Snyman and Botha, 1993; Keaton Energy, 2009).
Other coalfields of emerging importance are: Waterberg, Soutpansberg and Ermelo
coalfields (Keaton Energy, 2009).
Most of the countrys coal (about 83%) is currently mined in the Highveld, Witbank
and Ermelo coalfields located in the Mpumalanga province (Cairncross, 2001).
Geology has determined that the Witbank and Highveld coalfields are by far the most
important source of South Africas mined coal at present. However, the Waterberg
deposits, which extend into Botswana, are widely expected to become the countrys
principal future coal resource (Snyman and Botha, 1993, Cairncross, 2001);
particularly as this is the region expected to be home to many of the new generation of
thermal power stations (Eskom Annual Report, 2008; Keaton Energy, 2009). The
Ultretch and Klip River coalfields in Kwazulu-Natal are comparatively small and
production has been slowly declining overall. However, the area produces most of the
Chapter 1: General Introduction

4

countrys anthracites as well as a fair part of its coking coals (Snyman and Botha,
1993; Cairncross, 2001).
Falcon et al. (2010) reported that the concentrated coal mining operation in some of
the coalfields (especially the Highveld coalfield), has had the effect that the current
products from this area are of low quality, Grade D (Calorific value (CV) < 25.5
kJmol
-1
); with ash content up to or greater than 40%. The application of these low
grade coals in conversion and utilisation processes (combustion and gasification), in
conventional existing commercialized facilities is therefore, limited by its low
efficiency, poor burnout characteristics, and increased equipment investment (greater
than 30%) (Hu et al., 2004). Furthermore, an indisputable huge amount of emissions
(SO
X
and NO
X
and particulate matter) that are environmentally unfriendly are
produced which require expensive downstream processing (Marban, et al., 1995;
Spalding-Fecher et al., 2000; Hu et al., 2004; Kaitano, 2007).
Current stringent legislation has put fossil fuel utility plants to pollutant emission
limits. This puts immense pressure on coal utility plants and coal resource users to
evolve technologies and ways of reducing emissions drastically. One of these
legislations is the United Nations Framework Convention on Climate Change
(UNFCC) Clean Development Mechanism (CDM), which is an offshoot of the Kyoto
protocol (UNFCC Website).
The coal utilisation industries and the government are continuously investigating,
assessing and encouraging various technologies for the utilisation of these low-grade
coals. In this regard, the government and major industries dependent on coal offer
their support to researches and studies on new technologies to efficiently use this vast
reserve in the Highveld, Witbank, Waterberg, Soutpansberg, and Ermelo coalfields.


1.2.1 Clean Coal Technologies
The growing worldwide awareness on pollutants emanating from coal usage has had
the result that processes reliant on coal as feedstock had to evolve new technologies
not as harmful to the environment as the older processes. Such processes are
commonly referred to as Clean Coal Technologies (CCT) and defined as technologies
Chapter 1: General Introduction

5

designed to enhance both the efficiency and the environmental acceptability of coal
extraction, preparation and utilisation (WCI, Ecoal, 2003). The main motive for
developing these new coal conversion technologies is the need to achieve significant
improvements in the important areas of fuel effectiveness, technical performance and
environmental impact protection (Osborne et al, 1996; Koornneef et al, 2007).
Various Clean Coal Technologies had emerged of which fluidised bed combustion
(FBC) and fluidised bed gasification (FBG) are more suited to the utilisation of these
South African low grade coals. These technologies reduce emissions and waste and
increase the amount of energy produced from each tonne of coal processed (Grainger
and Gibson, 1981).
FBC and FBG have gained rapid acceptance and commercialisation; can be operated
at atmospheric and pressurised conditions and has been shown to be a viable
alternative to Pulverised Coal Combustion (PCC) (Koornneef et al., 2007). One major
advantage of the FBC is that most of the pollutants produced during the conversion
processes as oxides of nitrogen- NO
X
and N
2
O can be removed by in-situ reduction
and sorbents such as dolomite can be added to capture
X
SO in the process (Marban et
al, 1995). The disposal of the associated ash from the high ash feedstocks can also be
done more effectively due to the fluidised state of the ash particles. FBC has two main
categories: the Bubbling Fluidised Bed Combustion (BFBC) and the Circulating
fluidised Bed (CFBC) and both categories have been developed operating at pressures
between atmospheric and 100 atmospheres and temperatures in the range of 750 C -
950 C. Particle sizes <25mm are used in CFBC while bigger particle sizes <50mm
may be used in BFBC (Kulasekaran et al., 1998; Koornneef et al, 2007). Koorneef et
al., (2007) also noted that the major difference between the BFBC and CFBC is the
particle fluidising velocity which is higher in CFBC than in BFBC, but the BFBC has
the advantage that its operation is much simpler and is more suitable for poor quality
coals which could require longer burnout time (Kaitano, 2007). Integrated
Gasification Combined Cycle (IGCC) power generation involving both gas and steam
turbines with fluidised bed gasification has been examined extensively and
commercialised internationally for coal feed stocks very different (more reactive) to
typical South African coals (Littlewood, 1977; Watkinson et al., 1991; Marban et al.,
Chapter 1: General Introduction

6

1995; Osborne et al., 1996; Buskies, 1996; Campbell et al., 2000; Koornneef et al.,
2007; Lee, 2007; GTC, 2008; Podolski et al., 2008).
This research work was therefore motivated by the quest to contribute positively to
the understanding of the inherent utilisation deficiencies of these low grade coals
through a better knowledge of its gasification kinetics with respect to the coal and
char properties. Until recently, coal processing, conversion and utilisation studies
were concentrated on high quality northern hemispheres coals, which have low ash,
high vitrinite and volatile matter content. These coals are completely different to
typical Highveld coals characterised by their low volatile matter, high ash and
inertinite content (Falcon, 1989; Snyman, 1989; Snyman and Botha, 1993; Cloke and
Lester, 1994); which caused Snyman and Botha (1993) to described South African
coals as abnormal in their review. Evolution of solutions and designs to use these
low grade coals will translate to an extended reserve as well as an improved
environment.


1.3. Hypotheses of the Study
The working hypotheses of this investigation are:
i. That the maceral composition of parent coals: vitrinites and inertinites, and
their ratio (inertinite-vitrinite ratio) have some effect on the reactivity of the
subsequent chars during gasification. Liptinites content in the coal samples
were very low (< 4 vol. %, mmb) and thus were not considered in this study.
Besides, the coal samples were found to be deficient in them. It should be
noted that the ash contents of the parent coal samples selected for this study
are approximately equal.
ii. That the changes that occur in these coals during the charring process (both
physical and chemical), influences the gasification reactivity of the chars.
Chapter 1: General Introduction

7

iii. That changes in chemical composition and structure of the carbon crystallite
(carbon BSU) at a molecular level, during this coal to char transition, impacts
on the resultant char gasification reactivity.


1.4 The Objectives of the Study
The overall objective of this research work is to investigate the effects of chemical
and physical properties imparted on chars during pyrolysis at 900 C, on their
subsequent gasification kinetics, and identify coal and char properties that influence
char reactivity. A suitable kinetic model that can best describe the gasification
behaviour of these chars derived from low grade coals towards carbon dioxide will
also be evaluated. The findings can be used in the design of gasifiers or implemented
on existing technologies as retrofits to effectively use these low quality coals with an
improved efficiency, economic and minimal environmental impact.
The specific objectives of this study are therefore as follows:
i. To study the subsequent changes in chemical, physical and structural
properties from coal to char. Resultant char carbon forms and changes in
crystallite properties of both coals and chars were also investigated.

ii. To conduct CO
2
gasification reactivity experiments on the chars on a
laboratory scale under conditions similar to fluidised bed gasification
conditions.

iii. To correlate the parent coals petrographic properties: maceral compositions,
vitrinite reflectance, inertinite-vitrinite ratio, maceral and modified reactive
maceral indices with the char CO
2
reactivity result.

iv. To also correlate the consequent chemical, physical and structural changes,
including changes in carbon crystallite properties of the chars with their
observed CO
2
gasification reactivity.

v. To apply the resulting data from the experiments in the kinetic modelling and
validation of the gasification reaction.
Chapter 1: General Introduction

8

To achieve these objectives, the following were undertaken:
The original coal samples were analysed for chemical, petrographic, and
physical structural properties prior to intermediate char production.

The subsequent chars were also analysed for chemical, physical, structural
and petrographic properties to understand the changes imparted on them
during the transition. Petrography was used here to identify the resultant char
carbon forms.

An advanced analytic technique- XRD was used to analyse the carbon
crystallite properties of both the precursor coals and the chars to identify
transitional changes in crystallite parameters such as inter-layer spacing,
crystallite height and diameter, aromaticity, fraction of amorphous carbon as
well as degree of disorder index.

CO
2
gasification experiments were conducted on the chars on a laboratory
scale using a Thermax 500, a high pressure thermogravimetric analyser
(HPTGA). The reactivity experiments involved the gasification of the chars
with a CO
2
- N
2
mixture isothermally at atmospheric pressure (0.875 bar in
Potchefstroom).

Char conversion and reactivity was determined using the experimental results.

Correlation of the various parent coal and char properties was done to
determine their contribution to the overall reactivity.

Determination of the gasification reaction kinetic parameters and fitting of the
experimental results to a suitable kinetic model taking into account the rate
controlling resistance observed during the experiments.

Testing and validation of the model against the experimental results and the
evaluated kinetic parameters.

Chapter 1: General Introduction

9

1.5 Scope of the Research Work
The research route followed in this study is as shown in the flow chart of Figure 2.
A brief introduction of the research work with some background information,
motivation, hypotheses and objectives are dealt with in chapter 1.


Figure 1.2: Scope of the research work.

Chapter 1: General Introduction

10

In chapter 2, a detailed review of relevant publications and literatures as it pertains to
the coal gasification industry in general, and char-
2
CO gasification in particular is
presented.
The methods used and the results of the various conventional and advanced analyses
used to characterise the parent coal and the resulting chars are presented in chapter 3,
while a description of the experimental apparatus, procedures and materials used are
given in chapter 4.
The reactivity results from the char-CO
2
gasification experiments including
correlations with parent coal and char properties and discussions thereof are shown in
chapter 5. These results were compared with the results of other investigators.
The kinetic modelling, validation and determination of the relevant kinetic parameters
from the reactivity experimental data are presented in chapter 6. The results were also
compared with that of other researchers.
From the foregoing, conclusions on the investigation are drawn and recommendations
for further studies made based on the findings of this study. These are presented in
chapter 7.










11









CHAPTER CHAPTER CHAPTER CHAPTER 2 22 2






2.0 LITERATURE REVIEW


2.1 Introduction
The objective of this chapter is to provide a literature review of research conducted in
the area of coal gasification reaction, especially as it pertains to gasification with CO
2
.
Char-CO
2
reaction rates, mechanisms and kinetics are also reviewed. Reactivity of
chars is discussed in relation to various factors that may influence it. Microstructural
and microtextural changes impacted on the carbon crystallite of chars, including
changes in various crystallite parameters upon devolatilisation and their effects on the
Chapter 2: Literature Review

12

gasification reactivity are also presented. Structural models associated with char-CO
2

gasification reactions are also reviewed. The survey presented in this chapter was
considered to provide the necessary background for the execution of this research
work.


2.2 Coal Gasification
Coal conversion by any of the processes to produce a mixture of combustible gases is
referred to as coal gasification, even though a large number of chemical reactions
other than the so called gasification reactions are involved (Littlewood, 1977; van
Heek and Mhlen, 1985; Kristiansen, 1996; Lee, 2007). It encompasses a series of
reaction steps that convert coal containing C, H, and O, as well as impurities such as
N and S, into synthesis gas and other forms of hydrocarbons. This is accomplished by
introducing a gasifying agent, which can be oxygen, carbon dioxide, steam, air, and/or
a mixture of two or more, or all of the above gasifying agents into a reactor vessel
containing coal feedstock where the temperature, pressure and flow pattern are
controlled (Littlewood, 1977; van Heek and Mhlen, 1985; Kristiansen, 1996; Kabe,
2004; Lee, 2007).
The proportions of the various species: CO, CO
2
, H
2
, CH
4
, H
2
O, N
2
, H
2
S, S
2
O, etc, in
the final product gas depend on the type of coal and its composition, the gasifying
agent or medium, and the thermodynamics and chemistry of the gasification reactions
as controlled by the process operating parameters.
Coal gasification technology can be used in the following energy systems of potential
importance (Lee, 2007).
i. Production of fuel for use in electric power generation.
ii. Production of synthesis gas for use as chemical feedstock for liquid transport
fuel and chemicals production.
iii. Production of hydrogen for fuel cell applications.
iv. Production of synthetic or substitute natural gas (SNG) for use as pipeline gas
supplies.
v. Generation of fuel gas for industrial purposes.
Chapter 2: Literature Review

13

2.2.1 The Industry Context of Coal Gasification
The utilisation of coal using some of the older technologies has been marked with
some undesirable environmental impacts. The major pollutants include oxides of
nitrogen and sulphur, ash and slag, particulate emissions and green house gases such
as carbon dioxide and nitrous oxide (Marban et al., 1995; Spalding-Fecher et al.,
2000). The role of these pollutants in urban smog formation, acid rain and enhanced
greenhouse effect are significant (Spalding-Fecher et al., 2000). Consequently, there
is strong incentive to reduce emissions and improve fuel efficiency of coal utilisation
technologies (Hu et al., 2004).
To address these challenges, new coal utilisation technologies have developed
(Grainger and Gibson, 1981; Osborne et al, 1996; Koornneef et al, 2007). Some are
variants of the Pulverised-Fuel (PF) fired coal boilers, such as Ultra Super-Critical PF
units with advanced
X
NO abatement technique and flue gas desulphurisation.
Pressurised fluidized bed (PFB) combustion and gasification, and advanced variants
of these technologies that are emerging are used for low rank, more reactive coals
(Osborne et al, 1996; Koornneef et al, 2007). These low ranked coals are usually rich
in inertinites and have high ash content as observed in the coals from the Highveld
coalfields of South Africa. Major advances in gas turbine engineering have meant that
Integrated Gasification Combined Cycle (IGCC) technologies are now a viable means
of achieving high efficiency in coal based power generation with reduced emissions.
The use of pulverised coal injection (PCI) in blast furnaces is also gaining popularity.
It reduces the amount of high-cost coking coals required by the iron smelting
processes and increases the efficiency of blast furnaces (Claudius Peters technologies,
2001).
Most of the advanced coal utilisation technologies operate at high temperatures and
pressures, which increases the reaction intensity and improves the efficiency of the
process. Typical thermal efficiencies of PF fired power stations are <37% whereas
super-critical PF units can achieve net efficiency of 47% (Campbell et al., 2000).
Power generation using IGCC systems achieve a thermal efficiency of about 47% and
it is believed that efficiencies of >50% are possible (Buskies, 1996; Campbell et al.,
2000)
Chapter 2: Literature Review

14

An understanding of coal and coal char behaviour at conditions relevant to these new
gasification technologies is required to enable optimum design of efficient coal
utilisation technology, and allow the prediction of the gasification behaviour of the
coal in a given gasifier. These issues are particularly important for the vast South
African coals, given its economic importance as a major export commodity and the
extent to which they are relied on domestically. In order to effectively match these
coals with technology, the gasification process and how it is affected by process
conditions, coal and char properties and their influences on the overall process need to
be understood.



2.2.2 History of Coal Gasification
Coal gasification is not a new technology. Littlewood (1977) reported that as far back
as 1792, gas produced from the carbonisation of coking coal was used for lighting.
Originally, a process similar to coking was used. However, in the 1860s, a process
that converted non-coking coal via gasification was demonstrated. In the late 1880s,
the chemical potential of the gasification process was demonstrated in the production
of ammonia (Littlewood, 1977). This technology quickly spread through Europe,
Japan and the US. The gasification of coal throughout the early to mid 1900s was
used to supply piped gas (town gas') for cooking and heating, and was produced in a
fixed bed gasifier, similar in principle to early coke ovens. The subsequent use of oil
as fuel and the discovery of large reserves of the relatively clean natural gas meant
that the reliance on coking coals for town gas production declined.
In more recent times, the large reserve of coal in South Africa, and in particular the
environmental concerns with the conventional facilities, has increased the importance
of clean coal gasification processes for large scale power generation, liquid fuels and
chemicals production.
Chapter 2: Literature Review

15

2.2.3 Modern Coal Gasification
Modern coal gasification systems are used to produce liquid fuels, various industrial
chemicals and to provide a cleaner, more efficient alternative to combustion-based
power generation systems. Power generation from coal gasification has been of
interest in recent years as the development of advanced turbine systems has made the
integration of coal gasification with gas and steam turbines much more viable
(Osborne et al., 1996).
Considering the reactor configuration and the method of gas-solid feedstock contact,
modern gasification processes can be grouped into four different types: entrained
flow, fluidised bed and fixed bed gasifiers (Littlewood, 1977; Kristiansen, 1996; Lee,
2007); the fourth group is the molten salt bath reactor- currently utilised by the Atgas
process (van Heek and Mhlen, 1985; Kabe, 2004; Lee, 2007). It should be noted that
the conditions under which these systems operate vary according to the type of coal
on which they are designed to run on. The products that can be obtained from modern
coal gasification are shown in Figure 2.1.


Figure 2.1: Coal gasification products (GTC, 2008)
Chapter 2: Literature Review

16

Gasification has been used in the chemical, refining, and fertilizer industries for more
than 50 years and by the electric power industry for more than 35 years. Currently,
there are more than 140 gasification plants, with more than 420 gasifiers operating
worldwide (GTC, 2008). In South Africa, Sasol

uses fixed bed dry bottom (FBDB

)
gasification technology to produce synthesis gas which is a precursor in its liquid
transport fuels and chemicals production (van de Venter, 2005). The South African
power utility, ESKOM has also planned (if approved), the addition of a 350 MW
UCG- IGCC ultra-high efficiency unit to its Majuba power station which could
potentially be commissioned in the 2012 timeframe (Eskom Annual report, 2008).
Tables 2.1 and 2.2 show various plants around the world using gasification
technology for power generation, liquid fuels and chemicals (Note: The list of
gasifiers for power generation is limited to IGCC, while that for Chemicals and liquid
fuels is limited to Sasol

FBDB

gasifiers).


Table 2.1: Gasification based power generating plants (GTC, 2008).
Plant Name Location
Year
started Feedstock
Output
(MW)
Noun Buggenum, Netherlands 1994 Coal / Biomass 250
Wabash Terre Haute, IN, USA 1995 Coal / Petroleum coke 260
Tampa Electric Polk County, FL, USA 1996 Coal / Petroleum coke 250
Vrevosa
Vrevosa, Czech
Republic
1996 Coal / Petroleum coke 350
Schwarze Pumpe Lausitz, Germany 1996 Coal / Biomass 40
Pernis Refinery Rotterdam, Netherlands 1997 Visbreaker tar 120
Elcogas Puertollano, Spain 1998 Coal / Petroleum coke 300
ISAB Energy Sicily, Italy 2000 Asphalt 520
Sarlux Sardinia, Italy 2001 Visbreaker tar 545
Chawan IGCC Jurong Island, Singapore 2001 Tar 160
api Energia Falconara, Italy 2002 Visbreaker tar 280
Valero
Delaware City, DE,
USA
2003 Petroleum coke 160
Negishi IGCC Negishi, Japan 2003 Asphalt 342
Eni Sannazzaro Sannazzaro, Italy 2006 Oil residue 250
Fujian
Petrochemical
Quanzhou, China 2009 Oil residue 280
Eskom Majuba, South Africa 2012 ??? Coal 350
Total 4457

Chapter 2: Literature Review

17

Table: 2.2: Some of the Plants using FBDB

as Sasol

(van de Venter, 2005).


Plant name Location
Year
started Feedstock
Output
(MW) Products
Sasol Chemical
Industries
Sasolburg,
S. Africa
1955
100% Sub
bituminous coal
17
FBDB
Liquid
chemicals
Sasol Synfuels
Secunda,
S. Africa
1979
100% Sub
bituminous coal
80
FBDB
Liquid fuels and
chemicals
Dakota
Gasification
Company
Dakota,
USA
1985 100% Lignite
14
FBDB
Substitute
natural gas
Shanxi-Tianji
Coal Chemical
Company
Shanxi,
China
1978 Anthracite 5 FBDB
Ammonia for
fertilizer
production
Swartz Pumpe
Lausitz,
Germany
1996/2
000
80% Waste / 10%
Lignite / 10%
Bituminous coal
5 FBDB
Power /
Methanol
Yima
Yima,
china
2000
100% Sub
bituminous coal
2 FBDB Methanol
KFX, Gillete
Wyoming
Wyoming,
USA
2005
100% Sub
bituminous coal
2 FBDB
Coal
beneficiation



2.3 The Coal Gasification Process
The conversion of coal into fuel gas generally involves three essential steps
(Laurendeau, 1978; Watkinson et al., 1991; Yu et al., 2004):

i. The devolatilisation of the organic matter and mineral matter leading to
the formation of char.

ii. The homogeneous reactions of the volatile species from step1 with the
reactant gases.

iii. The heterogeneous reactions of the resultant char with the reactant
gases leading to the formation of gaseous products and ash.
The first step in which the coal is heated and rapidly loses hydrocarbons, tars and
other volatiles is referred to as pyrolysis or devolatilisation. This is by far the fastest
Chapter 2: Literature Review

18

of the three stages and as such does not play a major role in determining the overall
kinetics of the gasification system. It does, however affect the amount of char
produced and the structural properties of this char (Laurendeau, 1978; Cloke and
Lester, 1994). Thus, it has a strong influence on the levels of conversion achievable in
the gasifier. The combustion of the released volatiles also produces much of the heat
required for subsequent conversion of the char and this is the second step of the
gasification process. In the third step, the resultant char undergoes heterogeneous
reactions with the various gaseous species present in the gasifier. This stage produces
a gas that is typically rich in CO and H
2
. The reaction of char with O
2
, if present in the
gasifier also provide the much needed heat for the endothermic char-CO
2
and char-
H
2
O reactions that continue well after the O
2
has been consumed. The heterogeneous
reactions are much slower than the pyrolysis reactions.
Under fluidised bed combustion conditions (750 - 1000 C, 5 - 20 atm), pyrolysis is
completed in less than 10 seconds, while char burnout takes approximately 1000
seconds. In the case of fixed bed gasification (750 - 1250 C, 20 - 50 atm), pyrolysis
and char burnout is in the order of 1000 and 10,000 seconds respectively (Laurendeau,
1978).
The relative slowness of the heterogeneous char-gas reaction often means that it is the
rate-determining step during coal conversion (Cloke and Lester, 1994). Consequently,
this step plays a major role in the planning for the design of a gasifier, and in the
assessment of coal for use in a particular gasification technology. It is therefore
critically important that the char reaction kinetics are understood. Thus, the role
played by various properties and parameters as it relates to the parent coal and the
chars form the basis of this work. The following sections look at these factors.


2.4 Chemical and Physical Structure of Coal and Char
Before the fundamentals of char gasification kinetics and the factors that affect the
kinetics are discussed in detail, it is important to understand the nature of coal and
char in terms of their chemical and physical structures and typical properties such as
composition, pore structure, surface area, etc.
Chapter 2: Literature Review

19

2.4.1 Chemical and Physical Structure of Coal
Coal contains practically all of the natural elements, but the organic fraction consists
mainly of carbon and hydrogen, commonly referred to as macerals, with lesser
quantities of oxygen, sulphur and nitrogen. The distribution of these elements within
the organic matrix, impacts on the structure of the coal as well as influence its
pyrolysis and gasification behaviour. These macerals, which are optically
recognisable and distinguishable are derived from different precursors, and can be
significantly different chemically and physically (van Krevelen, 1981; Jones et al.,
1985; Falcon and Snyman, 1986; Snyman and Botha, 1993; Choi et al., 1989; Taulbee
et al., 1989; Blanc et al., 1991; Czechowski and Kidawa, 1991; dela Rosa et al., 1992;
Mastalerz and Marc Bustin, 1993; Cloke and Lester, 1994; Kabe et al., 2004;
Kruszewska, 2003; Everson et al., 2008b).
Macerals are grouped into three broad classes: vitrinites, liptinites and inertinites.
With increasing maturation, the structural differences between macerals become less
distinct, until at high rank (anthracite), they are almost impossible to distinguish
optically or chemically (van Krevelen, 1981; Falcon and Snyman, 1986). However, in
the bituminous rank range, the differences among maceral groups and in some cases
the differences among maceral components within a maceral group can be substantial.
The carbon aromaticity of macerals determined by
13
C NMR, typically decreases in
the order: inertinite > vitrinite > liptinite (Choi et al., 1989; Palmer et al., 1990). The
elemental composition also varies with the H/C ratio, decreasing in the order: liptinite
> vitrinite > inertinites

(Blanc et al., 1991; Mastalerz and Marc Bustin, 1993; Maroto-
Valer et al., 1994), although the values converge at higher rank (van Krevelen, 1981;
Kabe et al., 2004). The proton aromaticity, H
a
determined from
1
H NMR indicates
that maceral concentrates have H
a
values that increase in the order: liptinite vitrinite
<< inertinite with a low value of 0.5 and a high value of 0.76

(Vassallo et al., 1991).
This shows a significant difference in the distribution of hydrogen within the maceral
groups. From the elemental composition, and proton and carbon aromaticities, the
aliphatic atomic hydrogen to carbon ratio can be determined. There are some
ambiguities about these values, with values below 2.0 discounted by Love et al.
(1993) as unreasonable. However, these values for some coals are probably high as a
Chapter 2: Literature Review

20

result of the addition of the highly aliphatic liptinite maceral group which may be
significant for Northern Hemisphere coals. Furthermore, values > 2.0 would indicate
either a higher amount of
2
CH groups or a combination of and groups.
This would essentially lower the amount of hydroaromatics present. While atomic
C H values > 2.0 are possible for coals, vitrinites have been reported to have lower
values for most of the rank range due to their hydroaromatic nature (dela Rosa et al.,
1992; Hanna et al., 1992). Hence, maceral composition impacts on coal structure and
behaviour. Hanna et al. (1992) also reported that macerals from Australian coals may
be very diverse from macerals from similar ranks of coal from the Northern
Hemisphere or the Southern hemisphere and this further substantiates the complexity
of coal structural investigations.
The age or maturation of coal also influences the chemical structure and is best
represented by the parameter called rank of coal. The rank of coal is directly related to
the carbon content, with a highly graphitic material representing the ultimate product
of the coalification process. The oxygen content decreases with increasing rank, while
the hydrogen content remains relatively constant until about 89% elemental carbon
content, above which a sharp decrease occurs

(van Krevelen, 1981; Falcon and
Snyman, 1986; Kabe et al., 2004). Hatcher (1988) reported that the fraction of carbon
that is both aromatic and protonated for vitrinites has considerable scatter and as the
rank increases, there is a rapid decrease. With increasing maturation, aromaticity
increases, the size of aromatic carbon per cluster also increases slowly up to anthracite
at which point a rapid increase occurs (Maroto-Valer et al., 1994). Disordered or non-
crystalline carbon usually referred to as fraction of amorphous carbon also decreases
with increasing rank and decreasing atomic hydrogen content (Franklin, 1951; Hirsch,
1954; Lu et al., 2001; Maity and Mukherjee, 2006).
While the absolute quantities of the atoms- C , H, O, N and S contained in the
organic component of coal impacts on the physical structure, their spatial arrangement
and orientation in the molecule also has an influence. This is observed in the density
of the molecules and is referred to as helium density or skeletal density due to the
fact that helium which is the smallest inert molecule and as such fits best to penetrate
the pore system of coals is used. Gan et al. (1972) however noted that helium will not
penetrate closed porosity, thus giving a lower value than the true density. The
3
CH CH
Chapter 2: Literature Review

21

corrected helium density for macerals is rank dependent increasing in the order:
liptinite < vitrinite < inertinite (Falcon and Snyman, 1986). With increasing rank (%
C ), Gan et al. (1972) reported helium density values of approximately 1.4 gcm
-3
,
decreasing to a minimum of approximately 1.25 gcm
-3
at 80% C and then, increasing
sharply at around 90% C for US coals of various rank- from anthracite to lignite.
This agreed well with the investigations of Senneca et al. (1998) who investigated
among other parameters the helium density of petroleum coke and two different coals,
one of which is from South Africa. The elemental carbon content of the three carbon
materials are: 77.3, 68.0 and 89.05 %C and their determined helium densities are:
1.5, 1.4 and 1.6 gcm
-3
respectively. Strugala (1994) calculated the real densities of
hard coals from the proximate, ultimate and mineral analyses using empirical
formulae.
There have been substantial efforts to elucidate the molecular structure of coal. This
task is made difficult due to large variety of coals, the heterogeneity of a single coal
and the complexity of individual constituents. Despite these hindrances, there were a
number of early successful modelling efforts (Given, 1960; Shinn, 1984; Carlson,
1992). Given (1960) proposed a hypothetical structure for a bituminous coal vitrinite
with 82% carbon as one possible arrangement of the atoms present in the coal
molecule in line with the facts available at that time.
Major advances in the area of molecular modelling as applied to coal structure have
evolved over the years. These advances had helped in the investigations of Jones et al.
(1999); Mathews et al. (2001); and more recently Narkiewicz and Mathews (2008);
and Van Niekerk (2008). The molecular representations of two South African,
Permian-aged vitrinite-rich Waterberg and inertinite-rich Highveld coals were
elucidated by Van Niekerk (2008). He reported an elemental composition of
2 16 69 846 1000
S N O H C and
4 20 78 673 1000
S N O H C for the vitrinite-rich Waterberg and
inertinite-rich Highveld coal respectively. The molecular models of these coals are
shown in Figures 2.4 and 2.5 in line and van der Waals radii rendering (Van Niekerk,
2008). It should be noted that hydrogen is not shown, while the other atoms: carbon,
oxygen, nitrogen and sulphur present in the models are coloured as: green red, blue
and yellow respectively.
Chapter 2:



Figure 2.2: Molecular m



Figure 2.3: Molecular
2.4.2 Chemical and Physical
The exact nature of chars derived from coal
almost all the variables that describe the p
Literature Review
22
Molecular model for the inertinite-rich Highveld coal
2008).
Molecular model for the vitrinite-rich Waterberg coal
2008)
Chemical and Physical Structure of Coal Char
derived from coal is difficult to describe in general terms, as
variables that describe the physical properties of chars are
Literature Review


oal (Van Niekerk,

oal (Van Niekerk,
Char
general terms, as
are dependent on
Chapter 2: Literature Review

23

the conditions under which the char were made, the parent coal rank, maceral
composition and particle size. Unfortunately, the study of the chemical structure of
char has not been as rampant as the study of its physical structure. Generally
speaking, coal chars are porous solids consisting primarily of carbon and mineral
matter, incorporating small amounts of nitrogen, sulphur, hydrogen and oxygen into
this framework (Laurendeau, 1978).
The amounts of nitrogen, sulphur, hydrogen and oxygen in the char (relative to the
carbon) are often less than their amount in the parent coal, due to their removal during
devolatilisation. With increasing devolatilisation temperature and residence time, char
aromaticity increases with a corresponding loss of aliphatic groups (Pugmire et al.,
1991; Lu et al., 2001, 2002a and 2002b; Kawakami et al., 2006) and decreasing
fraction of disordered or amorphous carbon (Schoening, 1982 and 1983; Senneca et
al., 1998; Russell et al., 1999; Lu et al., 2001; Kawakami et al., 2006; Maity and
Mukherjee, 2006). Fletcher (1993) reported that rapid heating of Pittsburgh # 8 coal
showed little evidence of char cluster size growth, and suggested that cross-linking is
occurring at the same time as aliphatics are lost. The atomic H/C and O/C ratios
decreases (Fletcher, 1993), and the proton aromaticity increases (Chen et al., 1992).
Laurendeau (1978) also noted that the carbon in the char is usually more ordered than
in the parent coal giving rise to increased crystallinity and decreased fraction of
amorphous carbon, a result of the heat treatment during pyrolysis (Davis et al., 1995).
Ordering of the carbon crystallite structure decreases the occurrence of carbon edge
sites and unpaired surface electrons often associated with the reactivity of the char
(Kashiwaya and Ishii, 1991; Davis et al., 1995; Russel et al., 1999; Shim and Hurt,
2000; Takagi et al., 2004). The extent of structural ordering with respect to the
different annealing temperatures is represented in the schematic of Figure 2.4.

Chapter 2: Literature Review

24


Figure 2.4: A schematic representation of the structural changes that occurs upon heating
of coal (Russel et al., 1999).
Heat treatment of chars, in particular its effect on char reactivity, is discussed in detail
later in this review. However, fully devolatilised chars of various coal ranks have been
reported to be similar in their chemical structure (Fletcher, 1993; Chen et al., 1992).
The physical structural evolution of chars is related to thermoplasticity and hence,
char morphology (Cloke and Lester, 1994). Thermoplasticity influences particle size
and fragmentation which is very important for the hydrodynamics and heat and mass
transfer in fluidised bed and entrained flow gasifiers (van Heek and Mhlen, 1987).
The controlling feature of char morphology appears to be the extent of fluidity which
the particle can achieve. However, thermoplasticity is a transient property, and
appears to be more transient for certain macerals and for certain conditions. Upon
heating, thermal swelling generally occurs in vitrinites and liptinites (Cloke and
Lester, 1994; du Cann, 2008). The swelling can be substantial. Vitrinites swell to a
greater extent than the other macerals (Tsai and Scaroni, 1987). Fletcher (1993) also
noted that swelling can occur with little or no mass loss, prior to and during tar
release.
Generally, inertinites does not readily swell; although micrinites usually associated
with liptinites have been reported to vesiculate with liptinite during combustion.
Micrinites has also been known to yield more volatile matter than other inertinite
macerals, probably due to its connection to the liptinite fragments (Cloke and Lester,
1994).
Chapter 2: Literature Review

25

The porosity of chars depends to a large extent on the nature of porosity, maceral
content and rank of the precursor coals. The devolatilisation process, through the
softening of the particle and the release of volatiles, affects this pore structure
dramatically (Tsai and Scaroni, 1987; Cloke and Lester, 1994). The high temperatures
present during pyrolysis have been found to reduce the microporous nature of these
pores, as the plastic nature of the intermediate phase causes the closing of the fine
structures. The volatile matter released also impact on the porosity of the char, as the
release of volatile material through the softened particle can form larger pores and
voids. However, it is difficult to present a general description of the nature of the pore
structure of chars since the temperature, pressure and heating rate of pyrolysis all
seem to affect it in some way (Ng et al., 1988). Depending on these pyrolysis
conditions, the pores of a char range in size from nanometres to millimetres across.
They can be interconnecting with eventual paths to the external surface of the char, or
they can be the so-called blind pores, which are not connected to the surface but
which are often exposed after reaction as carbon is removed (Laurendeau, 1978). The
shapes of the pores in chars vary. Models depicting pore structure of chars often
simplify the case and use uniformly cylindrical pores. In reality, pores can be
cylindrical, spherical, or bottlenecked (pores that are large but have a restricted
opening or slits), or combinations of these. The size of pores is rarely uniform. Often
long cylindrical pores taper or expand, and the diameter of spherical pores is rarely
the same throughout a sample (Laurendeau, (1978).
Chars generally have higher
2
CO

and BET surface areas than the parent coal (Ng et
al., 1988) with values greater than 400 m
2
g
-1
reported for bituminous chars by Tsai
and Scaroni (1987). Surface area is clearly dependent on physical structure, with
possibly large differences in surface areas among chars of the same coal. Skeletal
density of chars are significantly higher than those of the parent coals (Nsakala et al.,
1978), with reported values in excess of 2.0 gcm
-3
(Ng et al., 1988), close to the
theoretical density of graphite- 2.26 g.cm
-3
. These high values are indicative of the
efficient packing of the carbon sheets, similar to that existing in graphite (Hirsch,
1954).
Not much effort has been put into the molecular modelling of char structure. Jones et
al. (1999) modelled the char structural evolution during the pyrolysis process in a
Chapter 2: Literature Review

26

wire-mesh reactor of Pittsburgh #8 coal. They were able to render the geometry
optimised structural conformations by energy minimisation as shown in Figure 2.5, of
the coal, its char and the intermediate volatiles. They also reported an average
molecular formula of
5 9 46 401 563
S N O H C for the parent coal. Average molecular
formulae of
2 6 26 193 469
S N O H C and
2 5 5 125 418
S N O H C

were reported for the chars with
holding times of 0.3 second and 1.0 second respectively (Jones et al., 1999).



Figure 2.5: Geometry optimised structural conformations of average coal and char
molecules and intermediates, in the coal to char pyrolysis reaction (Jones et al., 1999).


Chapter 2: Literature Review

27

2.4.3 The Crystallite Structure of the Carbon Basic Structural Unit
(BSU)
The crystallite structure of carbonaceous material refers to the microtextural and
microstructural characterisation of the carbon contained in them. Various
investigators gave a description of chars microstructure using analytic techniques as
HRTEM and XRD which can characterise the structure of the carbonaceous materials,
at atomic level (Franklin, 1951; Hirsch, 1954; Ergun, 1968; van Krevelen, 1981;
Rouzaud, 1990; Kashiwaya and Ishii 1991; Davis et al., 1995; Marsh, 1989; Lu et al.,
2001; Sharma et al., 2002; Feng et al., 2003; Takagi et al., 2004, Trejo et al., 2007).
They concluded that the crystallite of coal and char consists of polyaromatic basic
structural units (BSU) with size about 1 nm, formed by polyaromatic layers (4 to 10
rings) isolated or stacked by 2 or 3. The BSUs are ordered in stacked planes of the
aromatic layers often referred to as lamellae (Essenhigh, R. H. 1981; van Krevelen,
1981; Marsh, 1989; Kashiwaya and Ishii, 1991; Kabe et al., 2004), or molecular
orientation domains (MOD) or local molecular orientation (LMO) (Rouzaud et al.,
1988; Feng et al., 2003). A schematic representation of a crystallite of graphite is as
given in Figure 2.6. Inside the MODs the polyaromatic basic structural units (BSUs)
are either misoriented or locally orientated in parallel (Rouzaud, 1990).


Figure 2.6: Schematic representation of a crystallite of graphite (Essenhigh, 1981)
Chapter 2: Literature Review

28

The X-ray diffraction (XRD) pattern of a carbonaceous material such as coal and coal
chars shows diffuse peaks at (002), (10) and (11) bands that correspond to (002),
(100) and (110) reflections of graphite and strong low-angle or background scattering
(Franklin, 1951; Hirsch, 1954; Short and Walker, 1963; Schoening, 1982 and 1983;
Davis et al.,1995; Kashiwaya and Ishii, 1991; Lu et al., 2001, 2002a and 2002b;
Takagi et al., 2004; Kawakami et al., 2006; Bouhadda, 2007; Wu et al., 2008). The
diffuse peaks of the (002), (10) and (11) reflections indicate the presence of small
graphite-like domains (Franklin, 1951; Hirsch, 1954; Short and Walker, 1963; Lu et
al., 2001). It is also widely reported that the non-crystalline carbon or amorphous
carbon, forms the background intensity of X-ray diffraction pattern (Franklin, 1951;
Hirsch, 1954; Short and Walker, 1963; Schoening, 1982 and 1983; Davis et al.,1995;
Kashiwaya and Ishii, 1991;Lu et al., 2001, 2002a and 2002b; Takagi et al., 2004;
Kawakami et al., 2006; Bouhadda, 2007; Wu et al., 2008).
The crystallite is usually layered in the form ABABAB. The (002) reflection indicates
the stacking of the aromatic layers and both the (10) and (11) reflections correspond
to the two-dimensional lattices of the aromatic layers. The crystallographic parameters
that can be determined using the XRD pattern are: the aromatic layer stacking height
or crystallite height, L
c
; average crystallite size or diameter, L
a
and the inter-layer
spacing or inter-planar distance (d
002
) as annotated in Figure 2.6. Other parameters
that can be determined from the XRD diffractogram include the carbon aromaticity
and the fraction of amorphous carbon.
The aromatic layer stacking height and the distance between lamellae in the char
depend on the rank of the parent coal; as the coal rank increases the crystallite height
increases and the distance between lamellae decreases (van Krevelen, 1981; Marsh,
1989; Kabe et al., 2004). The average crystallite lattice parameters (L
c
and L
a
) can be
determined using empirical equation first derived by Scherrer (Kashiwaya and Ishii,
1991; Davis et al., 1995; Lu et al., 2001; Feng et al., 2003; Takagi et al., 2004; Trejo
et al., 2007; Van Niekerk, 2008):


002 002
cos .
K
L
c
= (2.1)

Chapter 2: Literature Review

29


10 10
cos .
K
L
a
= (2.2)

The interlayer distance between aromatic sheets (
002
d ) was calculated from the
maximum of the (002) band using the Braggs Equation (Lu et al., 2001; Takagi et al.,
2004; Van Niekerk, 2008):


002
002
2

Sin
d = (2.3)

The average number of crystallites in a stack
ave
N , was calculated using the relation
(Trejo et al., 2007; Van Niekerk, 2008):

1
002
+ =
d
L
N
c
ave
(2.4)

If it is assumed that the area under the (002) band the band (
002
A and

A ) is equal
to the number of aromatic carbon and aliphatic carbon atoms respectively, then the
aromaticity of the coal or char can be calculated using Equation 2.5 (Lu et al., 2001;
2002a and 2002b; Kawakami et al., 2006; Trejo et al., 2007; Van Niekerk, 2008):

A A
A
f
a
+
=
002
002
(2.5)

The fraction of amorphous carbon can also be calculated from the reduced X-ray
intensity of the d
002
peak of the sample using the relation originally proposed by
Franklin (1951) and used severally by Ergun and Tiensuu (1959), Lu et al. (2001,
2002a and 2002b). These are discussed in detail in Section 3.6.4.


( )
( ) s d N
s d N
s X
X I
ave
ave
A
A

002
2
002
2
2
sin
sin 0606 . 0
1
(2.6)
Chapter 2: Literature Review

30

2.5 Coal, Char and Gasification Reactivity
To be able to assess the suitability of coals for use in different gasification systems
and to aid in the design of reactors to suit specific coal types, a comprehensive
gasification model that can predict coal conversion levels under process conditions is
desired. The influence of the char gasification stage on the rate of conversion means
that it is a significant variable in the development of such a model.
Char conversion rates are a function of many variables: the more significant of these
are discussed in detail in later sections of this review. Some of these are physical
effects, and can be modelled adequately with a thorough understanding of the
gasification conditions and particle structure. Others relate more to the chemistry of
the char; consequently predicting such variables requires experimental data on which
models can be based.
The intrinsic reactivity of the char is such a property, and is the focus of much of the
work in this research. The intrinsic reactivity of char is defined by kinetic data
measured under conditions where chemical reactions alone control the conversion
rates. Data are usually normalised to some measure of the total surface area of the
char, so that they reflect the inherent nature of the carbon (and any influences of
mineral matter and other impurities). Such data are required in order to understand
conversion rates in processes that have some degree of chemical rate control,
regardless of the extent of pore diffusion limitations. This is particularly relevant for
the slower char-CO
2
and char-H
2
O reactions under gasification conditions. Chemical
kinetics may also have some degree of controlling influence for combustion systems
at the lower temperatures found in fluidised bed applications and for very fine
particles in PF systems. Intrinsic reactivity is characterised by kinetic parameters
determined from reaction rate measurements, both as measured (apparent or specific)
and normalised to the char surface area (intrinsic), measured under conditions where
chemical processes alone control reaction rates. This usually necessitates the use of
relatively low temperatures: for char-O
2
reactions, this is often below 500C while
char-CO
2
and char-H
2
O reactions are usually measured at between 800 and 1100 C
(Dutta et al., 1977; Radovic et al., 1985; Matsui et al., 1987; Harris and Smith, 1990;
Fu and Wang, 2001; Ochoa et al., 2001; Sharma et al., 2002).
Chapter 2: Literature Review

31

The conditions under which intrinsic data are generated mean that they are more
readily obtained than data generated under simulated process conditions. Intrinsic data
are particularly useful in the modelling of the char gasification system. Combining
intrinsic reactivity data with the appropriate diffusion, mass transfer and heat transfer
calculations and char structural considerations will allow the development of a
predictive process model that can be applied to the sample as it reacts under a variety
of conditions. This is useful as operating parameters of different gasifiers vary greatly.
By understanding the factors that cause the intrinsic reactivity of a char to vary,
particularly under the high pressure conditions that prevail in gasification
applications, a more accurate predictive model is possible.


2.6 Factors Influencing Gasification Reactivity
In addition to the reaction conditions, the heterogeneous nature of coal means that
there are a number of other factors that will affect the conversion rates of the
subsequent chars. These are as follows:

Parent coal properties
Pyrolysis conditions
Chemical structure and composition
Changes in carbon crystallite properties
Catalysis by mineral matter
Physical structural properties of the chars
Different chars will be affected to different extents by the factors listed above; the
magnitude of these effects will depend on the parent coal properties as well as on the
conditions under which it is reacting. For example, properties that affect the chemical
rate of reaction but not the rate of diffusion, such as catalysis and available surface
area, will be more important at lower temperatures than at higher temperatures, when
chemical processes have more influence on the rate of conversion of the char.
Chapter 2: Literature Review

32

Conversely, char porosity (which influences the rate of diffusion through the pores of
the sample) will not have a large effect at low temperatures but become quite
significant at higher temperatures.


2.6.1 Properties of the Parent Coal
The properties of parent coal have long been known to have some effect on the
reactivity of the char produced after pyrolysis. Some of these properties that need
consideration during gasification include: carbon content, volatile matter content,
mineral matter content, and the petrographic properties of the parent coal.


2.6.1.1 Volatile Matter Content
The volatile matter content of coal depends on the maceral present. Cai et al. (1998)
investigated some South African coals and reported that volatile matter content
decreases in the order: liptinite > vitrinite > inertinite; and influences both the quality
and quantity of chars formed. The evolution and combustion of volatiles during
devolatilisation affects the char porosity and hence overall reactivity and can cause
fragmentation of the coal particles (Kaitano, 2007). It is generally observed that the
conversion rate during gasification is less for chars that are produced from high-
ranked coals. In practical and pilot-scale systems, this has usually been associated
with the lower volatile content of the higher-rank coal leaving a larger amount of char
to be converted. The evolution rate of volatile matter in coal also depends on the
particle temperature, operating pressure and composition of the volatiles as noted by
Khairil et al. (2001).


2.6.1.2 Fixed Carbon Content
There is some evidence (Jenkins et al., 1973; Miura et al., 1989; de la Puente et al.,
2000) to suggest that the carbon in chars derived from high-rank coals is intrinsically
less reactive than carbon in chars made from low-ranked coals. This can be due to the
presence of more catalytically active mineral constituents in lower-ranked coals
Chapter 2: Literature Review

33

(Miura et al., 1989) or the difference in the fundamental structure of the carbon: low-
rank coal chars have a structure with a higher degree of disorder which provide more
active sites for reaction; whereas the higher-rank coal chars have less surface area
and a more ordered carbon lattice structure (Laurendeau, 1978) which may lead to
less sites on which reaction can take place. Thus higher rank coals exhibit higher
aromaticity and contains a lower fraction of carbon that is amorphous (Davis et al.,
1995; Lu et al., 2001).


2.6.1.3 Petrographic Properties of Coal
There has been much research on the effects of the petrographic composition of the
parent coals on the reactivity of the subsequent chars (Cai et al., 1998; Megaritis et
al., 1999; Alonso et al., 1999; Cloke and Lester, 1994). While there is experimental
evidence that chars derived from inertinite-rich coals exhibit longer burning times
than chars derived from coals that have a higher amount of vitrinite (Jones et al.,
1985; Sun et al., 2004; Kaitano, 2007; Everson et al., 2008a and 2008b), it is
suggested that these differences are primarily due to the effects of the coal
petrographic composition on the pyrolysis behaviour and the resultant char structure
as demonstrated by Cai et al. (1998). However, Cloke and Lester (1994) noted that
chars formed from inertinites (reactive- semifusinite and inertodetrinites) and
micrinites associated with liptinites are indeed, as reactive as vitrinites. These suffixes
to say that, not all inertinites are inert. Low reflecting inertinites are also more
reactive than higher reflecting inertinites (Cloke and Lester, 1994); which suggests
that inertinites from low ranked coals should be more reactive that their counterparts
from higher ranked coals.

During pyrolysis, vitrinites readily soften and release volatile material (Tsai and
Scaroni, 1987; Cloke and Lester, 1994; du Cann, 2008). This leads to chars with large
pores (resulting in particles with a low carbon density) and a relatively low (yet
accessible) surface area. Such morphology leads to particles that exhibit fast
conversion rates under combustion and gasification conditions. Inertinites, conversely,
have a tendency to not soften or release large amounts of volatile materials during
pyrolysis (du Cann, 2008; Everson et al., 2008b). Consequently, they are relatively
carbon-rich and retain more of the parent coals microporosity.
Chapter 2: Literature Review

34

The surface area of char made from inertinites is relatively high with a rather
uncorresponding increase in char porosity (Friesen and Ogunsola, 1995; du Cann,
2008). Under process conditions, these pores are also inaccessible, leading to
conversion rates that are low compared to chars made from vitrinite-rich coals. Bailey
et al. (1990), Benfell et al. (2000) and Yu et al. (2004) have grouped the chars that
result from the pyrolysis of coals rich in different macerals into 3 broad types: Group I
(very porous, thin shelled) through group II to group III (dense, with limited porosity).
Everson and co-workers (2008b) and du Cann (2008) proposed a classification system
for chars grouping the different resultant char carbon forms into seven broad groups
(Groups A to F) and further delineated each broad group into different categories. A
detailed discussion on these char carbon forms is presented in Section 3.7.6.


2.6.2 Pyrolysis Conditions and Heat Treatment
As well as determining the amount of char to be converted, the conditions of pyrolysis
can affect the structure and morphology of the resultant char (Cloke and Lester,
1994). This in turn can have influences on the reactivity characteristics of the char
during gasification reactions. Heat treatment of coals, the heating rates involved and
the pressures at which it is done are all important issues that need to be considered
when examining char reactivity.
When coal is heated to temperatures in the range of 400 - 900 C in an inert
atmosphere (pyrolysis), it decomposes into a hydrogen rich fraction (the volatiles),
and a carbon rich solid residue (the char) (Laurendeau, 1978; Kristiansen, 1996). The
pyrolysis reaction can be written as follows:

Gases Tars Char Coal
Heat
+ + (2.7)

The hydrogen rich fraction consists of tars and gases. The resultant char mainly
consists of fixed carbon and inorganic mineral matter (ash). The reactivity and the
structure of the char depend on the coal type, the temperature of pyrolysis, the heating
rate, the particle size, the residence time or holding time of the coal particle in the
Chapter 2: Literature Review

35

reactor, the pressure and the composition of the gas during pyrolysis. During
pyrolysis, the microporosity (the volume fraction of pores with size <2 nm) largely
increases (Laurendeau, 1978; Tsai, 1982). This is due to the volatiles released from
the previously inaccessible pores and restructuring of the aromatic clusters in the char
(Jones et al., 1999). The surface areas of mesopores (pore size, 2-50 nm) and
macropores (pore size > 50 nm) increases as well, probably due to gasification (steam
and carbon dioxide are produced in the pyrolysis process). Opening and coalescence
of micropores (pore size, <2 nm) may also lead to more mesopores.
It is generally agreed that the production of the volatiles takes place by the following
steps (Kristiansen, 1996):

i. Depolymerisation of the coal macro molecule. These results in the metaplast;
the fragments of the macro molecule.
ii. Repolymerisation of the metaplast molecules.
iii. Internal transport of molecules towards the surface and transport from the
surface to the bulk gas.

Mass transport inside the particle takes place by convection and diffusion in pores
when coal is non-softening. When the coal becomes plastic, the internal transport
takes place by liquid-phase and bubble transport. Plasticity occurs in coals with
carbon content between 81-92 % (Kristiansen, 1996), but this also depends on oxygen
and hydrogen content, heating rate and particle size. When the heating rate is high, the
repolymerisation reaction of the metaplast occurs very soon after the
depolymerisation reaction and the coal cannot plasticise.
It is well known that chars that have endured higher temperatures during pyrolysis
will generally exhibit a lower gasification reactivity, indicated by slow conversion
rates leading to longer burnout times (Jenkins et al., 1973; van Heek and Mhlen,
1985; Davis et al., 1995). Pyrolysis and post-devolatilisation heat treatment of coals
reduces the number of heteroatoms and crystalline irregularities leading to increased
aromaticity and crystalline carbon in the char, both of which play a role in
Chapter 2: Literature Review

36

determining gasification reactivity (Russell et al., 1999). Shim and Hurt 2000
investigated the effect of heat treatment on low temperature reactivity of char to O
2
.
Their findings correlated well with published results; the reactivity was observed to
decrease with increasing heat treatment temperature, which was attributed to the
annealing and structural ordering of the carbon in the char. The annealing was found
to be very fast, such that the process could be assumed to occur simultaneously with
char conversion. Wu et al. (2008) made similar findings in their CO
2
gasification of
chars annealed up to 1400 C.
It is not only the final temperature reached during pyrolysis, but the residence time or
holding time in a reactor at that temperature and the heating rate experienced by the
coal particle during conversion to char that are important. Van Heek and Mhlen
(1987) investigated the characteristic temperature at which a coal undergoing
pyrolysis will drastically lose reactivity. Radovic et al. (1983), also investigated the
reactivity of chars that were subjected to different residence times at a temperature of
1000 C. The reactivity of the chars dropped after residence times of a few seconds,
reached a minimum with a residence time of 5 minutes and could not be reduced
further, even at a residence times of up to 1 hour. This leads to the concept of a char
being dead burned; meaning that the pyrolysis temperature caused the reactivity of
the char to be substantially reduced and stabilised, in a very short residence time. This
reduction in reactivity is attributed to the ordering of the carbon lattice in the char at
higher temperatures, resulting in loss of potential number of active sites on which
reaction can occur.
The heating rate of coal during the pyrolysis stage will also have effects on the
reactivity of the resulting char. Hindmarsh et al. (1995) and Gale et al. (1996) studied
the effects of heating rate on the low-temperature reactivity of the resulting char.
Their results show a decrease in reactivity with both heating rate and pyrolysis
temperature. An increase in both of these parameters results in greater volatiles
release from the particles, to which they attributed the loss in subsequent char
reactivity. Different coals are affected to different extents by heat treatment. Some
coals undergo drastic changes of the chemical and physical properties during
pyrolysis. For some coals with strong coking properties, this may include the
formation of plastic intermediate phases. Coals that soften and also release large
Chapter 2: Literature Review

37

amounts of volatile matter will form chars that have structural properties quite
different to the parent coal. On the other hand, coals such as anthracites have a solid
structure that is less affected by carbonisation (Van Heek and Mhlen, 1987), and
therefore are less likely to be affected by a variation in pyrolysis conditions.
The pressure at which the parent coal is devolatilised also plays an important role in
the reactivity of the resulting char. Sha et al. (1990a) noted a significant decrease in
the reactivity of the char as the pyrolysis pressure was increased, postulating effects
on the pore structure as the cause. Two reasons for this are postulated: the
compression of the pyrolysis products failing to open the structure of the forming
char, and the other being an indirect temperature effect. Such a temperature effect
could arise from the exothermic hydrogasification reaction that may be occurring on
the surface of the charring particle, following or simultaneous with the hydropyrolysis
step. Cai et al. (1996) also found reactivities to decrease with hydropyrolysis
pressures up to 40 atm pyrolysis pressure and increased at pressures above this. The
eventual increase in reactivity was the result of some char conversion by
2
H at the
higher pressures exposing a fresh and enlarged carbon surface.
Lee et al. (1992) investigated the structure and reactivity of Illinois #6 coal char
following pyrolysis at elevated pressures. They found that increasing the pyrolysis
pressure slowed the rate of release of volatiles, increased the amount of char
remaining after pyrolysis and altered the composition of the volatile products. Their
data also demonstrated how pressure hinders the development of the mesopore system
that develops after the coal passes through the plastic phase of pyrolysis.
The increased fluidity that resulted from higher-pressure pyrolysis led to enhanced
ordering of carbon layers and the subsequent loss of gasification reactivity in the char
residue. An interesting effect of pressure on the development of char structure during
pyrolysis has been reported by Benfell et al. (2000), and Wu et al. (2000). Coal was
devolatilised over a range of pressures and various structural parameters of the
resulting char were analysed. They observed that as the pyrolysis pressure increases,
the proportion of morphological Group I chars- the open structured, porous group-
increases with a decrease in the proportion of Group II and III chars.
Chapter 2: Literature Review

38

2.6.3 Chemical Structure and Composition of Coal and Char
The complex coal and char structures (Figures 2.2, 2.3 and 2.5) requires analysis in
terms of their characteristic functional groups (Laurendeau, 1978). These fundamental
groups typifying the atomic species: C, H, O, and S, are summarised in Table 2.3.

Table 2.3: Basic structures and functional groups in coal (Laurendeau, 1978; Lahaye,
1998)



A typical coal structure consists of an aromatic / hydroaromatic cluster- average of 4
to 10 rings per cluster loosely joined together by methylene, ether and sulphide
Atomic Specie Fundamental group Typical example

Chapter 2: Literature Review

39

linkages, 1 to 3 carbon atoms in length (Laurendeau, 1978). Given (1960) reported
that the loose aliphatic groups allow cross-linking between clusters on different planes
and hence the development of an extensive pore structure. He also noted that the
aliphatic, hydroaromatic and heterocyclic bonds are quite susceptible to bond
breakage at pyrolysis temperature. This culminates to chars with less aliphatic and
more aromatic carbon.
Chars are characterized by carbon-rich, poly-nuclear aromatic structures. Edge carbon
atoms are at least one order of magnitude more reactive than the basal carbon atoms
(Walker and Hippo, 1975). Increased reactivity at carbon edges is presumed to be due
to the availability of unpaired electrons which are available to form bonds with
chemisorbed species (Laine et al., 1963). Furthermore, impurities which may catalyse
some carbon reactions tend to diffuse and concentrate at crystallite edges (Walker and
Hippo, 1975).
Enhanced activity found at defects in the char structure (such as vacancies) is
probably due to geometric or charge imbalances (Thomas and Thomas, 1967).
Laurendeau (1978) reported that oxygen and hydrogen sites on these functional
groups also promote char reactivity since chemisorption on non-aromatic sites is
usually favoured to aromatic sites noting that oxygen sites are thought to influence
reactivity via electron exchanges. Hydrogen sites are presumed to increase char
reactivity by preferential oxidation (Jenkins et al., 1973; Laurendeau, 1978). Nitrogen
and sulphur could promote ring structure attack since electrons are most available
at heterocyclic sites.


2.6.4 Changes in Carbon Crystallite Properties
The greater percentage of the organic portion of coal-derived char is composed of an
imperfect arrangement of aromatic layers of varying in sizes (Davis et al., 1995). The
effects of the changes that occur in these crystallites during devolatilisation and
gasification have been studied by various investigators (Sha et al., 1990(b);
Kashiwaya and Ishii, 1991; Davis et al., 1995; Sharma et al., 1999 and 2002; Feng et
al., 2003; Takagi et al., 2004; Arenillas et al., 2004). Chemical properties of chars
Chapter 2: Literature Review

40

most affected by these processes are the aromaticity and the fraction of amorphous
carbon. Physical and structural lattice parameters as inter-layer spacing, crystallite
height and diameter are also affected.
Char reactivity generally becomes low as the conversion increases. This has been
attributed to the microstructural and microtextural changes that occur in the char
crystallite and / or loss of catalytic activity of the minerals and inorganic matter. The
size and ordering of these layers can be related to the reactivity of the carbon present
in three fundamental ways (Davis et al., 1995; Russell et al., 1999; Sharma et al.,
2002; Arenillas et al., 2004).
First, as the crystallite size changes, the amount of hydrogen and the ratio of reactive
edge sites to unreactive interior sites decrease: carbon aromaticity thus increases.
Second, as the individual layers approach a more perfectly graphitic arrangement of
aromatic units, active sites associated with defects, heteroatoms (non-sp
2
-hybridized
or carbons) are lost and thus, the fraction of amorphous carbon decreases.
Third, as the aromatic layers reach a more compact stacking arrangement, the
microscopically available surface area decreases. Although the relationship between
reactivity and crystallite ordering is more complex than this simple scheme suggests,
reactivity in general, decreases with increasing degree of order across the spectrum of
carbon materials (Kashiwaya and Ishii, 1991; Davis et al., 1995; Russell et al., 1999;
Sharma et al., 2002; Feng et al., 2003; Arenillas et al., 2004; Wu et al., 2008). The
increasing degree of orderliness usually results to increase in aromaticity and a
corresponding increase in the fraction of crystalline carbon; while the fraction of
amorphous carbon decrease (Franklin, 1951; Hirsch, 1954; Schoening, 1983; Davis et
al., 1995; Lu et al., 2001, 2002a and 2002b; Maity and Mukherjee, 2006; Wu et al.,
2008).
The reduced reactivity of subsequent chars may also be explained by either a lower
surface area of carbon available for reaction, or lower intrinsic reactivity (the reaction
rate per unit area of pore surface in the absence of any mass transfer restriction), or
due to a lower number of active carbon sites (Smith, 1978). As the most reactive
carbon atoms are located at the edges of the lamellae not on the layer planes (basal
Chapter 2: Literature Review

41

planes), the density of accessible layer edges depends on the molecular orientation
domain (MOD) size. Therefore, the smaller the MOD size, the greater the free edge
density (Rouzaud et al., 1988). Essenhigh (1981) and Tomita (2001) however noted
that, reactivities at the carbon active sites located on different edges such as armchair
and zig-zag are different.
Kashiwaya and Ishii (1991) designed an experiment to observe the difference in
reactivity of carbon atoms located on the basal plane and the edges of the
polyaromatic layers. They measured the aromatic layer stacking height, L
c
, and
average crystallite diameter, L
a
, of a metallurgical coke at different temperatures
under inert gas (Ar) and reaction gas mixtures of Ar-CO-CO
2
. They found out that
crystallite stacking height L
c
, was affected only by temperature with no significant
difference between the L
c
, of both the annealed and reacted cokes, implying that the
reaction is very slow on the basal plane. The crystallite diameter L
a
, of both annealed
and reacted cokes also showed an increase with temperature, but the reacted coke had
a lower L
a
, than the annealed coke at similar temperatures, indicating that the reaction
occurs at the edges of the polyaromatic layers.
Sharma and co-workers (1999 and 2002) studied the structure of coal char during low-
temperature gasification using High-Resolution Transmission Electron Microscopy
(HRTEM). Different coal chars show different behavior in the change of ordering with
conversion. Chars from Pocahontas #3 low volatile bituminous coal contained both
ordered and disordered crystallites at the initial stage, and the ordered crystallites were
dominant after gasification (dry ash-free basis). Beulah-zap lignite chars mainly
consisted of a higher fraction of amorphous carbon, and the crystallite ordering
increased upon gasification. On the other hand, chars from high volatile bituminous
coal, Illinois #6, and all the demineralised chars from these coals, did not show a large
change before and after gasification.
Feng et al. (2003) determined the crystallite size of a char sample during reaction with
carbon dioxide at constant temperature (800 C), up to 90% carbon conversion.
Crystallite height, L
c
, did not change significantly below approximately 60% carbon
conversion but it decreased at greater conversion levels, whereas crystallite diameter
L
a
, decreased during gasification, even at an early stage. They assumed that initially
the reaction occurs predominantly at the edges of the polyaromatic layers, which
Chapter 2: Literature Review

42

implies a decrease of the crystallite diameter, and only at later stages of reaction that
the entire polyaromatic layers were consumed. This correlates to the findings of
Kashiwaya and Ishii (1991).
The aromaticity of chars has been found to increase while, the fraction of amorphous
carbon decrease, during conversion processes. In the investigations of Lu et al. (2001,
2002a and 2002b), chars produced at between 900 C and 1500 C were characterized
accordingly for various crystallite properties and subjected to combustion reactivity
tests. Their findings on the chars are similar to published results, showing an
increasing aromaticity and a decreasing fraction of amorphous carbon with pyrolysis
temperature. Their reactivity results also show that increasing aromaticity and
decreasing fraction of amorphous carbon impacts negatively on reactivity.
These are further corroborated by the recent findings of Wu et al. (2008) on CO
2

gasification of chars prepared at the different temperatures (950 C to 1500 C). Their
characterisation results show that the fraction of amorphous carbon decreases with
increasing charring temperature while the reactivity result shows that chars prepared
at 950 C with higher fraction of amorphous carbon are more reactive than chars
prepared at higher temperatures. They also found that reactivity increases with
increasing inter-layer spacing. Chars pyrolysed at 950 C has a higher inter-layer
spacing value and were more reactive when compared with their counterparts
devolatilised at 1200 C and 1400 C.


2.6.5 Catalysis by Mineral Matter
Coals contain a variety of inorganic mineral constituents, which are left behind as ash
after coal conversion processes. Inorganic impurities in coal occur in two forms-
mineral matter and trace elements. Mineral matter in raw coal consists of four major
types:
i. Alumino-silicates (clays), such as kaolinite- Al
3
Si
2
O
5
(OH)
4
and illite-
KAl
3
Si
3
O
10
(OH)
2
.
Chapter 2: Literature Review

43

ii. Oxides, such as lime- CaO, quartz- SiO
2
and hematite- Fe
2
O
3

iii. Carbonates, such as calcite- CaCO
3
, siderite- FeCO
3
, dolomite-
CaCO
3
MgCO
3
and ankarite- 2CaCO
3
MgCO
3
FeCO
3
.
iv. Sulphides and sulphates, such as pyrite- FeS
2
and gypsum- CaSO
4
2H
2
O.
Typically, mineral matter is randomly distributed in coal as inclusions about 2 micron
in diameter (Wigley et al., 1997). During pyrolysis, gasification or combustion,
mineral matter undergoes transformation and changes causing the mineral phases in
the chars to be entirely different to that in the precursor coal (Van Alphen, 2009).
Some 20 to 30 trace metals are also distributed throughout the coal structure; which
are either organically bound to the coal molecule such as boron, bonded inorganically
to mineral matter (for example zirconium, manganese), or occur in both organic and
inorganic forms such as copper. Typical concentrations of trace metals in coal are 5 to
500 ppm, although some elements such as- B, Ba, Sr, Cu, Mn, Sn and Zr often appear
at the 500 to 1000 ppm level (Lauredeau, 1978).
Although Tomita (2001) and Lee (2007) comfirmed that the mechanism of catalysis
by mineral matter in the gasification systems is not straightforward, Walker, et al.
(1968), had reported that catalysts might affect gasification in several ways:

i. It may affect both steps of the oxygen-exchange mechanism of the gasification
reaction by either increasing the number of active sites or lowering the overall
activation energy of the reaction;
ii. It may induce pits in the carbon basal plane and expose additional edge planes
for reaction;
iii. It may also bypass the oxygen-exchange mechanism of the gasification
reaction completely.

Mineral matter and trace metals can provide direct catalytic activity particularly
calcium, iron, and manganese compounds (Tomita, 2001; Lee, 2007). Laine et al.
(1963) reported findings that as little as 100 ppm Fe, can increase carbon-CO
2
gasification reactivity by a factor of about 150. Kayembe and Pulsifer (1976) found
Chapter 2: Literature Review

44

that 10% by weight of K
2
CO
3
reduced the activation energy of the steam gasification
of Bear coal chars from 254 kJmol
-1
to 145 kJmol
-1
at atmospheric pressure and at
temperatures between 600-850 C.
Different metals show different catalytic influences during coal gasification. Ye et al.
(1998) ranked four of the more prominent catalytically active materials as: Na > K >
Ca > Ni, and found that their effect on reactivity is not additive, since the expected
effect of the combined activity of several cations was less than the sum of their
individual effects. They also reported that the activity of Ca in particular was
dependent on the form it took in the char structure: Ca
2+
cations chemically
associated with the carboxyl groups are active, whereas cations that are not
chemically associated have only a slight effect on the reactivity.
Httinger and Nattermann (1994) investigated the rates of steam and carbon dioxide
gasification as a function of burn-off, and observed, particularly the catalytic effects
of the inorganic mineral constituents Ca and Fe. They found, for a range of coal
ranks, that the average rate over the extent of burn-out did not correlate well with the
molar ratio of Ca in the char. They did find that Fe was three times more active than
the Ca, and that its effects were additive with Ca. This is particularly relevant for
chars derived from low-ranked coals as it has been reported by many investigators
that their reactivity is very dependent on the presence of mineral matter (Lauredeau,
1978, Miura et al., 1989; Huang et al., 1991; Czechowski and Kidawa, 1991; Kyotani
et al., 1993; Ye et al., 1998; Samaras et al., 1996; Httinger and Nattermann, 1994;
Tomita, 2001; Kabe et al., 2004; Lee, 2007).
Huang et al. (1991) prepared chars from vitrinite- and inertinite-rich fractions from a
low rank coal (34% volatile matter). The vitrinite-rich char was more reactive than the
inertinite-rich char but after demineralisation the former was less reactive. The surface
area of the micropores and mesopores of the inertinite-rich char was greater than that
of the vitrinite-rich char. They concluded that the catalytic effect of the mineral matter
on the reaction rate is greater than the surface area effect. Czechowski and Kidawa
(1991), made similar observation in their study. They assumed that the greater
concentration of the elements: Ca, K and Na in the ash of the vitrinite-rich char is
responsible for its higher reactivity compared to that of the inertinite-rich char.
Chapter 2: Literature Review

45

Other salts found to have catalytic effects during low temperature gasification were in
decreasing order of effectiveness: NaCO
3
> Li
2
CO
3
> KCl > NaCl > CuO (Miura et
al., 1989; Samaras et al., 1996; Httinger and Nattermann, 1994; Tomita, 2001; Lee,
2007).
Tomita (2001) and more recently, Lee (2007) noted that catalytic activity of these
inorganic subtances containing Na, Ca, K, Fe, etc., depend not only on the type, but
also on the gasification conditions and their degree of dispersion in the coal or char.
Kyotani et al. (1993), however observed that the catalytic effects of minerals are less
significant at higher temperatures. This might be expected because, at higher
temperatures, chemical processes and their catalytic effects have less influence on the
reaction rate.



2.6.6 Physical Structural Properties of Chars
2.6.6.1 Total Surface Area
Considering the adsorption-desorption mechanism which is widely accepted for
heterogeneous char-gas reactions (Laine et al., 1963; Thomas and Thomas, 1967;
Walker, et al., 1968; Radovic et al., 1983; Kyotani, et al., 1988; Tomita, 2001,
Arenillas et al., 2004), it would be expected that the more carbon surface that a
particular sample has, the faster the reactions would proceed. Attempts to correlate the
observed apparent or specific reactivities to the total surface area of the char
sample have been met with mixed success. Reports of two chars with equal total
surface areas having different reactivities, and two chars with similar reactivities
having different surface areas exist (Radovic et al., 1985). Nonetheless, the
normalisation of measured reactivities to the total surface area of the char can help to
differentiate between physical and chemical effects on the reaction rates. The
objective of such an exercise is not to use surface area as an explanation for the
kinetic behaviour of the char, but to attempt to remove the influences of the physical
structure of the char and leave behind the effects of the intrinsic chemical properties
of the sample.
Chapter 2: Literature Review

46

Total surface area is commonly measured using two techniques: adsorption of
nitrogen gas onto the surface of the sample at 77 K, and adsorption of carbon dioxide
onto the surface of the sample at 273 K, sometimes lower than 273 K (Anderson et al.,
1965; Walker et al., 1988; Micromeritics, 2006). Often these two techniques will give
different results; the interpretation of these results can provide a lot of information
regarding the structure of the char. In chars with an extensive micropore network
(generally < 2 nm), adsorption of nitrogen does not always reveal the surface area
attributed to these micropores (Anderson et al., 1965; Walker et al., 1988). As the
2
CO adsorption experiment is performed at higher temperatures (usually 273 K as
opposed to 77 K) the
2
CO molecules have a higher average kinetic energy. This
increased average energy means that the gas molecules are more likely to penetrate
the smaller pores. Thus samples with high
2
CO areas and low
2
N areas indicate an
extensive micropore network, while chars with similar
2
N and
2
CO surface areas are
believed to have larger pores (Walker et al., 1988).


2.6.6.2 Active Surface Area
The fact that reactivity does not always correlate well with the total surface area of a
char led some researchers to investigate the proportion of this total surface that was
actually available for chemical reaction (Laine et al., 1963; Radovic et al., 1983;
Arenillas et al., 2004). This so-called active surface area (ASA) is attributed to
surface defects, heteroatoms (ring structures containing O, N, etc.), mineral matter and
disordered carbon atoms or carbon edges. It rarely includes the entire surface of the
char. Reliable measurement of the ASA is difficult, and this is reflected in the lack of
consistency between different researchers results and a lack of reproducibility of
individual techniques. ASA is usually determined using temperature programmed
desorption (TPD), according to the following procedures (Laine et al., 1963; Radovic
et al., 1983; Kyotani et al., 1988; Arenillas et al., 2004): Oxygen is adsorbed onto the
surface of the char at low temperatures (200 to 300 C) and allowed to equilibrate.
This equilibration time varies and can take up to 12 hours. The sample is then heated
in an inert environment at a known and constant rate and the desorbing surface
Chapter 2: Literature Review

47

complexes are measured as the carbon-containing gases CO and CO
2
. By assuming
that the oxygen chemisorbed to all the active sites is available for reaction, the amount
of desorbed product is believed to be proportional to the active surface area.
This technique has been met with limited practical success (Lizzio et al., 1990;
Lahaye, 1998). The major problem with the process appears to be the time allowed for
adsorption of oxygen. Though, an equilibrium between the char and the O
2
is required,
it is often not reached within a practical time-frame and consequently an arbitrary
adsorption period is allocated. This leads to inconsistencies between investigators.
The concept of active sites and more importantly the concentration of activated
surface complexes present during the reaction is a fundamental issue behind the
understanding of the behaviour of the gasification reactions over a wide range of
conditions. The ability to measure the amount of reaction intermediates or surface
complexes at any stage of gasification would therefore be an extremely useful tool
(Httinger, 1990; Httinger and Nill, 1990).


2.6.6.3 Surface Complex Concentration during Reaction
Investigations into char-
2
CO gasification reactions have measured the amount of
surface complexes present on the surface of the char during gasification (Httinger,
1990; Httinger and Nill, 1990; Khl et al., 1992). During these investigations, the
sample was gasified to a pre-determined level of conversion and quenched rapidly,
leaving the adsorbed surface complexes on the surface of the char. By heating the
sample slowly in an inert gas environment, the amount of desorbed carbon products
can be measured and used to estimate the amount of adsorbed surface complexes.
This is a very useful approach for the analysis of the processes involved in the
gasification reactions. It is not a means by which an active surface area can be
measured for the purposes of characterising a char because of the nature of the
measurement. This is due to the fact that the number of surface complexes present
during reaction is dependent on the reaction conditions of temperature and reactant
partial pressure. Thus, the so-called reactive surface area will be a function of the
reaction conditions.
Chapter 2: Literature Review

48

2.7 Methods of Measuring Gasification Reactivity
There are several methods by which the reactivity parameters that apply to char
gasification can be measured. These vary in their methods of particle suspension and
the conditions under which the sample reacts. It is important to match the desired
outcomes of a set of experiments to the type of apparatus used in order to ensure that
the data generated are appropriate, as different approaches are more suited to a
particular task than others. Many methods are available both at laboratory scale, pilot
scale and test-rig facilities for the measurement of reactivity of coal and char during
gasification. These include: entrained flow reactors (EFR) (Chitsora et al., 1987;
Hampartsoumian et al., 1993); fluidised bed gasifier (Httinger and Nattermann,
1994); drop tube furnace (DTF) simulating entrained flow gasification (Megaritis et
al., 1999; Liu et al., 2000; Kajitani et al., 2006); horizontal tube reactors (HTR)
(Harris and Smith, 1990, Lizzio et al., 1990; Sina et al., 2003); and
thermogravimetric analysers (TGA) (Dutta et al., 1977; Matsui et al., 1987; Mhlen et
al., 1985; Miura et al., 1990; Huang et al., 1991; Czechowski and Kidawa, 1991;
Ochoa et al., 2001; Kaitano, 2007; Everson et al., 2006 and 2008a) In this
investigation, a TGA was used.


2.7.1 Thermogravimetric Analysers
TGA is one of the more common techniques used for the analysis of heterogeneous
reaction kinetics. In investigations using TGAs, the char sample is held in a basket, in
a furnace, and the gasifying medium is flowed through the furnace where it
encounters the sample in the basket. The sample holder is attached to a sensitive
balance and the mass of the sample is monitored as a function of time and / or
temperature. Derivatives of the mass versus time and the reaction rate(s) can be
calculated. The details of individual TGA designs vary, with some having reference
sample holders serving to measure the sample temperatures and others have
thermocouples underneath the sample holder, both to aid in accurate temperature
change measurements.
Chapter 2: Literature Review

49

Some TGAs have a horizontally-orientated reaction tube; however, the majority are
vertical. Although, they are one of the commonest means by which attempts are made
to measure kinetics, TGAs are limited by the sensitivity of the balance mechanism
employed, the means by which the temperature of the sample is estimated, and the
ability to ensure intimate gas-sample contact. Since any physical contact with the
sample by the thermocouple will disturb the balance reading, the sample
thermocouple is usually situated a few millimetres below the basket, which introduces
some uncertainty in the sample temperature measurement. It needs to be confirmed
that the amount of sample used in the experiment is small enough so that the
thermodynamics of the reaction being studied make negligible contributions to the
discrepancies between the actual and the measured sample temperatures (Feng and
Bhatia, 2002; Patnaik, 2008).
TGAs have been used extensively in the past for studies of the kinetics of gas-solid
reactions (Dutta et al., 1977; Matsui et al., 1987; Mhlen et al., 1985; Mhlen and
Sulimma 1987; Miura et al., 1990; Huang et al., 1991; Czechowski and Kidawa,
1991; Ochoa et al., 2001; Kaitano, 2007; Everson et al., 2006 and 2008a). There are
pressurised TGA instruments commercially available, with the latest ones capable of
100 atm pressure and 1600 C sample temperature. Increasing the system pressure
does introduce problems, such as increased effects of drag, variations in buoyancy and
large heat transfer from the furnace walls to the gas. State of the art TGA designs and
the use of suitable control software is able to counter these effects, as well as being
able to provide smoothed and corrected differentiation of the raw TGA data ( Mhlen
and Sulimma, 1987).

In contrast to fixed-bed reactors, where the reaction rate can be directly measured
from the product gas concentration and the gas flow rates, TGAs produce an integral
of the rate data. These data are usually expressed as mass (or fractional mass loss)
versus time. Derivative calculations on the data generated by the TGA are required to
generate results that are representative of the reaction rate. Such calculations can
produce rate versus time or burn-out curves that have a poor signal to noise ratio.
Differentiation of corrected mass versus time data and plotting this function against
conversion gives us data that follow the reaction rate as the gasification of the sample
proceeds.
Chapter 2: Literature Review

50

The use of TGAs in the investigation of heterogeneous reaction kinetics needs to be
accompanied by a thorough understanding of the limitations of such a system. TGAs
are unsuited to the simulation of the extreme conditions that prevail in combustion
and gasification technologies. This is due primarily to the unreliability of the
measurement of the sample temperature at such high temperatures, but also has to do
with the static presentation of the sample and the associated problems of reactant
diffusion through the bed. Using TGAs under well-defined, well-controlled
conditions, however, produces data that are both applicable and reliable (Patnaik,
2008).
TGAs are particularly suited for the measurement of Regime I kinetics (chemical
reaction controlled kinetics), which necessitate the use of low sample temperatures,
and which by definition should be independent of the sample presentation such as
particle size, gas flow rates, and mass of sample. Such data can then be used as the
basis for kinetic models, or can be combined with structural data to estimate high-
temperature reactivities.


2.8 Char-CO
2
Gasification Reactions
This section of the review presents the mechanisms by which chars react with CO
2
.
The kinetics of these reactions, which have been reported in the literature, is also
discussed.


2.8.1 Char-CO
2
Reaction Mechanism
Char-CO
2
reactions occur via reactant adsorption, surface reaction and product
desorption. From an experimental point of view, CO
2
gasification reactions are easier
to study than combustion reactions as they are much slower and the chars do not have
the tendency to ignite, as the reactions are endothermic. A widely accepted
mechanism for the char-CO
2
reaction is the reversible oxygen-exchange mechanism
(Ergun and Mentser 1968; Turkdogan and Vinters, 1969; Httinger, 1990; Httinger
and Nill, 1990; Kashiwaya and Ishii, 1991; Hampartsoumian et al., 1993; Kristiansen,
1996; Lee, 2007):
Chapter 2: Literature Review

51

( ) CO O C CO C
f
+ +
2

2 1
, k k (2.8)

( )
f
C CO O C +
3
k (2.9)

It was observed by Turkdogan and Vinters (1969) and corroborated by Httinger and
Nill (1990), that the desorption reaction (Equation 2.9), is the rate-determining
reaction for the overall process, at least at atmospheric pressure. The adsorption
reaction influences the reaction rate by determining the number of adsorbed surface
complexes, C(O), through the forward reaction of Equation 2.8.
Experimental evidence has shown that carbon monoxide, the product of the CO
2

gasification reactions, inhibits the desorption reaction of Equation 2.9 (Ergun and
Mentser, 1968). This inhibition by carbon monoxide can be accounted for, by the
dynamic equilibrium shifts in the above oxygen-exchange mechanism of Equation
2.8. There are data to support this mechanism in the review by Ergun and Mentser
(1968) which has been confirmed by the results of Hampartsoumian et al. (1993).
This mechanism is currently the accepted mechanism for CO
2
gasification of
carbonaceous materials at atmospheric pressure. Hampartsoumian et al. (1993) also
suggest that CO inhibition in gasification reactions is less important at high
temperatures (up to 1527 C), suggesting that as the temperature increases the rate of
formation of CO decreases.


2.9 Heterogeneous Char-Gas Kinetics
2.9.1 Reaction Rate Models
There are two ways by which the variations of char gasification rates with temperature
and pressure can be represented. These are:

i. Rate equations based on stepwise reaction mechanisms (the so called
Langmuir-Hinshelwood equations, with stepwise rate constants that vary in
Arrhenius form with temperature)
Chapter 2: Literature Review

52


ii. An nth order approximation to this rate equation- the power rate law (an
overall Arrhenius-type representation).

By taking the reaction mechanisms proposed earlier (Section 2.8.1; Equations 2.8 and
2.9) and applying standard kinetic theory, the rates of each of these two reactions
would be:

( )
[ ] ( ) [ ]
CO CO f
y O C k y C k r
2 1 1
2
= (2.10)

( )
( ) [ ] O C k r
3 2
= (2.11)

Assuming that the rates of reaction:
( ) 1
r and
( ) 2
r are equal at equilibrium and given
that:

[ ] [ ] ( ) [ ] O C C C
t f
= (2.12)

an expression for the number of adsorbed surface complexes in terms of the gas
partial pressures and the total concentration of available carbon sites can be evaluated
using Equation 2.13:

( ) [ ]
[ ]
3 2 1
1
2
2
k y k y k
y k C
O C
CO CO
CO t
+ +
= (2.13)

The char-CO
2
reaction rate is believed to be determined by the rate of desorption of
surface complexes (Httinger and Nill, 1990). Therefore the reaction rate can be
written as:

( ) [ ] O C k r
CO 3
2
= (2.14)

and by substituting Equation 2.13 into Equation 2.14, the rate equation for the char-
2
CO reaction can be expressed as below (Blackwood and Ingeme, 1960; Smith et al.,
1991):
Chapter 2: Literature Review

53


[ ]
2
2
2
3
1
3
2
1
1
CO CO
CO t
CO
y
k
k
y
k
k
y k C
r
+ +
= (2.15)

Equation 2.15 is the Langmuir-Hinshelwood rate equation for atmospheric CO
2

gasification. There is the possibility that reactions not discussed here become
significant in the mechanism at high pressures (which is not within the scope of this
investigation).
By considering the reaction as an overall rate expression, an n
th
order approximation
to the rate equation can be used to describe the reaction rate. This is often called an
Arrhenius-type representation or the n
th
order rate equation (power rate law). Such an
expression is essentially empirical, relating the intrinsic reaction rate (which can be
the combination of several individual steps) in terms of a rate constant and the
concentration of reactant gases (expressed as concentration, partial pressure or mole
fraction) raised to an order, m (Smith et al., 1991):


m
CO s
ky r
2
= (2.16)

The rate constant, k, is exponentially dependent on the reaction temperature, T,
through the relationship:

|

\
|
=
RT
E
A k
a
exp (2.17)

Thus,

m
CO
a
s
y
RT
E
A r
2
exp |

\
|
= (2.18)

Many chemical reactions obey this type of relationship, and generally the gasification
reactions (steam and CO
2
) are no exception over a limited range of conditions (Dutta
et al., 1977; Laurendeau, 1978; Harris and Smith, 1990; Smith et al., 1991; Murillo et
al., 2004 and 2006). Kaitano (2007) reported that the power rate law is generally more
Chapter 2: Literature Review

54

applicable at lower partial pressures of gasifying agent. This is predominately due to
the inability of the values of the kinetic parameters to apply over extremes of pressure
and temperature conditions. At higher pressures, the Langmuir-Hinshelwood rate
equation is more applicable as it accounts for both the CO
2
partial pressure as well as
the formation of the product gas (CO).
It is likely, however, that in measuring such kinetic parameters over such a variety of
conditions, the influences of a number of other processes such as mass transfer, gas-
film diffusion, or product inhibition on the reaction rates are also measured. In order
to be able to quantify reaction rates at high temperatures and pressures, the influences
of pore and bulk diffusion must be accounted for, separate from the chemical
processes, either by mathematical estimation or experimental measurement.


2.9.2 Overall CO
2
Gasification Kinetics
The rate of the char-CO
2
reaction is several orders of magnitude slower than the char-
O
2
reaction (Harris and Smith, 1990). Due to these slower reaction rates, the processes
limiting gasification with CO
2
under process conditions are expected to be a
combination of chemistry and diffusion through the pores of the char, even at the high
temperatures experienced in practical gasification systems. There have been many
investigations into the reactivity to CO
2
of chars produced from a variety of coal types
both at atmospheric and higher pressures with reported activation energies ranging
from 100 to 280 MJmol
-1
(Dutta et al., 1977; Radovic et al., 1985; Matsui et al.,
1987; Miura et al., 1990; Ochoa et al., 2001; Everson et al., 2006and 2008a; Murillo
et al., 2006 Kajitani et al., 2006; Kaitano, 2007). Measured activation energies for the
char-CO
2
reaction are usually greater than that for char-O
2
reaction.
Overall reaction orders, however, are similar. The accepted value for the reaction
order for the char-
2
CO reaction is in the range 0.5 - 0.7 (Laurendeau, 1978; Harris
and Smith, 1990; Kaitano, 2007; Everson et al., 2008a), which is similar to the value
reported for the oxygen reactions. This may be expected because the surface
adsorption-desorption mechanism is similar for the oxygen and carbon dioxide
reactions (Radovic et al., 1983). There is some experimental evidence for this
provided by Httinger (1990); Httinger and Nill (1990); and Zhang and Calo (1996).
Chapter 2: Literature Review

55

They performed experiments at atmospheric pressure, in which the partial pressure of
2
CO was varied. The experiments were frozen at a known level of conversion, and
the sample was then heated at a known heating rate in an inert gas. The adsorbed
surface complexes, which remained on the surface when the reaction was stopped,
then desorbed and were able to be measured as evolved CO. They found that the
amount of adsorbed complex increased with the partial pressure of CO
2
. Httinger and
Nill (1990) however observed that the reaction rate normalised to the amount of
adsorbed complex is constant.


2.10 Homogeneous Gas-Phase Reactions
Although the heterogeneous char-gas reactions are important when considering the
burnout times for char during gasification, there are also homogeneous gas-phase
reactions that are important in determining the composition of the fuel or synthesis
gas produced in the gasification process. If there is oxygen in the gasifier, it
dominates the gas-phase reactions. These include:


2 2 2
1
CO O CO + (2.19)

O H O H
2 2 2
1
2
+ (2.20)

O H CO O CH
2 2 2 4
2 2 + + (2.21)

O H SO O S H
2 2 2 2
1
2
1 + + (2.22)

2 2 2 2
1
1 SO CO O COS + + (2.23)

Gas-phase reactions that occur after oxygen is removed are also important as they
influence the composition of the fuel / synthesis gas. In the presence of steam, the
water-gas shift reaction has a dominant influence on the resulting
2
H CO ratio:

2 2 2
H CO O H CO + + (2.24)
Chapter 2: Literature Review

56

And the methanation reaction (important at higher pressures) increases the calorific
value of the fuel gas (Lee, 2007):

O H CH H CO
2 4 2
3 + + (2.25)

Jntgen et al. (1981) also reported that methane can also react with steam as follows:

2 2 4
3H CO O H CH + + (2.26)

2 2 2 4
4 2 H CO O H CH + + (2.27)

which will affect the final ratio of hydrogen to methane. The sulphur species present
may also undergo the following homogeneous reactions with hydrogen and steam:

S H O H H SO
2 2 2 2
2 3 + + (2.28)

S H CO O H COS
2 2 2
+ + (2.29)

The formation of nitrogen oxides also occurs, but this is readily reduced (in-situ) so
that the form of nitrogen in the exit stream of a gasifier is principally molecular
nitrogen,
2
N with some
3
NH (Watkinson, 1991). The kinetics of these gas-phase
reactions are not discussed in this review. The focus is on the heterogeneous char-
2
CO reactions.


2.11 Structural Kinetic Models
The overall reaction rate for char-gas reactions depends on the intrinsic surface
reaction
s
r and the structure of the char particles (structure factor, ) (X f ), which are
determined by various factors including coal properties, devolatilisation conditions
and the reaction conditions (See section 2.6). This can be expressed as follows:

) ( ) , ( X f C T r
dt
dX
s
= (2.30)
Chapter 2: Literature Review

57

The surface of char particles is both physically and chemically heterogeneous. It is
expected, therefore, that different parts of the surface will behave differently under
varying conditions. Considering this, there is no single technique that can be used to
consistently remove all effects of the surface of the char from a reactivity
measurement. Thus, a measure of the true intrinsic reactivity of char is rather hard to
achieve. Instead, a value of the total surface area is introduced as a useful
measurement that can be used to get a reactivity parameter that is representative of the
average reactivity of the char particle.

As the reactions remove carbon from the surface of the pores, the surface area of the
char changes as conversion progresses. An understanding of how the coal char surface
evolves and, more importantly, how we can predict its behaviour is an important part
of the knowledge required for successful modelling of gasification reactions. To this
end, a number of models predicting the evolution of surface area have been
developed. Most of these models are based on the development of various pore
structures in the solid reactant in this case coal char. Some of these models
incorporating structural evolution during the reactions are: the volume reaction model
(Dutta et al., 1977; Murillo et al., 2004 and 2006; Wu et al., 2008), the shrinking core
model (Levenspiel, 1972; Bhatia and Perlmutter, 1980 and 1981; Mhlen et al., 1985;
Mhlen and Sulimma, 1987; Murillo et al., 2004; Everson et al., 2006; Zhang et al.,
2006), and the random pore model (Bhatia and Perlmutter, 1980 and 1981; Radovic et
al., 1985; Matsui et al., 1987; Miura et al., 1990; Hampartsoumian et al., 1993; Ochoa
et al., 2001; Sina et al., 2003; Kajitani et al., 2006; Zhang et al., 2006; Kaitano,
2007; Everson et al., 2008b; Sangtong-Ngam and Narasungha, 2008).


2.11.1 The Volume Reaction Model
The volume reaction model (VRM) is the simplest model that takes into account
structural changes that occur during the reactions and has been proposed by Dutta et
al. (1977) to describe char reaction with CO
2
(Murillo et al., 2004 and 2006; Wu et
al., 2008). The VRM assumes uniform gas diffusion in the entire solid particle and
simplifies the heterogeneous char-CO
2
reaction by assuming that the gas is reacting
homogeneously with the char. The kinetic expression for the reaction rate is (Dutta et
al., 1977; Murillo et al., 2004):
Chapter 2: Literature Review

58

) 1 ( ) ( X k X f
v
= (2.31)

and the reaction rate expression is given by (Bhatia and Perlmutter, 1980):

) 1 (
1
X
C k
dt
dX
o
g v

(2.32)

The limitations of the model are that it only predicts a monotonically decreasing
reaction rate whereas most gasification systems have been known to exhibit maximum
reaction rate.


2.11.2 Shrinking Core Model
The Shrinking core model (SCM) considers reaction(s) as occurring on the surface of
a shrinking non-reacted core within the solid particle (grain). The shrinking surface
can leave behind a layer of ash, which may offers resistance to the transport of
reactant and product gases. Under these conditions, it is assumed that there is
negligible penetration for the reaction in the core and that the reaction occurs within a
very thin layer on the surface of the shrinking particle surface. Levenspiel (1972)
assumed that a porous particle with negligible diffusion between individual grains
consists of an assembly of uniform non-porous grains and that the shrinking core
behaviour is confined to these grains. Most investigators, however, considered this
behaviour applicable to the whole particle and have been found to be effective in the
prediction of heterogeneous coal char-gas reaction kinetics (Levenspiel, 1972;
Everson et al., 2006). For a shrinking core model, with the surface chemical reaction
controlling and the particles assumed to be spherical, the structure factor is given
below (Mhlen et al., 1985; Mhlen and Sulimma, 1987; Murillo et al., 2004; Zhang
et al., 2006; Everson et al., 2006; Bhatia and Perlmutter, 1980 and 1981):

) 1 (
) 1 (
) (
3 / 2
o
o
X S
X f

= (2.33)

Chapter 2: Literature Review

59

And the overall reaction rate can be expressed as:


( )
3 / 2
0
0
) 1 (
1
X
S
r
dt
dX
S

(2.34)

It encompasses the initial structural properties of the char in terms of the initial
porosity and external surface area, but fails to predict a maximum behaviour observed
for some gasification experiment (Hashimoto et al., 1979). The SCM however, does
not account for structural changes (surface area and porosity) occurring during
reaction of the char.


2.11.3 Random Pore Model

The Random pore model (RPM), (Bhatia and Perlmutter, 1980 and 1981) considers a
transformation from non-overlapped cylindrical pores to overlapped ones during
gasification. It basically assumes that pores are of random sizes and orientations and
may adjust for intersections and can very well describe gasification systems that show
maximum reaction rate and those that do not. The models structural factor is given by
(Bhatia and Perlmutter, 1980):

( )
( ) ( )
0
0
1
1 ln 1 1


=
X X S
X f (2.35)

where,

( )
2
0
0 0
1 4
S
L


= (2.36)

Several investigators: Bhatia and Perlmutter (1980) and (1981); Kaitano (2007);
Everson et al. (2008b) on coal chars gasification and other heterogeneous reaction
kinetics have successfully used RPM and its various modifications in modeling their
gasification systems. Detailed discussion on RPM is presented in chapter 6.


60














CHAPTER 3 33 3





3.0 COAL AND CHAR CHARACTERISATION

3.1 Introduction
In this chapter, a detailed characterisation of the parent coal and subsequent char
samples are presented. These were done in order to relate the fundamental coal and
char properties to the CO
2
reactivity of the chars. Several techniques, both
conventional: proximate and ultimate analysis, total sulphur content and calorific
value determination; and advanced: XRD and XRF were used.
Chapter 3: Coal and char Characterisation

61

Detailed petrographic analysis (organic composition and reflectance properties) were
conducted on the samples, which provided a detailed maceral analysis on all the coal
samples; and char carbon forms, structural and textural analysis on the chars. Various
physical structural properties of the parent coal and char samples were also
determined using surface area and pore size analyser and helium pycnometry (HP).
The origin of the coal samples, methodologies and procedures of the various
characterisation techniques are also discussed.



3.2 Origin of Coal Samples
The four original coal samples used in this study, identified for confidentiality
purposes as coal samples; B, C, C2 and D2, were supplied by Eskom and originated
from mines in the Highveld area. The parent coal samples were received in particle
sizes ranging from fine powder to 5 cm.


3.3 Sample Preparation
Mechanical size reduction was employed to reduce the particle size of the coal
samples as received to the required size ranges for analyses and char preparation. For
each coal sample, 300 g of the +5 mm size fraction was separated and crushed with a
jaw-crusher (Samuel Osborne (SA) LTD, Model: 66YROLL) set at maximum jaw
opening initially to gradually reduce the particle size. The crushed sample was
screened with a 1.7 mm screen; then the jaw opening was decreased slightly and the
top-size coal particles from the 1.7 mm screen were passed through again. This
process was repeated until the entire 300 g sample passes through the 1.7 mm screen.
Further size reduction of the -1.7 mm particle size, was conducted using a Fritsch P-14
rotary mill (R-mill) with a mounted 280x230mm macro crusher (Model No. 46-126)
containing 10mm ceramic balls. The speed setting used was 600 rpm and allowed to
run for 30 minutes. Samples were screened as necessary to remove the required
particle size fraction and crushing continued. It should be noted that the mills were
Chapter 3: Coal and char Characterisation

62

cleaned properly with compressed air prior to use to avoid any contamination with
previous crushed samples. The different size fractions were then weighed out into 30
g samples which were flushed with nitrogen and appropriately sealed in vacuumed
bags to prevent oxidation.
The particle size requirement and screens used for size fractions of various analyses
and char production is given in Table 3.1.


Table 3.1: Size requirements for coal and char characterisation analyses.
Analysis technique Size requirement (m) Screen used (m)
Char production 1000 <dp<1120 1120;1000
Petrography
HP measurements
Gas adsorption
Proximate analysis 75<dp<212 212; 75
Ultimate analysis
Total sulphur analysis
CV determination
XRD mineral analysis <75 75
Dimeneralisation
XRD carbon fraction analysis




3.4 Char Preparation at 900 C
The already screened required size fraction of parent coals (1000 <dp<1120 m), was
used to produce the chars. This was done by the Coal Research Group, North-West
University.
Chapter 3: Coal and char Characterisation

63

3.4.1 Charring Apparatus and Procedure
The char production was conducted in a Packed Bed Balance Reactor (PBBR). The
schematic representation of the reactor is shown in Figure 3.1. The PBBR consists of
a vertical tubular furnace mounted on a very sensitive microbalance by an aluminum
tripod. Nitrogen gas was purged through the bottom of the reactor at a flow rate of
1500 cm
3
min
-1
measured by the help of a gas flow controller, and the temperature in
the furnace was controlled by a programmable heat controller attached to the furnace
unit. The furnace can be programmed to work both isothermally and non-isothermally
(ramp). The weight loss due to devolatilisation was observed via an online data
acquisition module with the attainment of a constant weight indicative of the
completion of the char production process.
This PBBR is dedicated to char production and is capable of handling up to 60 grams
of sample per batch. Char production in a separate and dedicated reactor prior to TGA
experiment was done in order to prevent the condensation of tars and volatiles on the
delicate TGA as noted by Kaitano, (2007). The char production conditions are given
in Table 3.2


Figure 3.1: Experimental setup for char preparation (Not drawn to scale) (Kaitano, 2007)
Table 3.2: Char production conditions.
Furnace
Reactor
Gas flow controller
Nitrogen gas
Balance
Computer
Thermocouple
Volatiles / flue gas
Module
N
2

Chapter 3: Coal and char Characterisation

64

Char production conditions
Mass of coal per batch 60 g (maximum)
Initial ramp 20 Cmin
-1

Final isothermal temperature 900 C
Holding time 70 min

The char production sequence from the parent coal samples are as follows: The coal
samples were sieved into the required 1mm particle size and put in an oven at 60 C
and for 1 hour so that the condensed moisture is driven off. The sample temperature
was equilibrated to ambient temperature and pressure in a flow of nitrogen.
It was then heated non-isothermally at 20 Cmin
-1
in a nitrogen atmosphere to the
target temperature of 900 C and then, held isothermally at the target temperature for
70 minutes. A final constant weight shows that all the volatiles have been driven off.
The furnace was turned off and the sample cooled down to ambient temperature and
pressure under nitrogen flow. After cooling down, the samples were analysed for char
yields and packaged in vacuumed plastic bags under nitrogen flow with the labels:
Char B, Char C, Char C2 and Char D2.
Further preparation was conducted on the char samples. Samples used for XRD
carbon crystallite and XRF ash components analyses were pulverized to -75 m;
demineralization was done on part of them as required by XRD carbon analysis. The
samples used for TGA experimentation were split to fractions of 50 1 mg using a
rotary sample splitter at a speed of 10 rpm. Fully prepared and or split char samples
were stored in a dessicator from which samples were taken for various analyses as
well as for experiments on the TGA.
The result of the char yield after devolatilisation of the precursor coals is presented in
Table 3.3. It can be seen from the char yield data that the moisture and volatile matter
content of the parent coal (based on proximate analysis of the coals (adb)) are almost
driven off completely. Discrepancies in the total percentage is probably due to the
presence of traces of volatiles still trapped in the char as was confirmed from the
proximate analysis (Table 3.8).
Table 3.3: Char yield after production.
Chapter 3: Coal and char Characterisation

65

Char Yield
Char ID B C C2 D2
Char yield (wt. %, adb) 73.5 75.6 78.9 72.4
Moisture and VM of parent coal (wt. %, adb) 27.3 24.8 21.1 28.1
Total (wt. %, adb) 100.8 100.4 100 100.5


3.5 Coal and Char Characterisation Analyses
The various characterisation analyses conducted on the four coal and four
intermediate char samples, including the laboratories they were conducted at, are
summarised in Table 3.4.


Table 3.4: Characterisation analyses conducted on the coal and char samples.
Analysis Property Laboratory Used
Chemical &
Mineralogical
Proximate Advanced Coal Technology / NWU
Ultimate Advanced Coal Technology
Calorific value Advanced Coal Technology
Total sulphur Advanced Coal Technology
Mineral (XRD) University of Pretoria
Ash (XRF) Council for Geosciences, Pretoria
Petrographic
Maceral composition Petrographics SA
Vitrinite reflectance Petrographics SA
Carbon-mineral Structure Petrographics SA
Char form Petrographics SA
Structural

Carbon crystallite (XRD) PANalytical XRD, Randburg
Gas adsorption (CO
2
) North-West University
HP measurements North-West University

Chapter 3: Coal and char Characterisation

66

3.6 Coal and Char Characterisation Equipment and Techniques
3.6.1 Chemical Analyses
Chemical analyses of the samples comprised of proximate and ultimate analyses,
calorific value and total sulphur content determination. The proximate analysis on all
the four char samples was done at the North-West University, while that of the parent
coals were done at the Laboratory of Advanced Coal Technology (ACT) Pretoria. The
ultimate and total sulphur analysis of all the samples was also done at ACT according
to the standard methods shown in Table 3.5 (du Cann, 2007 and 2008).



Table 3.5: Analytical methods used for chemical and mineralogical analysis.
Items Standard Method
Sample Preparation SABS 0135: Part 1 & 2. (1997)
Proximate Analysis
Inherent Moisture Content (%) SABS 925
Ash Content (%) SABS ISO 1171 (1997)
Volatile Matter Content (%) SABS ISO 562 (1998)
Fixed carbon Content (%) By Difference
Ultimate Analysis
Carbon (%) ISO 12902
Hydrogen (%) ISO 12902
Nitrogen (%) ISO 12902
Oxygen (%) By Difference
Total Sulphur Content (IR Spectroscopy) (%) ISO 19759
Calorific Value (MJ.Kg
-1
) SABS ISO 1928 (1995)
Grade (Based on CV, Air dry basis) CKS 561-1982
Mineralogical Analysis
Ash Analysis (XRF) ASTM 3682

Chapter 3: Coal and char Characterisation

67

3.6.2 X-ray Diffraction (XRD) Mineral Analysis
The X-ray diffraction (XRD) analysis was used to determine and quantify the
different minerals contained in the coal and char samples. This mineralogy analysis
was done at the Laboratory of the Geology Department, University of Pretoria
(courtesy of Dr. Sabine Verryn).
Already prepared coal samples (-75 m) was used for this analysis. Char samples
were further processed to get the required particle size. 40 g of the four char samples
(1 mm particles) were milled down to fine powder of -75m in a Fritsch P-14 rotary
ball mill with ceramic balls.
The powdered samples were transferred to suitable sample holders made of
aluminium alloy and tamped gently, but thoroughly with the edge of a glass slide. It
was important to fill the sample holders and the surplus sliced off with a glass slide
(approximately 50x70mm and 5 mm thick), whilst simultaneously compressing the
samples in the holders. The above procedures were repeated until smooth surfaces of
even texture were obtained.
Analysis on the prepared samples was done using a PANalytical XPert Pro powder
diffractometer with XCelerator detector and variable divergence- and receiving slits
with Fe filtered Co-K radiation. The analysis parameters and settings on XRD
system is shown on Table 3.6.

Table 3.6: Analysis parameters and settings on the XRD system for mineral analysis.
Anode material Cobalt target, : Co K=1.78901
Generator setting 40 kV; 40 mA
Angle of scan 2.0 2 135
Progressive divergence Slit 1.0 (fixed)
Anti-scatter slit 2.0
Soller slit 0.02 Radian
Scanning Continuous
Duration of scan 6- 8 hours


Chapter 3: Coal and char Characterisation

68

The crystalline phases of minerals present in the samples were identified using XPert
Highscore plus software. The relative phase amounts (weight %) was estimated using
the Rietveld method (Autoquan Program). Amorphous phases, if present, were not
taken into account in the quantification.


3.6.3 Ash Analysis (XRF)
In a bid to have more insight into the ash compositions of the high ash chars, major
element analysis using X-ray Fluorescence (XRF) spectroscopy was conducted on the
chars. This was done at the laboratory of the Council for Geosciences, Pretoria
according to ASTM 3682 standards.
The already micronised char samples (-75 m) were used for this analysis for better
results. The technique involves initial drying of the fine char samples at a temperature
of 110 C to a constant weight. The dried samples were then calcined in air at a
temperature of 500 C for one hour and at 815 C for four hours, in order to determine
the Loss on Ignition (LOI) value as well as drive off water and all organic components
and compounds originally contained in the char samples. The calcined sample was
converted into a solid solution by fusion with lithium tetraborate (Li
2
B
4
O
7
), LTB (one
gram of the calcined ash to nine grams of LTB).
The prepared solid solution and standard SARM-2, an international syenite certified
reference material from MINTEK were placed in the sample holders and put in the
sample compartment of the XRF spectrometer.
For quantification, the intensity of a characteristic line of element to be determined
was measured and concentration of the element in the sample was calculated from the
measured intensity (Loubser and Verryn, 2008; Matjie, 2008). Coal and char ashes
contains typically: Fe, Al, Mg, Mn, V, Ti, Si, Ca, Na, K, P, S and Cr, which are
reported as oxides by default: Fe
2
O
3
, Al
2
O
3
, MgO, MnO, V
2
O
3
, TiO
2
, SiO
2
, CaO,
Na
2
O, K
2
O, P
2
O
5
, SO
3
and Cr
2
O
3
(Loubser and Verryn, 2008; Matjie, 2008). Minor
and trace element analysis was not done on the samples. The ash chemistry of the
parent coal samples was also not determined.
Chapter 3: Coal and char Characterisation

69

3.6.4 X-ray Diffraction (XRD) Carbon Crystallite Analysis
XRD was also used to study the carbon crystallite properties of both the coal and char
samples. XRD analytical techniques have been established as a useful tool in
obtaining the structural information of coal and chars (Lu et al., 2001; Feng et al.,
2003; Gupta, 2007).
The pulverised (-75 m) coal and char samples for this investigation were
demineralised to reduce the amount of mineral matter present in them as well as to
minimise their influence during quantitative analysis (Lu et al., 2001, Maity and
Mukherjee, 2006; Van Niekerk, 2008).
The demineralisation was done according to the modified method adapted from
literature (Steel and Patrick, 2001; Lu et al., 2001; Maity and Mukherjee, 2006; Van
Niekerk, 2008). Twenty grams of each sample was added to 80 ml 5M concentrated
hydrochloric acid (HCl) in a glass beaker and agitated with a polyethylene coated
magnetic stirrer for 24 hours at 450 revolutions per minute (rpm). The acid was
separated from the sample by filtration and the precipitate was added to 80 ml 29M
(48 wt. %) hydrofluoric acid (HF) in a polyethylene beaker and agitated for 24 hours
with a polyethylene coated magnetic stirrer at 450 rpm. The acid was again separated
by filtration and the precipitate added again to 80 ml 5M hydrochloric acid for another
24 hours under the same conditions as the first HCl leaching. The acid was separated
by filtration and the precipitate was washed under a lower pressure atmosphere with
ultrapure water. In order to remove the adsorbed HCl from the residual sample
particles, large quantities of ultrapure water were used in the rinse stage until the pH
of the filtrate corresponded to 7.0. The remaining fully washed acid-treated samples
were dried in a vacuum oven at 60 C until a constant weight was obtained. All the
samples were flushed with nitrogen and sealed in plastic bags prior to XRD scans.
Proximate analysis was conducted on the demineralised samples to determine the
degree of removal of the minerals in the samples. The effectiveness of
demineralisation, E
d
, was introduced to quantify the extent of demineralisation. It is
the ratio of ash removed or demineralised to the original ash content of the sample
expressed in percentage:
Chapter 3: Coal and char Characterisation

70


100
|
|

\
|
=
i
f i
d
A
A A
E (%) (3.1)


The XRD scans on the demineralised samples were conducted on a PANalytical XRD
XPert Pro powder diffractometer using Co K radiation (courtesy of Dr. Sabine
Verryn). The analysis parameters and settings on the XRD system for carbon fraction
analysis are presented in Table 3.7.

Table 3.7: Analysis parameters and settings on the XRD system for carbon crystallite
analysis.
Anode Material Cobalt, Co target
Generator Settings 45 mA, 35 kV
Angle of scan (2) 4.0 2 120
K
1
() 1.78901
K
2
() 1.7929
K

() 1.62083
K
2
- K
1
ratio 0.5
Step Size (2) 0.017
Scan Step Time (second) 13.335
Scan Type Continuous
PSD Mode Scanning
PSD Length (2) 2.12
Divergence Slit Type Programmable
Divergence Slit Size (mm) 15
Specimen Length (mm) 10
Measurement Temperature (C) 25
Spinning Yes


X-ray intensities were measured and recorded automatically by the XPert Highscore
plus software. The observed intensity in arbitrary units was processed to get the
reduced intensity in atomic unit (a.u) through the following steps:
Chapter 3: Coal and char Characterisation

71

Obtaining the raw diffractogram of the sample.
Obtaining the diffractogram of spiked sample with a reference material of
known peak (pure silicon was used in this study).
Combining the two resulting diffractograms (spiked and unspiked) for any
shift in peaks and corrections if any.
Doing K
2
stripping was done on the resulting diffractogram, to obtain a final
diffractogram with X-ray intensity due to the K
1
radiation only.
The resulting final diffractogram was corrected for polarisation and
geometrical factors using the XPert Highscore plus application. The
absorption factor, A() was also corrected for using the Milberg equation,
shown below (Shiraishi and Kobayashi, 1973; Hubbell and Seltzer, 1996).


) 1 (
1
1 ) (


= e A (3.2)
where is evaluated from:

2 cos . . 2
'
ec A
|
|

\
|
= (3.3)
and
|
|

\
|

is the mass absorption coefficient for Co K radiation; and is the


bulk density of the sample.
The average carbon crystallite parameters for both the coal and char samples
such as: interlayer spacing, d
002
, were determined using Braggs Law (Equation
2.3, Section 2.5.3), while the crystallite height, L
c
, and average crystallite
diameter, L
a
, were evaluated using the Scherrer equation (Equation 2.1 and
2.2). The average number of aromatic layers per carbon crystallite, N
ave,
was
calculated from d
002
and L
c
, using Equation 2.3 (Section 2.4.3).
Chapter 3: Coal and char Characterisation

72

The aromaticity, f
a
, of the samples was calculated from the area under the d
002

peak and
002
side band using a combination of the Origin 6.1 software and the
HighScore Plus curve fitting facility.
By considering that the scattered X-ray intensity is a combination of intensities
contributed by the crystalline and amorphous carbon structure (Lu et al.,
2001), the reduced intensity of the coal and char samples, I, in atomic units
can be related to the two separate contributions. Thus:

am cr
I I I + = (3.4)

Since the intensity contributed by the fraction of amorphous carbon is constant
over the whole scattering region and does not contribute to the peak intensity
but reflected only in the background (Franklin, 1950; Ergun and Tiensuu,
1959; Lu et al., 2001), it is related to and equal to the fraction of amorphous
carbon, X
A
, contained in the samples. Equation 3.5 can thus be rewritten as:

A cr
X I I + = (3.5)

If only the reflections within the (002) peak is considered, as suggested by
Warren (1941) and referenced by Franklin (1950); Ergun and Tiensuu (1959)
and Lu et al. (2001), then:

A
X I I + =
002
(3.6)

By plotting the reduced X-ray intensity against s=2Sin/, the fraction of
amorphous carbon, X
A
, was evaluated from the symmetrical curve of the d
002

peak using the relation derived by Franklin (1950):
( )
( )

|
|

\
|
|

\
|
=
n
ave
ave
n
s d N
s d N
p
s
I
002
2
002
2
2
002
sin
sin 0606 . 0

(3.7)
Chapter 3: Coal and char Characterisation

73

Combining Equation 3.6 and 3.7 yields:

( )
( )

|
|

\
|

n
ave
ave
n
A
A
s d N
s d N
p
s
X
X I

002
2
002
2 2
002
sin
sin
0606 . 0 1
(3.8)

where s=2Sin/; P
n
, is the fraction of aromatic carbon contained within the
d
002
peak.
Since all the carbon within the (002) peak are completely crystalline and
aromatic, Pn=1, and introducing, I
max
, as the maximum reduced intensity
within the (002) peak, the solution for the fraction of amorphous carbon now
becomes:

( )
( ) s d N
s d N
s X
X I
ave
ave
A
A

002
2
002
2
2
max
sin
sin 0606 . 0
1
(3.9)

The degree of disorder index, DOI, of the samples was calculated from the
aromaticity, f
a
, and the fraction of amorphous carbon, X
A
, using the relation
(Lu et al., 2002a):

) 1 )( 1 (
a A A
f X X DOI + = (3.10)


3.6.5 Petrographic Analysis
The petrographic analyses of all the coal and char samples were done at Petrographics
SA, Pretoria (du Cann, 2007 and 2008). These analyses involve the microscopic
examination of the samples, which in conjunction with other techniques gives the
necessary details as it pertains to the age or maturity of the coal samples, organic
Chapter 3: Coal and char Characterisation

74

composition, minerals and organic matter associations, as well as the physical
properties and conditions analysis of the samples.

The following analytical methods were used for the analyses. Petrographic block
preparation of the samples (parent coals and chars) using epoxy resin was done
according to ISO 7404 - 2 (1985) standard. The resulting samples were grinded and
polished. The samples were then examined under the microscope (Leica DM4500P)
under a reflected light oil immersion lens. Vitrinite random reflectance measurements
to establish rank or degree of maturation of the samples were conducted in accordance
with the ISO Standard 7404 5 method (1994).

The macerals groups (mono-, bi-, and tri-macerals) were quantified by a 500 point-
count technique in accordance with the ISO Standard 7404 - 3 method (1994). The
reactive inertinite macerals were identified according to Smith, Roux and Steyn for
South African coals (Smith et al., 1983). A total maceral reflectance scan was also
undertaken on each of the sample (coals and chars). 250 random reflectance readings
were taken on all macerals over the polished surface of each petrographic block.

Coal microlithotype, carbominerite and minerite analyses to determine the organic /
inorganic associations were conducted in accordance with the ISO Standard 7404 - 4
method (1988). A condition analysis was also performed on all the coal and char
samples using the same 500 point-count technique as in the case of group macerals to
analyse the extent of visible changes in the organic constituents due to heat treatment
by the char production process.


3.6.6 Structural Analysis
Various analyses were conducted on the coal and char samples to ascertain their
physical and structural properties. Some of the properties identified for analysis in this
study are: the micropore surface area, average micropore diameter, skeletal density,
microporosity and carbon crystallite microstructural parameters.
Chapter 3: Coal and char Characterisation

75

3.6.6.1 CO
2
Adsorption Analysis
Micropore surface area of the parent coal and chars was measured by CO
2
adsorption
using the Dubinin- Radushkevich method (Anderson et al., 1965; Walker and Kini,
1965; Walker et al., 1988) on a Micromeritics ASAP 2020 Accelerated Surface Area
and Porosimetry System (Micromeritics, 2006). Carbon dioxide has been used most
often to measure micropore surface areas of coals/chars (Anderson et al., 1965;
Walker and Kini, 1965; Walker et al., 1968). Prior to CO
2
adsorption, the samples
(about 0.20 grams each) were degassed under vacuum (10.0 mHg) in order to
eliminate moisture or condensed volatiles, which could prevent the adsorbate
accessibility. Degassing of the coal samples were done at ambient temperature (25 C)
for 48 hours to avoid the release of volatiles and the formation of low temperature
chars; while that of the chars were conducted at 380 C. A ramp of 2 C min
-1
with
holding temperatures of 90 C (for 1 hour ) and 380 C (48 hour) was used for the
chars. The evacuated samples were analysed at 0 C in an ice bath. The results were
processed using the ASAP 2020 V3.01 software linked to the facility. The ASAP
2020 uses two independent vacuum systems, one for sample analysis and one for
sample preparation. This allows preparation and analysis to proceed simultaneously
without the inherent delay found in single vacuum system analyzers that must share a
vacuum pump.

The average micropore diameter and micropore volume was determined using the
Horvart-Kawazoe (H-K) method (Hovarth and Kawazoe, 1983; Jaroniec et al., 1996;
Kowalczyk et al., 2002; Micromeritics, 2006). Porosity of both coals and subsequent
chars (D
p
5) was determined using the CO
2
adsorption data. Porosity of a material,
usually expressed as a percentage, is the ratio of the total volume of voids available
for fluid transmission to the total volume of the porous medium (Webb, 2001;
Micromeritics, 2006). To determine the volume of open pores, the cummulative
volume of CO
2
adsorbed has to be calculated. This was achieved by integrating the
cumulative pore volume over the entire measured pore diameter range. Hence
porosity was evaluated using the relation (Webb, 2001; Micromeritics, 2006):


=
pore
pore
c o
dD
dD
dV
(3.11)
Chapter 3: Coal and char Characterisation

76

3.6.6.2 Helium Pycnometry
The skeletal or apparent density of a solid particle is the ratio of the mass of discrete
particles of the solid material to the sum of the volumes of the solid materials in the
particles and closed, blind or inaccessible pores within the particles (Gan et al., 1972;
Strugala, 1994; Webb, 2001). The determination of the skeletal density of the coal and
char samples was conducted, using the Quantachrome Stereopycnometer (Model:
SPY-4). The pycnometer is manually operated and is specifically designed to measure
the true volume of solid materials by using the Archimedes principle of fluid (gas)
displacement (Helium was used in this study) and the technique of gas expansion
(Webb, 2001; Quantachrome Instruments, 2009). From this volume, density is easily
determined using the sample mass. It has a digital pressure display resolution of 0.001
psi and an error of < 0.2% when properly prepared (both sample and equipment),
thermally equilibrated, and sample occupies greater than 75% of nominal sample cell
volume (Tamari and Aguilar-Chvez, 2004; Quantachrome Instruments, 2009).
Ideally, a gas is used as the displacing fluid since it needs to penetrate the finest pores
to ensure maximum accuracy. For this reason helium is recommended and was used
in this study, since its small atomic dimension enables entry into crevices and pores
approaching one Angstrom (10
-10
m). Its behaviour as an ideal gas is also desirable.
Other gases such as nitrogen can be used, often with no measurable difference. The
pycnometer also offers a choice of two interchangeable sample cells (20 and 135 cm
3
)
used in conjunction with a single reference volume.
About 10 grams of sample is placed in the 20 cm
3
sample cell and degassed by
purging with a flow of helium, by vacuum, or by a series of pressurization cycles. The
standard analysis is performed by pressurizing the sample cell to approximately 17 psi
and recording the value. The selector valve is rotated so the gas expands into the
reference or added volume and that lower pressure is recorded. From these two
readings, the sample volume can be quickly and accurately calculated (Tamari and
Aguilar-Chvez, 2004; Quantachrome Instruments, 2009).


Chapter 3: Coal and char Characterisation

77

3.7 Characterisation Results and Discussion

3.7.1 Chemical Analyses
Results of proximate, ultimate, and total sulphur analyses, as well as the hydrogen-
carbon (H/C) and the oxygen-carbon (O/C) atomic ratios of all the coal and char
samples are given in Table 3.8.
From the summary, the parent coals inherent moisture content ranges from 3.40 wt.
%, adb for coal C to 6.50 wt. %, adb for coal D2. These coals are low in volatile
matter (VM) content with coal B, C, C2 and D2 having VM values of 23.3, 21.4, 18.2
and 21.6 wt. %, adb respectively. Since the charring process drives away moisture and
volatiles in the parent coal resulting in a carbon enriched char, the inherent moisture
and volatile matter content of the chars reduced drastically. Moisture content
decreased from 4.00 wt. %, adb in coal B to 0.90 wt. %, adb in char B, while it went
down from 6.50 wt. %, adb in coal D2 to 0.37 wt. %, adb in char D2. Similar
downward trends were also observed in chars C and C2.
Volatile matter content also saw the same downward movement with a range of 18.2
to 23.3 wt. %, adb in the coals reduced significantly to a very low value range of 0.80
to 2.70 wt. %, adb in the chars. Chars have been known to contain traces of moisture
and VM matter especially when prepared at temperatures below 1200 C. Several
investigators have reported traces of moisture and VM in their chars prepared at or
below 1000 C (Dutta et al., 1977; Matsui et al., 1987; Ochoa et al., 2001; Sina et
al., 2003; Zhang et al., 2006; Murillo et al., 2006; Zhang et al., 2007; Everson et al.,
2008b)
The coals were also characterised as high ash coals with their respective ash content
approximately equal having values of: 25.6, 29.7, 28.6 and 29.0 wt. %, adb for coal B,
C, C2 and D2 respectively. The chars experienced a corresponding increase in ash
content with respective values of 28.5, 33.2, 34.0 and 36.7 wt. %, adb for char B, C,
C2 and D2. It is worthy to note that the approximate ash content of the parent coal
was one of the motivations for the selection of these four coal samples for this study.
Chapter 3: Coal and char Characterisation

78

Fixed carbon content of the coals were: 47.1 wt. %, adb for coal B, 45.5 wt. %, adb
for coal C, 49.4 wt. %, adb for coal C2 and 42.9 wt. %, adb for coal D2. Fixed carbon
content of the chars increased as should be expected after transiting from coal to
values of 69.4, 64.8, 61.2 and 60.2 wt. %, adb for char B, C, C2 and D2 respectively.
The upper heating value of these coals under laboratory conditions on air dry basis
usually referred to as the gross calorific value are low with values ranging from 18.1
MJ/kg for coal D2 to the highest value of 21.4 MJ/kg for coal B. On a calorific value
basis, the four coal samples are graded as Grade D-III coals with all CVs less than
21.5 MJ/kg (CKS 561-1982). These results are similar to the findings of Kaitano
(2007) and Hattingh (2009) on their investigations on some inertinite-rich Highveld
coals and discards. Gross calorific value of the four chars was not determined.
From the ultimate analyses, all four coal samples are rich in elemental carbon with
coals B, C and C2 having 73.1 wt. %, daf and coal D2 having the lowest elemental
carbon content of 68.1 wt. %, daf. This property makes them ideal as feedstock for
coal conversion processes as the elemental carbon can complement the heat required
for some of the endothermic processes in the gasifier. The slight rank difference of
coal D2 (Bituminous medium rank D) was manifested in the elemental carbon
content; as elemental carbon content increases with rank (van Krevelen, 1981; Falcon
and Snyman, 1986; Kabe et al., 2004). Thus, coal D2 that is of a slightly lower rank
than the other three coals, had the lowest elemental carbon content.
Elemental oxygen content of the coals ranges from 19.0 wt. %, daf for coal B to 24.9
wt. %, daf for coal D2. Total sulphur content ranges from 0.64 wt. %, daf for coal D2
to 1.49 wt. %, daf for coal B, while nitrogen content ranges from 1.44 wt. %, daf for
coal D2 to 1.77 wt. %, daf for coal B. On this basis coal D2 should be most desirable
environmentally, since it has minimum content of both total sulphur and nitrogen.
The elemental carbon content of the chars increased accordingly to a range of 89.7 wt.
%, daf for char C2 to 93.1 wt. %, daf for char D2; while total sulphur content
increased to values of 0.42 to 1.49 wt. %, daf. Nitrogen content decreased from parent
coals to chars with reported values ranging from 0.07 wt. %, daf for char D2 to the
highest value of 0.62 wt. %, daf for char B.
Chapter 3: Coal and char Characterisation

79


Table 3.8: Result of proximate and chemical analyses of coal and char samples.
Proximate Analysis Coal B Char B Coal C Char C Coal C2 Char C2 Coal D2 Char D2
(wt. %) adb db adb db adb db adb db adb db adb db adb db adb db
Inherent Moisture 4.00 - 0.90 - 3.40 - 1.20 - 3.80 - 1.58 - 6.50 - 0.37 -
Volatile Matter, VM 23.3 24.3 1.20 1.21 21.4 22.2 0.80 0.81 18.2 18.9 3.20 3.30 21.6 23.1 2.70 2.70
Ash 25.6 26.7 28.5 28.8 29.7 30.7 33.2 33.6 28.6 29.7 34.0 34.5 29.0 31.1 36.7 36.8
Fixed Carbon, FC 47.1 49.0 69.4 70.0 45.5 47.1 64.8 65.6 49.4 51.3 61.2 62.2 42.9 45.8 60.2 60.5
Total 100 100 100 100.01 100 100 100 100.01 100 99.9 99.98 100 100 100 99.97 100
Gross Calorific
Value (adb) (MJ/kg) 21.4 - 20.0 - 20.2 - 18.1 -
Grade (Based on
GCV, (adb)) Grade D III - Grade D III - Grade D III - Grade D III -
Ultimate Analysis (Ash free basis) (wt. %)
Carbon 70.2 73.1 91.2 92.0 70.6 73.1 90.6 91.7 70.4 73.1 88.3 89.7 63.7 68.1 92.7 93.1
Hydrogen 4.66 4.85 2.02 2.04 4.70 4.83 2.02 2.04 4.30 4.51 2.42 2.45 4.56 4.87 2.22 2.23
Nitrogen 1.71 1.77 0.61 0.62 1.61 1.65 0.29 0.30 1.60 1.67 0.55 0.56 1.34 1.44 0.07 0.07
Oxygen 18.2 19.0 3.80 3.85 18.7 19.4 4.75 4.80 19.5 20.3 6.70 6.80 23.3 24.9 3.54 3.55
Total Sulphur 1.26 1.32 1.47 1.49 0.99 1.02 1.10 1.12 0.40 0.42 0.46 0.49 0.60 0.64 1.05 1.05
Total 100.03 100.04 100 100 100 100 99.96 99.96 100 100 100.01 100 100 99.95 99.95 100
H/C atomic ratio 0.792 0.792 0.264 0.264 0.804 0.792 0.264 0.264 0.732 0.744 0.348 0.324 0.852 0.852 0.288 0.288
O/C atomic ratio 0.195 0.195 0.0315 0.0315 0.195 0.195 0.0413 0.039 0.210 0.210 0.057 0.057 0.278 0.278 0.0285 0.0285
Chapter 3: Coal and char Characterisation

80

The reduction in elemental nitrogen from coal to chars could be attributed to the
partitioning and speciation of the coal nitrogen to NO
X
,

that leaves as flue gas with the
volatiles. Conversely, the increase in elemental sulphur contents is an indication that
sulphur in the studied char samples is primarily bound to the organic matrix and the
minerals and therefore, relatively increased in proportion due to the evacuation of
moisture and volatiles during the devolatilisation reaction.
The hydrogen-carbon (H/C) atomic ratio in the coals ranged from 0.79 for coal B to
0.85 for coal D2 while the oxygen-carbon (O/C) atomic ratio were between 0.20 for
coal B and 0.29 for coal D2. These were within the range of values reported for H/C
and O/C atomic ratio in literature for inertinite-rich coals (de la Rosa et al., 1992;
Hanna et al., 1992; Mastalerz and Marc Bustin, 1993; Maroto-Valer et al., 1994). The
H/C and O/C atomic ratio are major factors that affect the aromaticity vis--vis extent
of structural disorderliness in coal and chars (Lu et al., 2001). Both the H/C and O/C
atomic ratios exhibited significant reductions in the transition from coals to chars.
This can be attributed to the removal of the liptinites, aliphatics and lower molecular
components of the coals during pyrolysis.


3.7.2 XRD Mineral Analyses
Minerals and or inorganic species present in both the coal and char samples was
studied using powder XRD technique and Rietveld Autoquan quantification program.
A summary of graphite (carbon) and total mineral content of the coal and char
samples are presented in Table 3.9, while a detailed result of the mineral phase
proportion expressed as a percentage of the total mineral content is given in Table
3.10. The idealised formulas of the mineral species (van Alphen, 2009; Web Mineral,
2009) are also included.
From Table 3.9, it can be observed that the total mineral phases reported for the coal
samples from XRD analysis are higher than their respective ash content from
proximate analysis. This is attributed to the ashing process of the proximate analysis
at high temperature which may result to losses of some volatile mineral phases,
despite the fact that the XRD technique measures the percentage of the crystalline
Chapter 3: Coal and char Characterisation

81

mineral matter only. A similar finding was made by Matjie, (2008). Total mineral
phases of the chars are lower than their ash counterpart from proximate analysis. This
is because the XRD technique does not account for the amorphous phases of the
minerals and decompositions (van Alphen, 2009) may have occurred during the
charring process.


Table 3.9: Percentage of graphite and total crystalline mineral phases of the coal and
char samples from XRD results.
Sample ID Graphite (carbon/ char) Total mineral phases
Coal B 66.31 33.69
Char B 83.07 16.93
Coal C 66.53 33.47
Char C 83.72 16.28
Coal C2 63.96 36.04
Char C2 89.23 10.77
Coal D2 59.50 40.50
Char D2 87.56 12.44

Graphite (carbon) makes up the greater proportion of the coal and char samples from
XRD mineral analysis. This further validates the results from both the proximate and
ultimate analysis. The graphite content in the coal samples ranges between 59.5 wt. %
for coal D2 and 66.5 wt. % for coal B.
Results from Table 3.10 show that all four coal samples are rich in major coal
minerals: kaolinite and quartz. Kaolinite content values (graphite free basis, (gfb))
ranges from 52.8 wt. %, gfb for coal D2 to 79.9 wt. %, gfb for coal C2; while quartz
content value are between 6.19 wt. %, gfb for coal D2 to 17.6 wt. %, gfb for coal B.
Other major mineral species dominant in the coal samples are calcite, dolomite
gypsum and pyrite. Minor minerals observed in the coal samples include small
amounts of muscovite and trace quantities of rutile and siderite. It can be observed
that coal D2 has less abundance of the clay minerals: kaolinite and quartz and very
high proportion of gypsum (29.6 wt. %, gfb) relative to the other coal samples and
may impart some catalytic effect during its utilisation.
Chapter 3: Coal and char Characterisation

82

Table 3.10: Mineral abundance of coals and chars (graphite free bases, (wt. %, gfb)).
Mineral Idealised Chemical
Mineral Proportion (wt. %, gfb)
Specie Formula
1

Coal B Char B Coal C Char C Coal C2 Char C2 Coal D2 Char D2
Calcite CaCO
3
4.03 - 4.31 0.61 2.52 4.00 2.52 1.93
Dolomite CaMg(CO
3
)
2
4.72 - 6.76 - 1.72 - 2.79 -
Gypsum CaSO
4
2(H2O) 4.86 - 7.66 - 1.08 - 29.6 -
Kaolinite Al
2
[Si
2
O
5
](OH)
4
55.5 - 55.6 - 79.9 - 52.8 -
Muscovite KAl
2
(AlSi
3
O
10
(OH)
2
) 11.0 - 10.6 - 6.63 - 3.72 -
Pyrite FeS
2
1.51 - 0.84 - 0.47 - 1.04 -
Quartz SiO
2
17.6 32.2 13.1 37.4 6.38 19.9 6.19 33.2
Rutile TiO
2
0.68 - 0.57 - 0.92 - 0.59 -
Siderite FeCO
3
0.12 - 0.54 - 0.36 - 0.79 -
Cristobalite SiO
2
- 20.1 - 24.0 - 38.9 - 25.1
Illite K
1-1.5
Al
4
[Si
6.5-7
Al
1-1.5
O
20
](OH
4
) - 14.2 - 7.13 - 8.91 - 2.33
Iron alpha -Fe - 2.36 - 1.54 - 1.02 - 4.18
Microcline KAlSi
3
O
8
- 20.1 - 20.8 - 22.6 - 18.2
Oldhamite Ca
0.9
Mg
0.05
Fe
0.05
S - 8.62 - 5.96 - 3.06 - 10.3
Sodalite Na
8
Al
6
Si
6
O
24
Cl
2
- 1.24 - 1.47 - 1.30 - 2.65
Troilite FeS - 1.18 - 1.11 - 0.28 - 2.10
Total 100.02 100.00 99.98 100.02 99.98 99.97 100.04 99.99
1
- van Alphen, (2009); Web Mineral Data, (2009)
Chapter 3: Coal and char Characterisation

83

The transition from parent coal to char via the charring process caused some loss
(water of crystallisation), decomposition and or transformation of the mineral species
in the parent coal as it changes to char at 900 C as seen in Tables 3.9 and 3.10.
Graphite percentage contents increased from parent coal to char in all four samples
ranging in value from 83.07 wt. % for char B to 89.23 wt. % for char C2 up from 59.5
and 66.53 wt. % for the parent coals. This is as expected since the devolatilisation of
the coal results in the enrichment of the resulting char in carbon and ash as seen in the
proximate and ultimate analyses results of the samples (Table 3.8).
On graphite free basis, quartz mineral (SiO
2
) increased by a factor of > 2 in the
transition which was more significant in chars D2 and C2. Quartz in Char D2
increased significantly from 6.19 wt. %, gfb in the parent coal to 33.2 wt. %, gfb. The
same trend was observed for char C2 with a quartz content of 19.9 wt. %, gfb up from
6.38 wt. %, gfb value of the precursor coal. This may be attributed to the
transformation of kaolinite and muscovite in the parent coal.
Calcite was not observed in char B and only very little abundance (0.61 - 4.0 wt. %,
gfb) was observed in the other chars. Calcite and quartz were the only mineral species
that were present in both the parent coals and the chars (except char B). Clay
minerals: illite and microline; and silicates: quartz and cristobalite were the most
abundant minerals in the chars. Pyrite were reduced to alpha iron and some
interactions with siderite in the parent coal yielded troilite in the chars. Other mineral
species in the chars are oldhamite and traces of sodalite.


3.7.3 Ash Analysis (XRF)
The normalised proportions (Loss on Ignition (LOI) and sulphur free basis) of major
inorganic composition of the char ashes from the XRF spectroscopy analysis are
presented in Table 3.11.
Considering the LOI and the total ash content, it can be seen that the ash content from
proximate analysis corresponds well with the total ash content from XRF result. This
is due to the similarity of the ashing process of the two techniques at high temperature
of 815 C. The XRF technique used is regarded as only semi-quantitative for sulphur
Chapter 3: Coal and char Characterisation

84

(Matjie, 2008; Loubser and Verryn, 2008), hence the sulphur contents were not
reported. From the result, SiO
2
and Al
2
O
3
are the most abundant chemical
components in the ashes of the chars which compare well with the XRD mineral
analysis results. These are derived from the clay minerals: quartz and kaolinite which
form the bulk of the mineral in coal and subsequent chars (Spears, 2000).
The presence of Fe
2
O
3
in al the char samples can be attributed to the pyrite and
siderite in the original coal as well as troilite and iron alpha detected in the chars. CaO
and MgO were also detected in all of the four char samples which is a reflection of the
dolomite and calcite phases in the parent coal samples and oldhamite in some
appreciable quantise in the chars, with a larger contribution to CaO content of char B
suspected to be due to the high content of gypsum in the parent coal B. The lowest
value of CaO content was observed in char C2 at 3.02 wt. %, lfb.

Table 3.11: Char sample ash chemistry on LOI and sulphur free basis (wt. %, lfb and
sfb ).
Sample ID
Mineral Species (wt. %, lfb and sfb)
Char B Char C Char C2 Char D2
SiO
2
50.9 49.9 50.8 46.9
Al
2
O
3
29.4 31.7 39.7 34.5
Fe
2
O
3
5.03 3.59 2.50 4.75
CaO 8.78 8.16 3.02 8.18
MgO 1.90 2.16 0.68 2.30
MnO 0.06 0.05 0.03 0.08
TiO
2
1.68 1.91 2.38 2.29
Na
2
O 0.23 0.23 0.26 0.14
K
2
O 0.74 0.98 0.37 0.48
P
2
O
5
1.18 1.31 0.23 0.33
Cr
2
O
3
0.09 0.04 0.05 0.04
Total (wt. %, lfb) 99.99 100.03 100.01 99.99
LOI (wt. %, adb) 71.3 67.5 68.9 67.3
Ash content of chars (wt. %, adb) 28.5 33.2 34.0 36.8
Alkali Index (-) 5.92 6.15 2.55 7.17
Chapter 3: Coal and char Characterisation

85

It is worthy to note that this analysis was confined to the chars only as the parent coals
were not used for the reactivity experiments; while the motivation for the analysis is
to study the inherent catalytic influence, if any, of the high ash components of the
chars on their CO
2
reactivity.
The presence of K
2
O in all the chars with the highest value in char C is related to the
muscovite in the parent coals and illite and microcline in the chars, while Na
2
O and
TiO
2
are derived from the sodalite in the chars and rutile in the parent coals
respectively.
A significant contribution of this analysis is the study of the catalytic effects due to
the presence of these elemental components by determining the respective lumped
parameter, alkali index AI, of the chars which are also presented in Table 3.11. It had
been established by various investigators: Walker and Hippo (1975), Miura et al.
(1989), Httinger and Natterman (1994), Tomita (2001), Zhang et al. (2006), and Lee
(2007) reported that, elemental components such as: Fe
2
O
3
, CaO, MgO, Na
2
O, K
2
O
impact some catalytic effect on the gasification reaction of coals and chars. The AI,
refers to the ratio of the total weight fraction of the basic species in the ash (CaO,
MgO, K
2
O, Na
2
O and Fe
2
O
3
) to the total weight fraction of the acidic compounds
(SiO
2
and Al
2
O
3
) in the ash, multiplied by the ash contents in weight percent of the
chars. This formulation is shown as Equation 3.12; while the AI results of the chars
are presented in Table 3.11.


%
3 2 2
3 2 2 2
ash
O Al SiO
O Fe O Na O K MgO CaO
AI
|
|

\
|
+
+ + + +
= (3.12)


3.7.4 X-ray Diffraction (XRD) Carbon Crystallite Analysis
The second aspect of analysis involving the use of the X-ray diffraction technique was
the study of the carbon crystallite properties of the coal and char samples. To reduce
noise and the effects of mineral matter on the XRD diffractogram and for a simplified
study of the carbon fraction, a three stage HCl-HF-HCl demineralisation was
conducted on the samples. The outcome of this, presented as proximate analysis of the
samples including the effectiveness of the process, is summarised in Table 3.12.
Chapter 3: Coal and char Characterisation

86

Table 3.12: Proximate analysis of raw and demineralised coal and char samples (wt.
%, db).
Sample
ID
VM
(wt. %, db)
Ash
(wt. %, db)
Fixed Carbon
(wt. %, db)
E
d
1
(%)
Coal B
Raw 24.3 26.6 49.0
96.7
Demin
2
27.0 0.88 72.2
Char B
Raw 1.21 28.8 70.0
89.1
Demin 5.79 3.13 91.1
Coal C
Raw 22.1 30.7 47.1
97.9
Demin 25.3 0.66 74.0
Char C
Raw 0.81 33.6 65.6
93.9
Demin 5.59 2.06 92.4
Coal C2
Raw 18.9 29.7 51.3
98.3
Demin 20.3 0.5.0 79.2
Char C2
Raw 3.25 34.5 62.2
89.8
Demin 5.49 3.52 91.0
Coal D2
Raw 22.4 28.8 48.7
96.7
Demin 27.2 0.94 71.8
Char D2
Raw 2.71 36.8 60.5
90.5
Demin 5.55 3.48 91.0
1
- E
d
Effectiveness of demineralisation
2
- Demineralised sample

It is obvious from the result that, higher effectiveness of demineralisation was
achieved on the parent coals (96.7- 98.3%) than on the chars (89.1- 93.9%) with the
corresponding ash contents reduced to 0.5 - 0.94 wt. %, db and 2.06 - 3.52 wt. %, db
respectively. This is to be expected since the pyrolysis reaction leading to the chars at
900 C, apart from driving off the volatiles, hardens the chars which may prevent the
demineralisation agent from reaching the mineral inclusions easily. This will also
culminate to an increase in both the skeletal and bulk density of the resulting chars.
An increase in both the fixed carbon and volatile matter content was also observed for
the demineralised coal and char samples. Similar results were reported by Lu et al.
(2001), Maity and Mukherjee (2006), Kawakami et al. (2006), and Van Niekerk
(2008).
Chapter 3: Coal and char Characterisation

87

The raw diffractograms, corrected for polarisation and geometrical factors from the
XRD analysis for the demineralised coal and char samples, are shown in Figure 3.2.
The background due to the amorphous carbon fraction was further removed and the
diffractograms smoothened with the HighScore Plus peak analysis tool and are
presented in Figure 3.3. The diffractograms of all the demineralised coal and char
samples investigated in this study possess the same graphitic features as those
reported in literature (Franklin, 1950 and 1951; Hirsch, 1954; Alexander and Sommer,
1959; Shiraishi and Kobayashi, 1973; Kumar and Gupta, 1995; Lu et al., 2001, 2002a
and 2002b; Aso et al., 2004; Kawakami et al., 2006; Maity and Mukherjee, 2006; Wu
et al., 2008).


Figure 3.2: Raw diffractograms of coal and char samples.
0
1000
2000
3000
4000
5000
6000
0 20 40 60 80 100 120
I
n
t
e
n
s
i
t
y
,

I
(
c
o
u
n
t
s
)
2 () CoK
Coal B
Coal C
Coal C2
Coal D2
0
1000
2000
3000
4000
5000
0 20 40 60 80 100 120
I
n
t
e
n
s
i
t
y
,

I
(
c
o
u
n
t
s
)
2 () CoK
Char B
Char C
Char C2
Char D2
Chapter 3: Coal and char Characterisation

88

The graphitic features were established by the presence of the (002) band
corresponding to the (00l) position of graphite. The (00l) position is related to the
inter-layer spacing of graphite and its resemblances in the diffractograms ((002) band)
occur at 29.56 2 30.05 for the coals; and 28.66 2 29.71 for the chars.
The (10) and (11) bands were also observed, which correspond to the (hk0) lines of
graphite, related to the hexagonal ring structure. The (10) bands occur at position
50.90 2 51.60 and 51.90 2 52.90, while the (11) band was observed at
98.21 2 98.59 and 96.70 2 97.45 for the coals and chars respectively.
These peak positions are significant in the calculation of the crystallite lattice
parameters.


Figure 3.3: Corrected and smoothened diffractograms of coal and char samples.
0
500
1000
1500
2000
2500
0 0.2 0.4 0.6 0.8 1
I
n
t
e
n
s
i
t
y
,

I
(
c
o
u
n
t
s
)
s=2Sin/ (
-1
)
Coal B
Coal C
Coal C2
Coal D2
0
300
600
900
1200
1500
1800
0 0.2 0.4 0.6 0.8 1
I
n
t
e
n
s
i
t
y
,

I
(
c
o
u
n
t
s
)
s=2sin/ (
-1
)
Char B
Char C
Char C2
Char D2
(002)
(10)
(11)
(002)
(10)
(11)
-band
-band
Chapter 3: Coal and char Characterisation

89


The assymetry exhibited by the (002) band had necessitated the dilineation of the peak
into the d
002
peak and the -side band which appears as a shoulder bands in the
difractograms (Franklin, 1950 and 1951; Hirsch, 1954; Ergun and Tiensuu; 1959;
Schoening, 1983; Lu et al., 2001; Lu et al., 2002a). The former is associated with the
aromatic ring stacking while the later is due to the aliphatic side chains. The small
spikes on the diffractogram are peaks of traces of minerals still remaining in the
demineralised samples as indicated by the effectiveness of demineralisation, E
d
,
(Table 3.12).
The inter-layer spacing was calculated from the (002) peak position using the Braggs
equation; while the crystallite height, L
c
, and diameter, L
a
, were determined from the
peak positions, and the full width at half maximum (FWHM) of the (002) and (10)
peaks respectively. The (11) band was not used due to its diffuse and obscure nature
which make quantitative analysis difficult.
The characteristics of the diffractograms and the annealing effects of the transition
from parent coal to char at 900 C (samples B and C) are expounded in Figure 3.4.
The (002), (10) and (11) peaks of the chars are broader than that of the precursor
coals, while the -shoulder-band is more prominent in the coals than in the chars. The
broadening of the (002) peak in the diffractograms obtained for the chars is due to the
closer packing and structural re-ordering, re-orientation and better alignment of the
aromatic carbon layers, which results in an increase in the inter-layer spacing and a
decrease in both the crystallite height and diameter of the chars, compared to that of
the original coals. Broadening of peaks is evidence of a decrease in carbon crystallite
size, while sharper peaks are signs of growth of the carbon crystallite (Kuroda and
Akamatu, 1959; Short and Walker, 1963; Kumar and Gupta, 1995; Lu et al., 2002a
and 2002b; Wu et al., 2008).
Crystallite condensation, as observed in this study, was also observed by Takagi et al.
(2004) for their chars produced at 760 C and 900 C. A careful analysis of the scatter
of the results reported by Kawakami et al. (2006), also show an increase in inter-layer
spacing between chars prepared at 900 C and 1200 C. It should be noted that
significant crystallite growth usually starts from 1600 C (Kuroda and Akamatu,
1959; Kumar and Gupta, 1995; Lu et al. 2002a and 2002b; Wu et al. 2008), with the
Chapter 3: Coal and char Characterisation

90

experiments conducted by these investigators, confined to temperatures above 1200
C. Takagi et al. (2004) however noted that broadening of peaks at lower temperatures
of heat treatment (350 - 920 C) is more significant for lower rank coals, while for
higher rank coals, sharper peaks may be observed.


Figure 3.4: Comparison of coal and char diffractograms for samples B and C.


The diffuse and in some cases almost absent -sideband in the diffractograms of the
chars is due to the loss of the aliphatic side chain during the char production process.
This culminates in a more ordered structure (Russell et al. 1999; Davis et al. 1995; Lu
et al. 2001; Van Niekerk, 2008), which impacts on the chemical properties of the
0
300
600
900
1200
1500
1800
0 0.2 0.4 0.6 0.8 1
I
n
t
e
n
s
i
t
y
,

I
(
c
o
u
n
t
s
)
s=2sin/ (
-1
)
Coal B
Char B
0
300
600
900
1200
1500
1800
0 0.2 0.4 0.6 0.8 1
I
n
t
e
n
s
i
t
y
,

I
(
c
o
u
n
t
s
)
s=2Sin/ (
-1
)
Coal C
Char C
-band
(10)
(11)
-band
(10)
(11)
(002)
(002)
Chapter 3: Coal and char Characterisation

91

char, as aromaticity and the fraction of amorphous carbon and other crystallite
parameters.
Van Niekerk (2008), however reported a contrasting result for the investigated
inertinite-rich Highveld coal sample, as a prominent -side band was observed, which
was entirely missing in the diffractogram of the vitrinite-rich Waterberg coal. This
means that the vitrinite-rich coal is structurally more ordered than the inertinite-rich
sample. The result contradicts the high aliphatic content of vitrinite-rich coals (Choi et
al., 1989; dela Rosa et al., 1992) and may have led to his inability to determine the
aromaticity of the studied vitrinite-rich coal sample using XRD technique.


3.7.4.1 Determination of Aromaticity of Coal and Char Samples
The aromaticity of both coal and char samples were determined from the areas under
the (002) peak and the -side band (Lu et al., 2001, 2002a and 2002b; Maity and
Mukherjee, 2006). This was done using a curve fitting analysis tool on the HighScore
Plus application. A confirmation was conducted on the result by using the peak
analysis, curve fitting and the data analysis tool of Origin 6.1 to determine the areas
under the (002) peak and the -side band. The curve fitting and analysis of the two
peaks to get the respective peak areas for coal B and Char C2, using HighScore Plus,
is shown in Figure 3.5.
A comparison of the results from the two methods described above is summarised in
Table 3.13. It can be seen from the table that the results obtained from the two
different data analysis applications are very similar to each other, thus validating the
results.

Table 3.13: Comparison of aromaticity results from HighScore Plus and Origin 6.1.
Determination of Aromaticity (fractional values (-))
Coal
B
Char
B
Coal
C
Char
C
Coal
C2
Char
C2
Coal
D2
Char
D2
By HighScore Plus 0.7993 0.8822 0.8106 0.9091 0.8527 0.9513 0.7539 0.8489
By Origin 6.1 0.8011 0.8839 0.8113 0.9192 0.8517 0.9483 0.7513 0.8470
Chapter 3: Coal and char Characterisation

92


Figure 3.5: Determination of area under d
002
and - band using HighScore Plus for coal B
and char C2.


3.7.4.2 Determination of Fraction of Amorphous Carbon of the
Coal and Char Samples
The determination of the fraction of amorphous carbon contained in both the coal and
char samples were carried out by normalising the diffractograms shown in Figure 3.4
to obtain the reduced intensity curve of the (002) peak (Figure 3.6), according to the
method proposed by Franklin (1950) and variously used by Ergun and Tiensuu
(1959); Short and Walker (1963); Lu et al. (2001, 2002a, and 2002b); Kawakami et
al. (2006).
Chapter 3: Coal and char Characterisation

93

The symmetrical section of the (002) band used for this analysis for coal C2 and char
C is shown in Figure 3.6. In Table 3.14, the I
max
, S
max
, and other parameters for the
calculation of the fraction of amorphous carbon in coal C2 and char C are presented.
It should be noted that the results of X
A
, are returned as negative and reported as
absolute values; and the reciprocal of S
max
should be equal to the inter-layer spacing,
d
002
, if the (002) band symmetric profile is correct.


Figure 3.6: Determination of amorphous fraction of carbon, X
A
, from (002) profile of coal
C2 and char C.


0
2
4
6
8
10
12
0.2888 0.2889 0.289 0.2891 0.2892
R
e
d
u
c
e
d

I
n
t
e
n
s
i
t
y
,

I
0
0
2
(
a
.
u
)
s=2Sin/ (
-1
)
(002) Profile- Coal C2
0
2
4
6
8
0.283 0.2831 0.2832 0.2833 0.2834 0.2835
R
e
d
u
c
e
d

I
n
t
e
n
s
i
t
y
,

I
0
0
2
(
a
.
u
)
s=2Sin/ (
-1
)
(002) Profile- Char C
I
max

S
max

I
max

S
max

Chapter 3: Coal and char Characterisation

94

Table 3.14: Determination of amorphous fraction of carbon for coal C2 and char C.
Sample ID Coal C2 Char C
I
max
12.06 8.405
d
002
3.460 3.530
N
ave
5.900 5.410
S
max
0.289 0.283
I-X
A
/1-X
A
7.860 6.960
X
A
0.613 0.242

Results from the carbon fraction analysis, using XRD techniques, are summarised in
Table 3.15.


Table 3.15: Result on carbon crystallite analysis using XRD.
Sample
ID
Coal
B
Char
B
Coal
C
Char
C
Coal
C2
Char
C2
Coal
D2
Char
D2
d
002
() 3.49 3.58 3.48 3.53 3.46 3.49 3.46 3.50
L
c
() 15.5 10.5 15.4 11.4 16.9 11.4 14.8 11.9
N
ave
(-) 5.45 3.92 5.41 4.23 5.90 4.26 5.28 4.40
L
a
() 9.32 7.52 12.2 10.1 12.6 11.2 12.0 10.7
f
a
(-) 0.80 0.88 0.81 0.92 0.85 0.95 0.75 0.85
X
A
(-) 0.49 0.28 0.51 0.24 0.61 0.23 0.66 0.48
DOI
1
(-) 0.59 0.37 0.60 0.30 0.67 0.27 0.75 0.56
1
- Degree of disorder index


It can be observed from the results, that apart from the inter-layer spacing, d
002
, and
aromaticity, f
a
, which generally increased from coal to char, the other parameters: the
average crystallite diameter, L
a
, average crystallite height, L
c
, average number of
aromatic aromatic layers per carbon crystallite, N
ave
, fraction of amorphous carbon,
X
A
; and the degree of disorder index, DOI generally decreased from parent coals to
chars.
Chapter 3: Coal and char Characterisation

95

The increase in the inter-layer spacing from parent coal to char is as a result of the
loss of moisture and volatiles and the annealing of the carbon crystallites causing the
disordered carbons of the coal to be more structurally ordered and orientated in the
chars. The decrease in lattice parameters: L
c
, L
a
, and N
ave
, is as a result of a more
compact, ordered and orientated structure impacted on the char by the heat treatment
process (Davis et al., 1995; Russell et al., 1999). This hardens the char and influences
both its skeletal and bulk density. Similar findings were made by Short and Walker
(1963); Takagi et al. (2004); Kawakami et al. (2006); Bouhadda, et al. (2007) and Wu
et al. (2008).
It is generally known that the inter-layer spacing increases with elemental carbon
content vis--vis rank of coal (Lu et al., 2001; Takagi et al., 2004; Maity and
Mukherjee, 2006). This was not observed on the coal samples and thus not reported,
and may be attributed to the fact that all four coal samples are about the same rank.
Coal B, C, and C2 are bituminous medium rank C with the same elemental carbon
content of 73.1 wt. %, daf; while coal D2 is classified as bituminous medium rank D
with elemental carbon content of 68.1 wt. %, daf.
The dependency of aromaticity on the atomic hydrogen-carbon ratio is presented in
Figure 3.7(i) and follows an almost linear trend of decreasing aromaticity with
increasing hydrogen-carbon ratio. This can be explained by the fact that hydrogen
content in coal is associated with the low molecular mass and aliphatic group and
increase in the proportion of this group will result in a higher volatile matter content
and lower aromaticity (Choi et al., 1989). Similar trends were reported by Solum et
al. (1989), Maroto-Valer et al. (1994); and validated by the result of Lu et al. (2001).
The correlation of the fraction of amorphous carbon in the coal samples to the atomic
hydrogen-carbon ratios yielded a wide scatter (not reported) but increased with the
oxygen-carbon ratios (Figure 3.7 (ii)). The plot of aromaticity versus the atomic
oxygen carbon ratio exhibited a wide scatter and was not reported as well.
It may thus be concluded that, while aromaticity of coals depend on the atomic
hydrogen ratio, the fraction of amorphous carbon may be influenced appreciably by
the atomic oxygen-carbon ratio which needs to be investigated further. However, most
investigators have noted that the mechanism of structural ordering is not well
understood (Franklin, 1951; Hirsch, 1954; Kumar and Gupta, 1995; Lu et al., 2001).
Chapter 3: Coal and char Characterisation

96


(i) Relationship between aromaticity and hydrogen (ii) Relationship between fraction of amorphous
-carbon ratio of coals carbon and oxygen-carbon ratio of coals

Figure 3.7: Relationship between aromaticity and fraction of amorphous carbon and the
atomic ratios of hydrogen and oxygen to carbon in coal samples.

The increase in aromaticity from coals to chars is related to the loss of aliphatic side
chains in the parent coal due to the charring process evident from the obscurity and
near-loss of the -sideband as observed in the diffractograms of the chars. Since the
fraction of amorphous carbon refers quantitatively to the fraction of disordered
carbon, it should be expected that its content decreased from the parent coal to the
annealed chars, whose structure is more ordered due to the annealing of the
turbostratic structure of the parent coals and the formation of a more crystalline
structure in the chars. The results of Senneca et al. (1998); Lu et al. (2001, 2002a, and
2002b); Kawakami et al. (2006); Maity and Mukherjee (2006) also correlate well with
the outcome of this study.
The decrease of the degree of disorder index (DOI) from coal to char is an indication
that the chars are more ordered and more crystalline than the parent coal. This was
corroborated by an increase in the aromaticity, and the decrease in the average
crystallite height and diameter, a further indication of the compactness and the degree
of orderliness of the chars (Lu et al., 2002a).
0.74
0.76
0.78
0.8
0.82
0.84
0.86
0.7 0.75 0.8 0.85 0.9
A
r
o
m
a
t
i
c
i
t
y
,

f
a
(
-
)
H/C atomic ratio (db) (-)
Coal B Coal C Coal C2 Coal D2
0.47
0.51
0.55
0.59
0.63
0.67
0.18 0.2 0.22 0.24 0.26 0.28 0.3
F
r
a
c
t
i
o
n

o
f

a
m
o
r
p
h
o
u
s

c
a
r
b
o
n
,

X
A
(
-
)
O/C atomic ratio (db) (-)
Coal B Coal C Coal C2 Coal D2
Chapter 3: Coal and char Characterisation

97

Correlations were made of some of the carbon crystallite parameters of the chars, as
shown in Figure 3.8. Both the fraction of amorphous carbon and the degree of
disorder index was found to decrease with increasing aromaticity (Figure 3.8(i) and
(ii)). Increasing aromaticity is an indication of increasing structural orderliness, hence
a decreasing DOI and fraction of amorphous carbons (Davis et al., 1995; Maity and
Mukherjee, 2006; Lu et al., 2002a; Wu et al., 2008).


(i) Relationship between aromaticity and fraction (ii) Relationship between aromaticity and degree
amorphous carbon in chars of disorder index (DOI) in chars

(iii) Relationship between fraction of amorphous carbon and degree of disorder index of chars

Figure 3.8: Relationship between various crystallite parameters of char samples.
0.2
0.25
0.3
0.35
0.4
0.45
0.5
0.84 0.87 0.9 0.93 0.96
F
r
a
c
t
i
o
n

o
f

a
m
o
r
p
h
o
u
s

c
a
r
b
o
n
,

X
A
(
-
)
Aromaticity, f
a
(-)
Char B Char C Char C2 Char D2
0.22
0.28
0.34
0.40
0.46
0.52
0.58
0.84 0.87 0.9 0.93 0.96
D
e
g
r
e
e

o
f

d
i
s
o
r
d
e
r

i
n
d
e
x
,

D
O
I
(
-
)
Aromaticity, f
a
(-)
Char B Char C Char C2 Char D2
0.20
0.25
0.30
0.35
0.40
0.45
0.50
0.2 0.3 0.4 0.5 0.6
F
r
a
c
t
i
o
n

o
f

a
m
o
.

c
a
r
b
o
n
,

X
A
(
-
)
Degree of Disorder Index, DOI (-)
Char B Char C Char C2 Char D2
Chapter 3: Coal and char Characterisation

98

The fraction of amorphous carbon was found to increase with degree of disorder index
(DOI) (Figure 3.8(iii)). This is not unusual as an increase in both parameters is
indicative of a higher relative abundance of disordered carbon (Lu et al., 2002a)
Although not reported because char D2 was out of trend, the aromaticity (considering
chars B, C and C2) was observed to decrease with an increase in both the inter-layer
spacing and fraction of amorphous carbon. This is to be expected, given the fact that
as d
002
increases further away from the inter-layer spacing of graphite (3.35 - 3.44 )
(Franklin, 1951, Hirsch, 1954, Ergun, 1968, Kumar and Gupta, 1995; Takagi et al.,
2004), the disorder and non-alignment of the carbon structure increases. It is also
expected that the fraction of amorphous carbon will correspondingly increase with
increasing inter-layer spacing; an indication of the level of structural ordering and the
amount of aromatic carbons. However, char D2 was out of trend in this regard and
was not reported.


3.7.5 Petrographic Analyses
Results and discussions relating to the petrographic analysis of the coal and char
samples are presented in this section. Results presented include the reflectance
properties of both vitrinite- and total maceral reflectance scans and organic
component analysis which include maceral, microlithotype, carbominerite and char
form analysis. Results from minerite analysis which describes the inorganic
components of the samples under the petrographic microscope and general condition
analysis are also presented.


3.7.5.1 Reflectance Properties
Reflectance analysis has been distinguished as a vital tool in determining the rank of
coal using the rank parameter based on vitrinite random reflectance (Cameron and
Hunt, 1985; Falcon and Snyman, 1986; Snyman, 1989; Cloke and Lester, 1994). For
this study, the coal vitrinite maceral, collotelinite was used as the reference material as
its reflectance increases uniformly as the coalification process progresses and is
independent of the vitrinite content and coal grade (Falcon and Snyman, 1986; du
Chapter 3: Coal and char Characterisation

99

Cann, 2007). Reflectance measurements were conducted in oil immersion in
accordance with ISO 7404 - 5 (1994). Summary of the major reflectance properties of
the coal and char samples are shown in Table 3.16.
The vitrinite random reflectance results showed that according to ISO-11760 (2005)
Classification of coals, three of the coal samples: coal B, C and C2 were characterised
as Bituminous Medium Rank C (0.65 - 0.78 Rr %); while coal D2 was classified as
Bituminous Medium Rank D (0.56 - 0.58 Rr %) (du Cann, 2007).
The mean vitrinite random reflectance values of the char samples are 0.67, 0.72, 0.75
and 0.56 Rr % respectively for coals B, C, C2 and D2. Vitrinite class distribution
ranges from V5 - V9 for both coals B and C and V4 - V12 and V4 - V7 for coal C2
and D2 respectively with a standard deviation of < 0.1. This is an indication that the
parent coals are single seam, non-blend coals (terminology of the ECE-UN
International codification System for Medium and High Rank Coals), except coal C2
whose vitrinite-class distribution was similar to that of a blend with standard deviation
of 0.129. Only coal C2 was observed with some abnormalities possibly caused by in
situ thermal effects with an exhibition of slightly extended vitrinite-class distribution
(du Cann, 2007). Similar findings were made by Kaitano 2007; Everson et al.
(2008b). The vitrinite random reflectance scan histograms of the coal samples are
presented in Appendix A. This property was not determined for chars due to
significant changes impacted on the coal vitrinite during the charring process at 900
C. Kaitano (2007) and Phiri (2010) reported comparable observations on their
investigated chars.
The total maceral reflectance scan analysis was also conducted on the coal and char
samples. This involved taking 250 readings on all organic components: vitrinites,
liptinites and inertinites and results obtained are also summarised on Table 3.16. The
overall assessment of the general reflectivity of the macerals show that the carbon-
rich char has attained higher varying levels of reflectance. Thus, very significant shifts
were displayed in the mean random reflectance (Rsc %) values of the chars (varying
up to 5.3 Rsc %) compared to the original coals (varying up to 1.6 Rsc %). Increasing
reflectance of the chars indicates decreasing volatiles, increasing carbon content,
increasing molecular ordering and a corresponding increase in ignition temperature
and time for burn-out (Kaitano, 2007; du Cann, 2008; Everson et al., 2008b).
Chapter 3: Coal and char Characterisation

100








Table 3.16: Reflectance properties of coal and char samples.
Sample ID Coal B Char B Coal C Char C Coal C2 Char C2 Coal D2 Char D2
Vitrinite Random Reflectance Scan
Mean vitrinite random reflectance, Rr (%) 0.67 - 0.72 - 0.75 - 0.56 -
Range of readings 0.4 - 0.8 - 0.5 - 0.9 - 0.4 - 1.2 - 0.4 - 0.7 -
Standard deviation, (%) 0.082 - 0.092 - 0.129 - 0.062 -
Vitrinite-class distribution V5 - V9 - V5 - V9 - V4 - V12 - V4 - V7 -
Rank of coal
Bit. Med.
Rank C -
Bit. Med.
Rank C -
Bit. Med.
Rank C -
Bit. Med.
Rank D -
Total Maceral Reflectance Scan
Mean random reflectance of scan, Rsc, (%) 1.27 5.28 1.30 5.19 1.63 4.43 1.15 5.07
Range of readings 0.10 - 2.8 3.7 - 6.8 0.10 - 2.6 3.4 -6.5 0.10 - 2.6 1.7 - 6.4 0.10 - 2.3 1.7 - 6.3
Standard deviation, (%) 0.542 0.588 0.492 0.596 0.382 1.127 0.458 0.715



Chapter 3: Coal and char Characterisation
101

These have profound effects on the performance of the various carbon conversion
processes (Falcon and Snyman, 1986; Snyman, 1989; Cloke and Lester, 1994; du
Cann, 2007). The total maceral reflectance scan histograms for the coal and char
samples are included in Appendix A and were found to agree very well with the
results of Kaitano (2007); Everson et al. (2008b); and Phiri (2010) on their various
studies of chars derived from inertinite-rich, high ash, South African low grade coals.


3.7.5.2 Maceral Analysis of Parent Coals
The analysis of the macerals or organic compositions and visible minerals of the
parent coals were conducted by point-count technique and results reported as volume
percent on a mineral matter basis according to ISO 7404 - 3 (1994). A summary of
this analysis including the inertinite-vitrinite ratio of the coal samples is given in
Table 3.17, while the detailed results are presented in Table 3.18. All four coal
samples were characterised as inertinite-rich with total inertinite content ranging from
54 vol. %, mmb for coal B to as high as 79 vol. %, mmb for coal C2. The total vitrinite
content was observed to be low in the samples, with coal B having the highest values
of 28 vol. %, mmb and coal C2, a very low value of 3 vol. %, mmb. Liptinite content
of the coals was also observed to be very low; < 4 vol. %, mmb in all the samples.
The ratio of inertinite to vitrinite macerals of the parent coals (on mineral matter
basis) are: 1.93, 3.00, 26.3 and 4.70 respectively for coals B, C, C2 and D2.


Table 3.17: Maceral component summary of the coal samples (vol. %, mmb).
Sample ID Coal B Coal C Coal C2 Coal D2
Vitrinite 28 19 3.0 13
Liptinite 3.0 3.0 3.0 4.0
Inertinite 54 57 79 61
Inertinite-vitrinite ratio 1.93 3.00 26.3 4.70
Chapter 3: Coal and char Characterisation

102




Table 3.18: Maceral compositions of coal samples (vol. %, mmb).

NOMENCLATURE:

VIT : Vitrinite RSF : Reactive semifusinite TI : Total inertinite
PV : Pseudo-vitrinite ISF : Inert semifusinite Total reactives = TV+TL+RSF+RINT
TV : Total vitrinite F/SEC : Fusinite / Secretinite
S/R/C : Sporinite / Resinite / Cutinite MIC : Micrinite
ALG : Alginite R INT : Reactive Inertodetrinite
TL : Total liptinite I INT : Inert Inertodetrinite
Coal ID
Vitrinite (vol. %,mmb) Liptinite (vol. %,mmb) Inertinite (vol. %,mmb) (vol. %,mmb)
VIT PV TV S/R/C ALG TL RSF ISF F/ MIC R I TI Visible Total
SEC INT INT Minerals Reactives
B 27 1 28 3 0 3 7 13 3 1 6 24 54 15 44
C 19 0 19 3 0 3 7 13 3 1 9 24 57 21 38
C2 3 0 3 3 0 3 14 17 4 2 14 28 79 15 34
D2 13 0 13 4 0 4 10 19 4 1 8 19 61 22 35
Chapter 3: Coal and char Characterisation
103

On a maceral component basis, pure vitrinite (3 - 27 vol. %, mmb), inert semifusinite
(13 - 19 vol. %, mmb) and inert inertodetrinite (19 - 28 vol. %, mmb) were found to be
most dominant in all the samples. Reactive semifusinite content was low: 14 vol. %,
mmb in all four samples. Visible minerals under the microscope were in the range of
15 to 22 vol. %, mmb for the four coals, which showed a relative similarity to the ash
content from proximate analysis except for coal C2 with a relatively lower visible
minerals and higher ash content. This was probably due to its high content of denser
inertinite maceral and some maceral-mineral associations which may introduce some
problems in the identification of the mineral grains (Kruszewska, 2003). Total
reactives which refer to the maceral constituents that possess the propensity to react to
heating ranged from 35 to 44 vol. %, mmb for all the coal samples. Kaitano (2007);
Everson et al. (2008b); and Phiri (2010) reported similar findings.


3.7.5.3 Microlithotype Analysis of Parent Coals
Microlithotypes refers to the association of macerals within the coal matrix and is
done by more number of points (approximately 500) than the maceral analysis.
(Cameron and Hunt, 1985; Falcon and Snyman, 1986; Snyman, 1989). This was done
for a completeness of petrographic data as well as for a better understanding of the
subsequent chars properties. It is well known that the association and specific
interactions of the different maceral components has a significant influence on the
properties of the resultant chars after pyrolysis (Falcon and Snyman, 1986; Tsai and
Scaroni, 1987; Cloke and Lester, 1994; du Cann, 2008). Analysis of the different
microlithotypes present in each of the four coal samples were conducted in
accordance with ISO 7404 - 4 (1998), and the results are presented in Table 3.19.
The proportion of pure vitrinite, vitrite (vitrinite > 95%) was generally low (< 13 vol.
%, mmb) in coals B, C and D2 and just 2 vol. %, mmb in coal C2; while pure inertinite
in the form of inertite (inertinite > 95%) accounted for > 32 vol. %, mmb of the coal
macerals, with the highest concentration of 59 vol. %, mmb observed in coal C2. None
of the four coals exhibited the presence of pure liptinite- liptites (liptinite > 95%)
within the coal matrix. Apart from coal C2 with a substantially low intermediate
maceral (bi-macerals and tri-macerals) volume (8 vol. %, mmb), the other three coals:
B, C and D2 have maceral mixtures concentrations of between 21 to 27 vol. %, mmb.
Chapter 3: Coal and char Characterisation

104




Table 3.19: Microlithotype analysis of coal sample (vol. %,mmb).

Coal ID
Pure Mono-macerals (%) Intermediates macerals/ maceral mixtures (%) Carbominerite (%) Minerite (%) Total (%)
Bi-macerals Tri-macerals
Vitrite Liptite Inertite Clarite Durite Vitrinertite Tri-macerite
B 13 0 32 3 5 11 8 19 9 100
C 11 0 29 2 5 8 6 26 13 100
C2 2 0 59 0 2 3 3 24 7 100
D2 8 0 40 1 6 8 6 22 9 100

NOMENCLATURE:
Vitrite : Vitrinite > 95% Vitrinertite : Vitrinite + Inertinite > 95%
Liptite : Liptinite > 95% Tri-macerite : Vitrinite, Inertinite, Liptinite > 5%
Inertite : Inertinite > 95% Carbominerite : Total inorganic/ organic microlithotypes
Clarite : Vitrinite + Liptinite > 95% Minerite : > 60 vol. % minerals
Durite : Inertinite + Liptinite > 95%


Chapter 3: Coal and char Characterisation
105

Significant differences were not observed in the carbominerite (mineral-maceral
associations) and minerite (mineral-rich particles) relative amounts in the four
samples with values ranging from 19 to 26 vol. %, mmb and 7 to 13 vol. %, mmb
respectively. This correlates well with the ash content of the four coals from
proximate analysis where no sharp distinction was observed. There is a good
agreement between this result and the observations of Kaitano (2007) and Phiri
(2010).


3.7.5.4 Carbominerite and Minerite Analysis of Parent Coals
Carbominerite and minerite analyses were conducted for completeness of
petrographic information and to give an insight into the mineral associations and
deportments with the macerals, within the parent coals organic matrix, as well as the
visible minerals (Everson et al., 2008a). Furthermore, it was used to validate other
results relating to the mineral matter and ash components of both coal and char
samples. While the carbominerites refers to minerals associated with the macerals
(coal organic matrix), the minerites are the dispersed inorganic matter in the coal
(Falcon and Snyman, 1982; du Cann, 2007; Everson et al., 2008b).

The relative abundance of the different carbominerites and minerites expressed as a
percentage of the total carbominerite and total minerite are shown in Table 3.20. From
these results, it is clear that the carbo-argilite and carbo-silicate mineral-maceral
associations occur most commonly in all four coals as finely dispersed clay minerals.
Carbonates were observed to be intimately mixed with the macerals in carbo-ankarite
with its highest occurrence in coal D2, but its relative abundance in the coals are more
contributed by the inorganic carbonate group.

The relative occurrence of carbo-pyrite is much the same for all the samples,
indicative that most of the pyrite in the coals studied is associated with the organic
components of the coal matrix. This confirms the result of increasing elemental
sulphur contents from coals to char, reported in the ultimate analysis result (Table 3.8,
Section 3.7.1). However, low occurrence of pyrite in the mineral-rich particles of coal
C2 (0.43 vol. %, mmb) and coal D2 (0.11 vol. %, mmb) was observed. The clay and
quartz group minerals occurred most commonly in all the coals and were frequently
intimately associated with the coal macerals.
Chapter 3: Coal and char Characterisation
106

The higher values of carbominerites (19 - 26 vol. %, mmb) compared to minerites (7 -
13 vol. %, mmb) from Table 3.20, shows that there may be more organic than
inorganic association of minerals within the organic matrix of the coal samples.
Carbo-argilite, carbo-silicate, clay and quartz groups are the most abundant
carbominerites and minerites in the coals. This compares well with the predominance
presence of kaolinite, clay and quartz groups in the XRD minerals analysis and the
XRF results. Similar findings have been reported by Kaitano (2007); Matjie (2008);
Hattingh (2009), in their investigations on South African coals.

Table 3.20: Carbominerite and minerite results as percentage of total carbominerite
and minerite (vol. %, mmb).

NOMENCLATURE:
Carbo-Argilite : Coal + 20 to 60 vol. % clay minerals.
Carbo-Silicate : Coal + 20 to 60 vol. % quartz
Carbo-Pyrite : Coal + 5 to 20 vol. % sulphides
Carbo-Ankerite : Coal + 20 to 60 vol. % carbonates
Carbo- Minerite : Total organic/inorganic microlithotypes
Minerite : > 60 vol. % minerals


3.7.5.5 General Condition of Coal Samples

From the petrographic analyses, the four coal samples were generally fresh in
appearance. However extensive cracks and micro-fissures were very common in the
coal macerals (du Cann, 2007). This may have occurred during handling and
preparation due to the brittle nature of the coal samples, which is characteristic of
coals of this rank. Du Cann (2007) also noted that the tendency of a coal to crack may
Sample ID COAL B COAL C COAL C2 COAL D2
Carbominerite
Carbo-Argilite and Carbo-Silicate 89.5 84.6 79.2 72.7
Carbo-Pyrite 5.26 3.85 4.20 4.55
Carbo-Ankerite 5.26 11.5 16.7 22.7
Total Carbo- Minerite 100.02 99.95 100.1 99.95
Minerite
Clay and Quartz Groups 66.7 76.9 71.1 88.8
Pyrite 11.1 7.70 0.43 0.11
Carbonate Group 22.2 15.4 28.5 11.1
Total Minerite 100 100 100.03 100.01
Chapter 3: Coal and char Characterisation
107

lead to widespread development of low temperature thermal stress fractures (passive
deflagration) on charring. The presence of deflagrating particles is an added
advantage during combustion and gasification, as this results in increased fines, and
thus higher surface areas for the reactions. Signs of severe weathering was quite often
observed in coal D2 and occasionally seen in coals B, C, and C2. There is a lot of
similarity between the coals used in this study and those studied by Kaitano (2007);
Everson et al. (2008b) and Phiri (2010) as far as it pertains to the general condition of
the samples.


3.7.6 Char Carbon Forms Analysis
Heat treatment of bituminous rank coals above 350 C usually causes the maceral
components: liptinite, vitrinites and some inertinites to soften, plasticise and expand
(Falcon and Snyman, 1986; Tsai and Scaroni, 1987; Cloke and Lester, 1994; du Cann,
2008). Most of the inertinite component and minerals are not much affected. The
applied heating rate and the final temperature attained are major determinants of the
degree of expansion. Vesicles may also be formed in the reactive maceral components
as volatiles are released (Cloke and Lester, 1994; du Cann, 2008). This leads to the
formation of vacuoles and cellular structures resulting in porous chars. Chars of this
type possess increased surface areas on which reaction occurs. Du Cann (2008) noted
that most inertinites, and particularly fusinites do not soften, degasify or develop into
porous structures. Instead, they form substantial dense chars with poor ignition and
burn-out characteristics (Kaitano, 2007).

However, not all inertinites are inert and not all vitrinites are reactive. Cloke and
Lester, (1994) in their review pointed out that some inertinite components are indeed
reactive; noting that microlithotype association has a major influence on the
propensity of these maceral components to react to heating. Inertinite maceral
association with liptinites often results in the formation of porous chars during
pyrolysis (Cloke and Lester, 1994).

The char carbon forms analysis was carried out on the basis of a 500 point-count in
accordance with ISO 7404 - 3 (1994) and microscopic constituents of the samples
were differentiated on the basis of their size, colour, reflectance, morphology, degree
Chapter 3: Coal and char Characterisation
108

of anisotropy and extent of devolatilisation (du Cann, 2008). The resulting char
carbon forms were classified according to the classification system proposed by du
Cann (2008). An outline of this classification system is presented in Appendix A,
while the result of the char carbon forms analysis is given in Table 3.21 on a mineral
matter basis (mmb).

The following categories of char carbon form constituents based on the classification
system were observed as reported by du Cann (2008). Photomicrographs of the
observed components taken with a partially cross polarized reflected white light in oil
immersion are also presented in Figures 3. 9; 3.10; and 3.11. It is worthy to note that
these photomicrographs are not peculiar to any of the chars, but provides a general
illustration of observed characteristics of any of the four chars.

Category A1: Chars with frequent tiny gas pores

These refer to chars that largely retain the original coal maceral structure
with some fine pores. They originate from both coal- vitrinites and inertinites.

i. Char carbon forms from coal vitrinite: These carbon forms are
characterised by frequent tiny gas pores with no evidence of well
developed devolatilisation vesicles. A depiction of this char carbon form is
shown in the photomicrograph of Figure 3.9(i).
ii. Char carbon forms from coal inertinites: These carbon forms originate
from parent coal inertinites that had not softened, expanded or opened
up to any appreciable degree on charring, causing them to largely
preserve their original coal maceral shape and form. They are also
commonly marked by tiny open pores resulting to some fine porosity.
Features of this group of char carbon forms are shown in the
photomicrograph of Figure 3.9(ii).

Category A2 and A3: Char networks displaying varying degree of porosity
This category describes char made up of fine walled and thicker walled
networks (Figure 3.10(i)). Char carbon forms exhibiting a fine walled network
are derived from reactive-rich coal macerals that had opened up to varying
extents during charring and possesses high internal surface areas. Thicker
walled networks originate from the inerts-rich coal particles and are less
porous.
Chapter 3: Coal and char Characterisation

109



Table 3.21: Char carbon forms analysis result (vol. %, mmb).
Char ID
Category Char carbon forms (vol. %, mmb) Maceral of origin Photomicrograph (Fig.) B C C2 D2
A 1 (i). Char with frequent Tiny gas pores Vitrinites Fig. 3.9(i) 14 11 1 10
A1 (ii). Dense Char with frequent Tiny gas pores Inertinites Fig. 3.9(ii) 27 24 31 22
A2. More open network - Fine walled Reactives Fig. 3.10(i) (above) 7 8 6 4
A3. More open network - Thicker walled Inerts Fig. 3.10(i) (below) 24 22 20 28
B7. Coke- Isotropic thick walled Vitrinites Fig. 3.10(ii); 3.10(iii) 6 5 1 4
D12. Partially reacted macerals - from coal - Fig. 3.10(iv); 3.10(v); 1 1 17 2
Fig. 3.10(vi); 3.11(i)
E13. Inorganic matter - from Coal - Fig. 3.11(v); 3.11(vi) 21 29 24 30
Total (vol. %, mmb) 100 100 100 100

Groups A and B : Very highly reflecting, oxidation restricted
Group C : Relatively low reflecting, exhibiting oxidation rims
Group D : Low reflecting little or no change
Anisotropic Exhibiting optical properties of different values when viewed under crossed Nicols.
Isotropic Exhibiting optical properties that are the same in all directions when viewed under crossed Nicols.

Chapter 3: Coal and char Characterisation
110

Category B4 - B7: Coke forms
These char forms originate from vitrite, the pure vitrinite mono-maceral of the
parent coals. Due to the fact that vitrinites and other reactive maceral soften
and degasify during the charring process, pores and vacuoles are formed. This
are visualised in Figure 3.10(ii). As the released gases within the pores and
vacuoles increase in volume because of temperature effects, the softening
walls expand giving rise to an increase in volume and surface area in the
material. The coke identified in the char samples are of isotropic forms
specifically Category B7 exhibiting substantially well developed
devolatilisation vesicles and most usually quite thick coke walls which are
again frequently marked with very fine-sized open pores. A typical highly
porous isotropic coke char carbon form from reactives is shown in the
photomicrograph of Figure 3.10(iii). None of the other coke categories: fine
and medium coke -circular anisotropic and coke -incipient anisotropic
(Category B4 - B6) were identified in any of the char samples.

Category C8 - C10: Oxidation effects
Evidence of oxidation on the maceral components are usually characterised by
the occurrence of dark rims and zones on particle edges around gasification
pores. This was not observed on any of the four char samples.

Category D11: Unaffected coal
Unaffected coal refers to materials in the parent coals unaffected by heat and
was not encountered in any of the chars. Vitrinites in this category usually
exhibit typical bituminous Medium Rank C reflectance ranges.

Category D12: Partially reacted macerals
Materials classified under this category display reflectance levels above those
of the parent coal vitrinites, but quite lower than those of the fully charred
components. Partially reacted macerals are usually pitted with intermediate
reflectivities in the range of 1% - 4% Rr. This is an indication of limited
devolatilisation and porosity is relatively poorly developed. The
photomicrographs presented in Figures 3.10(iv), (v), and (vi) and 3.11(i) are
good illustrations of char carbon forms associated with partially reacted
macerals.
Chapter 3: Coal and char Characterisation
111


Category E13: Visible minerals from parent coal
Category E13 (Visible minerals from parent coal) refers to the inorganic
material derived from the parent coal and was identified in substantial
quantities (21 - 30 vol. %, mmb) in all the char samples. Figures 3.11(v) and
3.11(vi) are good description of char form containing visible minerals.

Neither inorganic matter derived from other sources (Category E14), nor
process derived depositional carbon (Category F15) were observed in any of
the char samples.

General condition of the chars
Depreciated organic materials, cracks and fissures were observed as a common
feature in all the four char samples. These are due to thermal stress fracturing
coupled with passive deflagration (du Cann, 2008). Severe weathering often
observed in the parent coals may have also contributed to the depreciation of
the organic materials in the subsequent chars.





(i) Category A1: Char with frequent tiny gas (ii) Category A1: Dense char from inertinite
pores from vitrinite which displays no major which has not opened up and increased
in change in original parent coal maceral structure porosity

Figure 3.9: Photomicrographs of different categories of char carbon forms (du Cann, 2008)
(magnification=210
5
).

Chapter 3: Coal and char Characterisation
112



(i) Category A2: Fine walled from reactives- (ii) Category B7: Isotropic coke from
rich parent coal macerals (above); Category A3: vitrinite in the parent coal with well
Relatively thicker walled from inert coal macerals developed devolatilisation pores
(below)



(iii) Category B7: Highly porous isotropic coke (iv) Category D12: Partially reacted vitrinite-
formed from reactives (i.e., vitrinite, liptinite, Restricted devolatilisation
and reactive inertinite) in the parent coal




(v) Category D12: Partially reacted particle (vi) Category D12: Maceral bands with fine
showing thermal stress fractures but little clay inclusions displaying little in change
change in maceral structure and limited structure gas pore formation

Figure 3.10: Photomicrographs of different categories of char carbon forms (du Cann,
2008) (magnification=210
5
).
Chapter 3: Coal and char Characterisation
113




(i) Category D12: Fusinite showing deformation (ii) Extensive cracking in vitrinite but little
due to compression but no increase in porosity change in maceral structure




(iii) Char from inertinite displaying severe (iv) Inertodetrinites showing passive
fissuration deflagration and disintegration giving rise
to increased surface area




(v) Fine inertinite particles associated with clay (vi) The remains of carbonates inter-grown in
minerals from carbominerite in the parent coal inertinite

Figure 3.11: Photomicrographs of different categories of char carbon forms (du Cann,
2008) (magnification=210
5
).


Chapter 3: Coal and char Characterisation
114

From the summary given in Table 3.21, all the four char samples show varying
contributions of char carbon forms resulting from different macerals that substantially
corresponds to the proportion of such macerals in the parent coals. It is also evident
that there is a larger abundance of char carbon forms from coal inertinites (Category
A1 and A3) in all the chars. Category A1 char forms were most abundant in the char
samples, with chars with tiny gas pores from inertinites contributing between 22 - 31
vol. %, mmb. Thicker walled more open network char forms from coal inerts
(Category A3) was next in abundance and were considerably high in all the char
samples ranging in value from 20 vol. %, mmb in char C2 to 28 vol. %, mmb in char
D2. The proportion of chars with frequent tiny gas pores from vitrinites ranged from
10 - 14 vol. %, mmb in char B, C and D2 with an exceptionally low value for char C2
at just 1 vol. %. Fine walled more open network char carbon forms from coal reactive
macerals (Category A2), though low in abundance, was much the same in all the chars
(4 - 8 vol. %, mmb). Minor proportions of isotropic coke derived from vitrinites
(Category B7) were identified in approximately the same amount in chars B, C and
D2 at 6, 5 and 4 vol. %, mmb respectively, with again an exceptionally low value for
char C2 at 1 vol. %, mmb only.
The very low concentration of char forms from coal reactives and vitrinites in coal C2
is a good indication that it may be the least reactive of all the chars. Partially reacted
macerals from coal (Category D12 char carbon forms) was observed to be
substantially high in coal C2 at 17 vol. %, mmb, but very minimal in chars B, C and
D2 at < 2 vol. %, mmb. It is again suspected that the relatively large amount of
partially reacted maceral in char C2 may have caused its very low concentration of
Category A1 and B7 char carbon forms. Volume of inorganic matter derived from
coal (Category E13) was substantially high (21 - 30 vol. %, mmb) in the samples,
which is a good corroboration of the high ash content from the proximate analysis
results of the coals and respective chars. Char D2 containing the highest amount of
ash (36.8 wt. %, db) from proximate analysis, has the highest proportion of Category
E13 char carbon forms, while char B with the lowest ash content (28.8 wt. %, db) was
marked with the lowest volume of inorganic materials from coal.
The char carbon forms from: fine and medium coke- circular anisotropic (Category
B4- 6), oxidised from inertinites, vitrinites and low reflecting networks (Category
Chapter 3: Coal and char Characterisation
115

C8-10), unaffected original coal vitrinites and inertinites (Category D11), inorganic
matter derived from other sources apart from coal (Category E14), and process
derived depositional carbons (Category F15) were not observed on any of the samples
and thus were not reported in Table 3.21.
Comparisons were made of the total reactive and inert macerals of the parent coals
and the total reactive and inert components and / or char carbon forms of the chars. It
is generally accepted that the total reactive components of coal comprises of total
liptinite, total vitrinite, reactive semifusinite and reactive inertodetrinite, while inert
maceral components are made up of inert semifusinite, inert inertodetrinite, fusinite /
secretinite and micrinite (Su et al., 2001; Helle et al., 2003; Kaitano, 2007; du Cann,
2007 and 2008; Phiri, 2010).
In the context of this study, total reactive char carbon forms were considered to
encompass chars carbon forms with frequent tiny gas pores from vitrinite (Category
A1(i)), fine walled more open network char carbon forms from reactives (Category
A2) and isotropic coke from vitrinites (Category B7). Total inert char components
include dense char carbon forms with tiny gas pores from inertinite (Category A1(ii))
and thicker walled more open network from inerts (CategoryA3). Thus:

RINT RSF TV TL TRC
COAL
+ + + = (3.13)
7 . 2 . ) ( 1 .( B Cat A Cat i A Cat TRC
CHAR
+ + = (3.14)
and,
IINT ISF MIC SEC F TIC
COAL
+ + + + = (3.15)
3 . ) ( 1 .( A Cat ii A Cat TIC
CHAR
+ = (3.16)

A detailed summary of this comparison is presented in Table 3.22. Generally, some of
the total reactive components (TRC) in the coals were lost in the transition to chars
than the total inert components (TIC). The TRC experienced percentage losses of 38.6,
36.8 and 48.6% in samples B, C and D2 apiece; while an unprecedented 76.5% loss
Chapter 3: Coal and char Characterisation
116

was observed in sample C2. This unexpected high loss in TRC may be due to the
presence of partially reacted macerals of about 17 vol. %, mmb in char C2, which was
neither classified as part of the TRC nor TIC. Most of the TRC losses are in form of
volatile after softening and expansion of the original coals (Tsai and Scaroni, 1987),
and passive deflagration during devolatilisation (Kaitano, 2007). The resultant higher
density and higher maceral maceral reflectance of the chars are also evidence of the
loss of lower molecular weight TRC components and / or conversion to higher
molecular weight, more aromatic and denser TIC components in the transition from
coals to chars (Cloke and Lester, 1994; du Cann 2008).

Table 3.22: Total reactive and inert components of coal and char samples (vol. %,
mmb)
Components /
Sample ID
Total Reactive Components (TRC)- Coal and Char
Coal
B
Char
B
Coal
C
Char
C
Coal
C2
Char
C2
Coal
D2
Char
D2
Total Vitrinite 28 - 19 - 3 - 13 -
Total Liptinite 3 - 3 - 3 - 4 -
Reactive Semifusinite 7 - 7 - 14 - 10 -
Reactive Inertodetrinite 6 - 9 - 14 - 8 -
Category A1 (i) Char forms
1
- 14 - 11 - 1 - 10
Category A2 Char forms - 7 - 8 - 6 - 4
Category B7 Char forms - 6 - 5 - 1 - 4
Total 44 27 38 24 34 8 35 18
% loss of TRC in chars (%) 38.6 36.8 76.5 48.6
Total Inert Components (TIC)- Coal and Char
Inert Semifusinite 13 - 13 - 17 - 19 -
Fusinite/ Secretinite 3 - 3 - 4 - 4 -
Micrinite 1 - 1 - 2 - 1 -
Inert Inertodetrinite 24 - 24 - 28 - 19 -
Category A1(ii) Char forms - 27 - 24 - 31 - 22
Category A3 Char forms - 24 - 22 - 20 - 28
Total 41 51 41 46 51 51 43 50
% gain of TIC in chars (%) 24.4 12.2 0.00 16.3
1
- See sections 3.7.5.2 and 3.7.6 and Appendix A for descriptions
Chapter 3: Coal and char Characterisation
117

The majority of the inert components of the coal rarely softens, expands or opens up
to any appreciable extent on charring (du Cann, 2008). They therefore retain most of
the features of the parent coal macerals. Unlike the TRC, the total inert components
(TIC) of the samples generally exhibited gross gains in TIC between 12.2 and 24.4%,
except for sample C2, with neither loss nor gain in TIC (0%). The gains in TIC in
samples B, C, and D2, imply that some of the TRC of the parent coals may have been
transformed to inert components char carbon forms by the pyrolysis reaction. The
observation of neither loss nor gain in TIC in Sample C2 suffixes that the parent coal
inert macerals was not affected by the charring process; which is generally unrealistic.
This can, once more, be attributed to the high volume of partially reacted maceral char
carbon form in sample C2 that was not identified as part of either the TRC or the TIC
in the subsequent char carbon forms.
Further illustrations are shown in Figure 3.12. A comparison of the parent coal
vitrinites and inertinites with their resultant specific char carbon forms was made and
the result is a decreasing concentration in the transition from coals to chars for all the
samples (Figure 3.12(i) and 3.12(ii)). The parent coals- and subsequent chars- TRC
and TIC were also compared graphically in Figure 3.13(i) and 3.13(ii). TRC decreased
from coals to chars while an opposite trend was observed for the TIC, except for
sample C2 (0%).


(i) Parent coal vitrinite and resultant (ii) Parent coal inertinite and subsequent
Category A1 (i) char forms Category A1 (ii) char forms
Figure 3.12: Comparison of parent coals macerals and their specific char carbon forms in
the chars.
0
5
10
15
20
25
30
B C C2 D2
V
o
l
u
m
e

%
,

(
m
m
b
)
Sample ID
Coal Char
0
10
20
30
40
50
60
70
80
B C C2 D2
V
o
l
u
m
e

%
,

(
m
m
b
)
Sample ID
Coal Char
Chapter 3: Coal and char Characterisation
118


(i) Comparison of coals- and chars- TRC (ii) Comparison of coals- and chars- TIC

Figure 3.13: Comparison of the total inert and reactive components in the parent coals and
the resultant chars.


3.7.7 Physical Structural Analysis: Coal and Char Samples
Characterisation of the physical structural properties of the samples were carried out
on both coal and char samples. Results of the skeletal density of the samples are
shown in Figure 3.14 and Table 3.23.


Figure 3.14: Skeletal density of coal and char samples.
0
Coal
B
Char
B
Coal
C
Char
C
Coal
C2
Char
C2
Coal
D2
Char
D2
V
o
l
u
m
e

%

(
m
m
b
)
Total Reactive Components
Total Vitrinite Total Liptinite
Reactive Semifusinite Reactive Inertodetrinite
Category A2 Char forms Category A1 (i) Char forms
Category B7 Char forms
0
20
40
60
Coal
B
Char
B
Coal
C
Char
C
Coal
C2
Char
C2
Coal
D2
Char
D2
V
o
l
u
m
e

%
,

(
m
m
b
)
Total Inert Components, TIC
Inert Semifusinite Fusinite/ Secretinite
Micrinite Inert Inertodetrinite
Category A1(ii) Char forms Category A3 Char forms
B C C2 D2
Coal 1533 1585 1606 1568
Char 1875 1904 1928 1971
0
250
500
750
1000
1250
1500
1750
2000
S
k
e
l
e
t
a
l

d
e
n
s
i
t
y


(
k
g

m
-
3
)
Chapter 3: Coal and char Characterisation
119

Even though the skeletal density of the coal samples did not show much variation
with their values between 1533 to 1606 kgm
-3
, it is important to note that it increases
with increasing inertinite contents and decreases with increasing vitrinite contents of
the coals. The higher density observed in coal C2 was thus suspected to be due to its
large content of high density inertinite macerals (79 vol. %, mmb) and lower volatile
matter content (18.9 wt. %, db). These correlate well with the results obtained by
Senneca et al. (1998) and Hattingh (2009) in their various investigations of some
South African coals. An increase in skeletal density (1875 - 1971 kgm
-3
) was
observed in the chars. This is expected since the charring process evacuates volatiles
and low molecular mass species from the coals, leaving behind heavier carbon-rich
and mineral-concentrated chars. The high skeletal density of 1971 kgm
-3
exhibited by
char D2 can be attributed to its high ash content (36.8 wt. %, db), in addition to its
high volume of inertinite char carbon forms. Unlike the parent coals, no specific trend
was found between the resultant char carbon forms and the skeletal density of the
chars which varied only slightly.

The micropore surface area was determined according to the Dubinin-Radushkevich
(D-R) method, using carbon dioxide as adsorption gas (Anderson et al., 1965; Walker
and Kini, 1965; Walker et al., 1988) on both coal and char samples. The results of
these analyses are given in Figure 3.15. A detailed summary including mono-layer
pore volume, determined by the Dubinin-Radushkevich method and average
micropore diameter, determined by the Horvath-Kawazoe method (Horvath and
Kawazoe, 1983; Jaroniec et al., 1996; Kowalczyk et al., 2002) are presented in Table
3.23.


Figure 3.15 Micropore surface area of coal and char samples.
B C C2 D2
Coal 102.4 93.42 111.5 122.7
Char 187.4 174.6 131.3 307.0
0
50
100
150
200
250
300
M
i
c
r
o
p
o
r
e

s
u
r
f
a
c
e

a
r
e
a

(
m
2

g
-
1
)
Chapter 3: Coal and char Characterisation

120




Table 3.23: Physical structural properties of coal and char samples.
Property / Sample ID
Method
Used
Coal
B
Char
B
Coal
C
Char
C
Coal
C2
Char
C2
Coal
D2
Char
D2
Micropore surface area (m
2
g
-1
) D-R 102.4 187.4 93.42 174.6 111.5 131.3 122.7 307.0
Mono-layer capacity

(cm
3
g
-1
) D-R 22.42 41.02 20.45 38.22 24.4 28.75 26.87 67.20
Average micropore diameter

() H-K 4.161 3.651 4.115 3.724 4.111 3.868 3.935 3.613
Micropore volume x10
-2
(cm
3
g
-1
) H-K 2.167

5.283 2.047 4.486 2.389 3.101 3.377 6.636
Skeletal (Apparent) density

(kg m
-3
)
HP 1533 1875 1585 1904 1606 1928 1568 1971
Porosity

(%) (D
p
5) AD 2.882 9.970 2.906 8.247 3.346 5.188 5.424 12.75

NOMENCLATURE:
D-R : Dubinin-Radushkevich method
H-K : Horvath-Kawazoe method
HP : Helium Pycnometry
AD : CO
2
Adsorption data




Chapter 3: Coal and char Characterisation
121

The micropore surface areas of the coals did not exhibit much variations with values
ranging between 93.42 m
2
g
-1
for coal C and 122.7 m
2
g
-1
for coal D2. The softening,
plasticizing, expansion and the release of volatiles from the parent coal (Falcon and
Snyman, 1986; Tsai and Scaroni, 1987; Cloke and Lester, 1994; du Cann, 2008)
during the charring process resulted in the opening of closed, blind and inaccessible
pores (Laurendeau, 1978); formation of vacuoles, vesicles and cellular structures
(Cloke and Lester, 1994; du Cann, 2008) and the enlargement of pores (micropores to
mesopores and mesopores to macropore) (Laurendeau, 1978). The combined impact
of these is an increase in surface area and porosity as observed in the chars. Unlike the
parent coals, sharp differences were observed in the micropore surface areas of the
chars. The micropore surface areas of char B and C were very close to each other at
187.4 and 174.6 m
2
g
-1
. Char C2 exhibited a very low micropore surface area,
resembling that of parent coal at 131.3 m
2
g
-1
. This is possibly due to the large amount
of inertinite maceral (not easily affected by heat) in the parent coal (79 vol. %, mmb)
and high volume of partially reacted macerals in the char (17 vol. %, mmb). Char D2
exhibited an exceptionally high micropore surface area of 307 m
2
g
-1
up from 122.7
m
2
g
-1
of the precursor coal. This may be due to its higher microporosity (Walker et
al., 1988). The mono-layer capacity is a function of the surface area and varies
accordingly with it. Phiri, (2010) made similar findings. However, it was observed
that the increasing order of the micropore surface area in the coals (coals C < B < C2
< D2), changed in the chars (chars C2 < C < B < D2). This shows that the parent coals
are different in terms of porosity to the subsequent chars; although there is an inter-
play of many factors as far as char properties is concerned (Franklin, 1951; Cloke and
Lester, 1994; Davis et al., 1995; Russell et al., 1999; Lu et al., 2002a and 2002b).

It is generally accepted that carbon dioxide adsorption gives a better characterisation
of micropores < 20 (Anderson et al., 1965; Walker and Kini, 1965; Walker et al.,
1988) as compared to nitrogen adsorption (Garrido et al., 1987). This is due to the
faster diffusion of carbon dioxide into finer micropores at 273 K than nitrogen
diffusion at 77 K (Walker et al., 1988; Garrido et al., 1987). The adsorption isotherm
plots of both coals and chars are shown in Figure 3.16. Comparison of the adsorption
isotherms of the parent coals and the respective subsequent chars are further
illustrated in Figure 3.17. The higher quantity of adsorbed CO
2
in the chars as
compared to the coals is indicative of the higher D-R micropore surface areas attained
Chapter 3: Coal and char Characterisation
122

by the chars. Despite the fact that the micropore surface area of the chars increases
with decreasing skeletal density with respect to chars B, C and C2; which was not
observed for the parent coals; char D2 was completely out of trend in this correlation
and was thus, not reported in this dissertation.


(i) CO
2
adsorption isotherm plots for coal samples (ii) CO
2
adsorption isotherm plots for char samples

Figure 3.16: CO
2
adsorption isotherm plots for coal and char samples.

The average micropore diameter as determined by the Horvath-Kawazoe method was
found to generally decrease from the coals (3.935 - 4.161 ) to the chars (3.613 -
3.868 ). This is due to opening of closed and inaccessible micropores due to thermal
treatment and the release of volatiles, while the highest value of 3.868 observed for
char C2 may be attributed to the lowest volatile matter content of its parent coal (18.2
wt. %, adb) and its higher content of partially reacted macerals in the char. The pore
volume (micropores: D
p
5) (Micromeritics, 2006) increased from coals to chars
and the increments are consistent with the increase in the micropore surface area. It
can be seen from Table 3.23 that char D2 has the highest adsorbate pore volume; a
good indication of its exceptional surface area with respect to the other three chars.
Micro-porosity of both coal and char samples (D
p
5 ) determined from CO
2

adsorption data show that the parent coals exhibited very little porosities (< 5.5%)
with coal B having the lowest porosity of 2.9%. Although porosity generally
increased from coals to chars, porosity values of the chars are quite low as well. This
is probably due to the high inertinite maceral content of the parent coals (Laurendeau,
0
4
8
12
16
20
0 0.01 0.02 0.03
Q
u
a
n
t
i
t
y

o
f

a
d
s
o
r
b
e
d

C
O
2
(
c
m
3

g
-
1
)
Relative Pressure (P/P
0
)
Coal B Coal C Coal C2 Coal D2
0
5
10
15
20
25
30
35
0 0.01 0.02 0.03
Q
u
a
n
t
i
t
y

o
f

a
d
s
o
r
b
e
d

C
O
2
(
c
m
3

g
-
1
)
Relative Pressure (P/P
0
)
Char B Char C Char C2 Char D2
Chapter 3: Coal and char Characterisation
123

1978; Falcon and Snyman, 1986; Snyman and Botha, 1993), and the relatively high
TIC components in the chars (Cloke and Lester, 1994; du Cann, 2008). The low
porosity of char C2 is probably due to the precursor coals very low volatile matter
content (Tsai and Scaroni, 1987; Cloke and Lester, 1994) and a higher volume of
dense and partially reacted maceral in its char carbon forms. It was also found that the
increasing order of porosity in the coals (coal B < C < C2 < D2) changed in the chars
(char C2 < C < B < D2). This further strengthens the fact that there are significant
variations between coal and char properties with respect to porosity, even though; the
later is derived from the former.


Figure 3.17: Comparison of the CO
2
adsorption isotherm plots for coals and chars: B, C,
C2 and D2.

0
5
10
15
20
25
30
0 0.01 0.02 0.03
Q
u
a
n
t
i
t
y

o
f

C
O
2
a
d
s
o
r
b
e
d

(
c
m
3

g
-
1
)
Relative Pressure (P/P)
Sample B
Coal B Char B
0
5
10
15
20
25
0 0.01 0.02 0.03
Q
u
a
n
t
i
t
y

o
f

C
O
2
a
d
s
o
r
b
e
d

(
c
m
3

g
-
1
)
Relative Pressure (P/P)
Sample C
Coal C Char C
0
3
6
9
12
15
18
0 0.01 0.02 0.03
Q
u
a
n
t
i
t
y

o
f

C
O
2
a
d
s
o
r
b
e
d

(
c
m
3

g
-
1
)
Relative Pressure (P/P)
Sample C2
Coal C2 Char C2
0
5
10
15
20
25
30
35
0 0.01 0.02 0.03
Q
u
a
n
t
i
t
y

o
f

C
O
2
a
d
s
o
r
b
e
d

(
c
m
3

g
-
1
)
Relative Pressure (P/P)
Sample D2
Coal D2 Char D2
Chapter 3: Coal and char Characterisation
124

3.8 Summary
Analysis of the char yield after char production clearly shows that the mass of the
resultant char produced from each batch of the process is equivalent to the sum of the
fixed carbon, and ash contents only, of the parent coals. This clarifies that, the
moisture and volatile matter contents of the coals are almost evacuated completely by
the charring process.
The original four coals were characterised as high ash coals, with ash contents in the
range of 26.7 to 34.5 wt. %, db, while the respective chars exhibited higher ash
content values after pyrolysis with values between 28.8 to 36.8 wt. %, db. All the coal
samples were relatively low in volatile matter (<24 wt. %, db), which were
significantly driven off during the charring process, with the chars having VM values
between 0.81 to 3.30 wt. %, db. It should be noted that, although significant
devolatilisation took place during char production, most of the char carbon forms
(especially the inert char carbon forms) did not increase in porosity to any
appreciable degree under the petrographic miscroscope. Moisture content went down
to 1.58 wt. %, adb in the chars from the precursor coals values of 3.40 to 6.50 wt.
%, adb; while fixed carbon in the coals (42.9 - 49.4 wt. %, db) went up to values
between 60.2 to 69.4 wt. %, db in the chars. The gross calorific value of the parent
coals, though low, was found to be approximately the same within the small range of
18.1 MJkg
-1
- 21.4 MJkg
-1
.
Elemental carbon content was the same for coals B, C and C2 at 73.1 wt. %, daf;
while coal D2 has a lower value at 68.1 wt. %, daf. The elemental hydrogen-carbon
and the oxygen-carbon ratios for the samples were rather low and were reflected in
the higher aromaticity and lower fraction of amorphous carbon in both coals and
chars.
From petrographic analysis, the mean vitrinite reflectance of the parent coals ranged
from 0.56% to 0.75% Rr, hence the coal samples were characterised as bituminous,
medium rank C (coals B, C and C2) and bituminous, medium rank D coals (coal D2),
according to the ISO 11760 method (2005). The four coals total maceral reflectance
scans had values ranging from 1.15 to 1.63% Rsc. The devolatiliastion process shifted
the total maceral reflectance of the subsequent chars to substantially higher values
Chapter 3: Coal and char Characterisation
125

within the narrow range of 4.43 to 5.28% Rsc. The coals were characterised as
inertinite-rich with inertinite maceral content between 54 to 79 vol. % mmb and
inertinite-vitrinite ratios of: 1.93, 3.00, 26.3, and 4.70 respectively for coal B, C, C2
and D2.
It was observed from the microlitotype analysis, that the pure mono-maceral, inertite
was the dominant maceral component in all four parent coals (>32 vol. %, mmb), with
a value as high as 59 vol. %, mmb for coal C2. The liptinite mono-maceral, liptite was
not identified in any of the coal samples.
It was deduced from the carbominerite and minerite analysis that both organic and
inorganic association of minerals existed with the macerals of the coal samples. The
coal samples were pitted with higher fractions of carbominerites (19 - 26 vol. %,
mmb) relative to the minerite fraction at 7 - 13 vol. %, mmb. The carbominerite and
minerite analysis results were consistent with the results from XRD mineral analysis
of the parent coals and the XRF ash analysis of the chars with the dominant presence
of kaolinite, clay and quartz species in all the samples.
The transition from coals to chars at 900 C, led to the formation of various char
carbon forms in the chars. Coal reactive macerals formed minor proportions of
isotropic coke and fine char networks; while some vitrinites and the majority of the
coal inertinites resulted in denser char carbon forms, which formed a greater volume
of the subsequent chars having a marked increase in random maceral reflectance
values. The total reactive components (TRC) generally decreased from coals to chars,
with losses >36.5%. The total inert components (TIC) exhibited a gain in the
transition from coals to chars (12.2 to 24.4% loss), except char C2 with 0% gain. This
may have resulted from the conversion of some of the reactive macerals in the coals
to inert char carbon forms during the char formation process.
Demineralisation of coals and chars prior to XRD carbon crystallite analysis gave
higher demineralisation efficiencies on the parent coals (96.7 - 98.3%) than on the
chars (89.1 - 93.3%). The XRD diffractograms of both the demineralised coal and
char samples was observed to posses graphitic features characteristically similar to
that of graphite by the presence of (002), (10), and (11) peaks.
Chapter 3: Coal and char Characterisation
126

With respect to the carbon crystallite analysis, it was observed that the chars
possessed a more compact, better aligned and more structurally ordered carbon
content than their respective original coals. Lattice parameters: crystallite height, L
c
,
and crystallite diameter, L
a
, decreased, while the inter-layer spacing, d
002
, increased
from coals to chars. Thus the carbon structure of the chars were more condensed and
smaller that that of the parent coals. Aromaticity was observed to increase; while the
fraction of amorphous carbon decreased from coals to chars. It could also be inferred
from the XRD carbon crystallite analysis results that the structural disorderliness of
the chars increases in the order: Char C2 < Char C < Char B < Char D2. Hence, the
degree of disorder index, DOI, decreased from char D2 down to char C2.
The charring process also impacted on the physical structural properties of the
subsequent chars. The D-R micropore surface areas increased from values of 93.42
m
2
g
-1
to 122.7 m
2
g
-1
for the coals to 131.3 m
2
g
-1
to 307.0 m
2
g
-1
in the chars; while
the H-K average micropore diameters decreased from coal to chars. Skeletal density
was found to generally increase from coals (1533 - 1606 kgm
-3
) to chars (1875 - 1928
kgm
-3
) and was probably influenced significantly by the inertinite content of the
parent coal. Apart from maceral content, ash content also affected the skeletal density
as char D2 with the highest ash content has the highest density. Char D2 also
exhibited an exceptionally higher D-R micropore surface area with respect to the
other chars.
Microporosity of both coals (2.882 - 5.424%) and chars (5.188 - 12.75%) were rather
low, which can be attributed to the high volumes of total inert components (TIC) in
both coals and chars. This is because TICs do not react to heating easily. The high
volume of partially reacted maceral (char carbon form) in char C2 (17 vol. %, mmb)
and the parent coals lower volatile matter content (18.2 wt. %, adb) may have
contributed to its very low porosity with respect to the other chars. Average micropore
diameter and micropore volume were observed to generally decrease from coals to the
subsequent chars.
Mineral analysis by XRD analysis showed that significant mineral interactions and
transformations occurred in the coal to char transition at 900 C. Out of 16 crystalline
mineral phases identified in the coal and char samples, only quartz and calcite was
observed in both the coals and chars. Kaolinite and quartz were the dominant mineral
Chapter 3: Coal and char Characterisation
127

phases identified in the coals; while quartz, cristobalite and microcline formed greater
proportions of minerals in the char samples. However, significant variations were not
observed in the relative abundance of the various mineral phases in the coals and
respective chars as their values vary within a small range.
Ash components analysis revealed that SiO
2
and Al
2
O
3
were the dominant inorganic
species in the residual ash of the chars. Other inorganic species of catalytic
importance detected in the ashes of the chars are: Fe-, Ca-, Mg-, Na- and K- species.
The alkali index of the chars based on the detected inorganic components in
increasing order of magnitude was: 2.55 for C2; 5.92 for char B; 6.15 for char C; and
7.17 for char D2.
On the condition analysis of the samples, the coal samples were found to be generally
fresh in appearance, with frequent micro-fissures in macerals which are evidence of
weathering. Cracks, fissures and deteriorated organic material were common in all
four chars. These features common to the chars may be attributed to thermal stress
fracturing and passive deflagration.
From the characterisation results, it was observed that apart from the maceral contents
and the inertinite-vitrinite ratios, that coal B, C, and C2 did not exhibit sharp
variations in their investigated properties; while coal D2 was significantly different.
However, the subsequent chars exhibited substantial variations in both their chemical
and physical structural characteristics. The increasing order of magnitude of the
micropore surface area, microporosity, fraction of amorphous carbon, and the extent
of disorderliness of the carbon crystallite; all changed in the transition from parent
coals to the subsequent chars. This shows that coals are comparatively different to the
chars derived from them.






128













CHAPTER 4 44 4





4.0 EXPERIMENTAL: CHAR GASIFICATION
WITH CARBON DIOXIDE

4.1 Introduction
The experimental methodology and procedures employed to study the effects of
various chemical and physical properties of the selected chars on the char-CO
2

gasification kinetics are presented in this chapter. A description of the thermal
analysis equipment, specifications and experimental design and schedule are
discussed. Materials and reaction gases used in the course of the experiments are also
presented.
Chapter 4: Experimental: Char Gasification with Carbon Dioxide
129

4.2 Materials Used
Materials used in this study are coal/char and reaction gases, which are carbon dioxide
(CO
2
) and nitrogen (N
2
).


4.2.1 Coals and Subsequent Chars
The four low grade coal samples used for this investigation was selected from the coals
supplied by Eskom to the Coal Research Group, North-West University,
Potchefstroom Campus and originated from coalfields in the Highveld area of South
Africa. The particle size of the samples as received, ranged from fine powder to as
large as 5 cm.
These samples were prepared according to methods described in Section 3.3 and
devolatilised following the procedure outlined in Section 3.4.1, to prepare the
subsequent chars. The prepared chars have average particle diameters of 1000 m < dp
< 1120 m. The coal and char samples were identified as: coals- or chars- B, C, C2
and D2 respectively. Characterisation of the coals and the resultant chars were
presented in Chapter 3.


4.2.2 Reactant Gases
Reaction gases used for the char-CO
2
gasification reaction experiments were mixtures
of CO
2
and N
2
(See Table 4.3, Section 4.5 for reaction gas composition). Afrox
supplied both gaseous reagents. N
2
was used as inert diluent gas to control the
concentration of CO
2
, and a high flow rate of the mixed gases was used to eliminate
external diffusion effects during experimentation. The specifications of the two
reactant gases are summarised in Table 4.1.


Table 4.1 Specifications of gaseous reagents.
Reagent Afrox item No. Grade Purity
Carbon dioxide, CO
2
40-RC Technical 99.0%
Nitrogen, N
2
511204-SE-C Nitrogen Baseline 5.0 > 99.999%
Chapter 4: Experimental: Char Gasification with Carbon Dioxide
130

4.3 Reactivity Equipment: Thermogravimetry
A thermogravimetric analyser (TGA) was used to investigate the char-CO
2

gasification reactivity of the respective char samples. Dutta et al. (1977); Mhlen et
al. (1985); Matsui et al. (1987); Mhlen and Sulimma (1987); Kajitani et al. (2006);
Kaitano (2007); Everson et al. (2006 and 2008a), to mention a few investigators, have
successfully used TGAs in the study of gas-solid reaction kinetics. A detailed
discussion on thermogravimetry was presented in Section 2.7.1.


4.3.1 Thermax 500 Thermogravimetric Analyser (TGA)
A Thermax 500 TGA supplied by Thermo Fischer Scientific, Karlsruhe, Germany
was used. The TGA system is illustrated schematically in Figure 4.1 and a photograph
of the system is given in Figure 4.2. Like other TGAs, it measures weight change of a
sample over a given temperature and pressure range under specified environmental
conditions. The specifications of the TGA are presented in Table 4.2.
The sample is loaded into a miniature platinum gauze basket, which is connected to
and suspended from a sapphire extension rod or hang-down after opening the joint
coupling, joining the furnace vessel and the pressure balance. The controlled
temperature zone of the furnace is 50 mm. Furnace temperature follows a temperature
profile sequence established by the user in a method in the Thermax software. The
software periodically records time, weight, temperature and pressure according to
user-defined sequence to a file on the computer and allows the user to perform
analytical functions on the experimental data as well as save to other formats that can
be processed on other applications for various uses.
The Thermax 500 TGA is composed of three main parts: the mainframe, console and
the pressure and flow controllers. The mainframe holds the equipments hardware
including the pressure balance, the furnace vessel, the stand and the elevator in place
as shown in Figure 4.2. The console houses the equipment electronics, displays data
and warnings and connects the facility to the computer. The pressure and flow
controllers controls pressure in the instrument, flow rates for the three different gas
Chapter 4: Experimental: Char Gasification with Carbon Dioxide
131

channels, indicates pressure in the instrument and connects to the computer via an RS
232 serial transmission cable.


Figure 4.1: Schematic representation of Thermax 500 TGA showing the essential parts and
gas flow system (Not drawn to scale).


The Pressure Balance: The microbalance used by Thermax 500 TGA is a pressurised
balance consisting of two parts: a D-101 balance and a pressure head with a capacity
of 100 g and sensitivity of 1 g. However, the constraints of high pressure will usually
limit the effective sensitivity to about 100 g. Specifications of the pressure balance
are given in Table 4.2. The balance also has a closed loop servo network, which
Pressure balance
Pressure balance
counter weight
Purge gas (CO
2
+ N
2
)
Reaction gas (CO
2
+ N
2
)
Furnace gas (N
2
only)
Sapphire rod hang-down
Joint coupling
Furnace vessel
Temperature controlled zone
Platinum gauze basket
Thermocouple
Reaction gas
Purge gas
Furnace gas
Flue gas
Chapter 4: Experimental: Char Gasification with Carbon Dioxide
132

automatically compensates for weight changes in the sample. The electrical current
necessary to return the balance beam to its null position is directly proportional to
weight. Since this is a null balance, the sample always remains in a uniform
temperature zone.




Figure 4.2: Photograph of Thermax 500 TGA showing the essential parts.

The Furnace Vessel: The furnace vessel, which was machined from solid 316
stainless steel, was constructed and inspected according to the American Society of
Mechanical Engineers (A.S.M E) boiler and pressure vessel Code with the acceptance
stamp on the side of the bottom plate. Black anodised spacers minimize dead volume
in the balance chamber to approximately 150 cm
3
. Specifications of the furnace are
presented in Table 4.2. An internal quartz tube, sealed at both ends with O-rings,
separates the reaction chamber from the furnace and prevents reactive gases from
entering the furnace. A joint coupling ring was used to seal the furnace vessel to the
pressure balance. The sample is loaded and unloaded when the joint is opened and the
Pressure balance
Mainframe
Joint coupling
Thermax 500 console
Furnace vessel
Mainframe tripod stand
Elevator
Mass flow meter
Chapter 4: Experimental: Char Gasification with Carbon Dioxide
133

furnace is lowered using the elevator. A cooling fan that is suspended from the
balance isolates the microbalance from the heat generated by the furnace. This fan
assures a stable reading from the balance. The fan is computer controlled by the
Thermax software; however, a fan switch on the console allows for manual control as
well.

Table 4.2: Thermax 500 TGA specifications.
Pressure Balance
Maximum load 100 g
Weighing range 10 g
Sensitivity 1 g
Temperature drift 10 gC
-1

Accuracy at ambient temperature >99.98 %
Repeatability >99.999 %
Full range zeroing Yes
Furnace
Temperature range
At ambient pressure Up to 1100 C
At 69 up to 100 bar Up to 1000 C
Maximum ramp 25 Cmin
-1

Thermocouple K- type (1/8" Chromel-Alumel)
Temperature repeatability 3 C
Pressure Level
Maximum pressure at ambient 100 bar
Maximum pressure at 1000 C 69 bar
Vacuum 5 x 10
-4
torr
Reaction chamber
Sample Volume 35 cm
3

Reactor tube volume 300 cm
3

Controlled temperature zone 50 mm
Oxygen free environment Yes
Atmospheres O
2
, H
2
, CO, CO
2
, H
2
S, SO
2
, H
2
O
Chapter 4: Experimental: Char Gasification with Carbon Dioxide
134

The Thermax software (according to an established temperature profile), controls the
temperature of the furnace. A standard type-K thermocouple located close to the
sample detects and conveys the current temperature to the computer as well as the
console. The furnace allows a temperature profile of ramps and isotherms, established
by the user under constant computer control. The ramps are adjustable from 0.1
Cmin
-1
up to 25 Cmin
-1
depending on the users intention.
The Stand and Elevator: The stand is a rigid tripod support with a plate and leveling
screws for aligning the pressure balance. The legs of the tripod must make firm
contact with the floor and resist tilting and shifting. In cases where there is extensive
vibration (up to 17 Hz), a rubber mat or other isolation system may be required so that
the pressure balance is not affected. The elevator allows easy raising and lowering of
the furnace vessel and is controlled by a hand release system. Manual control is
sometimes required to align the components. Wheels on the elevator allow the unit to
be moved. However, it is recommended that the lift be fixed using the adjustable feet
to align the lift mechanism with the pressure balance.


4.3.2 Gas Supply
The reaction gas mixture (CO
2
and N
2
) premixed in various concentrations as required
for the experiment was supplied to the furnace vessels reactor section by means of
interconnected stainless steel piping. The flow rate of the furnace gas was controlled
by the furnace gas mass flow controller in the TGA console, while that of the purge
and reaction gases were controlled with a control unit connected to two Brooks-5850
mass flow controllers (for mixing) before connecting to the appropriate gas channel of
the TGA console. Furnace gas was N
2
only, while the same gas mixture was used for
both the purge gas and the reaction gas.


4.3.3 Data Acquisition Interface
A computer is required to operate the Thermax 500 TGA with the Thermax software.
The test criteria, established by the user in the software are displayed on the monitor,
Chapter 4: Experimental: Char Gasification with Carbon Dioxide
135

allowing the user to fully consider them before starting the experiment. The data
acquired during the experiment (mass, current temperature and pressure) is shown on
the computer display in real time and stored in computer files, which can be exported
to external storage devices for further processing.


4.4 Experimental Procedure
50 1 mg of sample (1000 m < d
p
< 1120 m) was used for each experiment on the
TGA. N
2
gas flow rate of 750 mlmin
-1
was used for the furnace gas, while the purge
gas and reaction gas composed of CO
2
and N
2
in the required concentrations were also
supplied to the TGA at 750 mlmin
-1
. This was done to eliminate any dilution effect on
the reaction gas composition that may arise if N
2
only was used for the purge gas.
Furthermore, CO
2
does not have significant corrosive side effects and thus was not
feared to affect the pressure balance mechanism and electronics.
Procedures for the experiments are as follows:
All the required settings on the TGA: initial dynamic ramp that was 25 Cmin
-1

for all the experiments, final isothermal temperature (900 - 950 C), and expected
duration of the experiment: set to 50 hours (180,000 seconds) were set. Pressure
mode was set off and TGA mode put on, as all the experiments were done at
ambient pressure (0.875 bar in Potchefstroom). The joint coupling on the TGA
was opened and the furnace vessel lowered using the manual leverage control.
The empty platinum gauze sample basket was cleaned of any particles using
compressed air, and then hung on the sapphire rod sample hang-down, and the
pressure balance was tared from the Thermax software on the computer.

50 1 mg of char sample was weighed on the Mettler Toledo AB204-S top
loading balance with 0.001 mg sensitivity. The sample basket was removed
gently from the hang-down, the weighed char sample was then loaded and hung
back on the sapphire rod holder. The pressure balance was cross checked for any
misalignment. If any was observed; the alignment knob will be used to correct it,
otherwise; the furnace vessel was raised to touch the pressure balance and joint
Chapter 4: Experimental: Char Gasification with Carbon Dioxide
136

coupling put back and tightened with six hexagonal bolts. Sample mass displayed
on the TGA console is supposed to be the same with that obtained from the
Mettler Toledo balance, or else the process was restarted.

The TGA furnace was degassed by flushing with N
2
gas for 15 minutes on the
three different gas channels. After this, the Themax software was used to start the
experiment after entering a run identificator label. The data logging programme
controls the TGA in line with the preset ramp, isotherm and run duration. Traces
of volatiles and or moisture (probably adsorbed by the char during and or
between storage and usage) may be observed during the ramp and initial
isothermal period. The TGA usually ramps to the isothermal temperatures (900,
917, 933 and 950 C) in 35 to 37 minutes. The char was left for an extra 23 to 25
minute (one hour in total) to make sure that there was no observable mass loss
due to any volatile or moisture, at the stipulated isothermal temperature.

After confirming the constant mass of the char at the required isothermal
temperature has been attained, the reaction gas (CO
2
+ N
2
or CO
2
only for 100 %
CO
2
experiments) was switched over manually. The purge gas, which is of the
same composition as the reaction gas was switched over as required. The reaction
was allowed to proceed until there was no apparent change in mass of the
reacting char sample; an indication that all the reactive carbon and carbon
compounds in the char are fully reacted and that the unreacting or
uncombustible residue corresponds to the ash. The ash primarily consists of the
mineral matter but may also contain some unreactive carbons.


4.5 TGA Experimental Programme
Experiments were planned to include two different variables: isothermal reaction
temperature and CO
2
concentration in the reaction gas mixture. Four different char
samples were used. The experimental data from the TGA Thermax software were
evaluated in detail to determine the kinetics of the gasification reaction as well as the
contribution(s) of the various char properties on the overall reaction. These results
Chapter 4: Experimental: Char Gasification with Carbon Dioxide
137

were used to deduce a suitable kinetic model to fit the experimental data. The outline
of experimental design followed in this investigation is given in Table 4.3. It should
be noted that all experiments were conducted at ambient pressure (0.875 bar).


Table 4.3: Reaction conditions for char-CO
2
gasification experiments.

Variable Ranges and compositions
Reaction Gasification of char with CO
2

Char samples B, C, C2 and D2
Char production temperature (in PBBR) 900 C
Char sample mass 50 1 mg
Char particle diameter 1000 < d
p
< 1120 m
Initial ramp (N
2
atmosphere) 25 Cmin
-1

Isothermal reaction temperatures 900, 917, 933 and 950 C
Reaction gas flow rate 750 mlmin
-1

Reaction gas mixture (CO
2
+N
2
) 25, 50, 75 and 100 % CO
2

Reaction Pressure Ambient (0.875 bar)
Number of Experiments without reproducibility 64 experiments














138







CHAPTER 5 55 5




5.0 GASIFICATION WITH CARBON DIOXIDE:
RESULTS AND DISCUSSION


5.1 Introduction
The results of the char-CO
2
gasification experiments are presented in this chapter. The
chars were prepared at 900 C according to a procedure described earlier in Section
3.4.1. The gasification reaction experiments were carried out in the Thermax 500
TGA following the method already discussed in Section 4.4. The main focus of this
chapter is: (i) to show the relative impact of the operating conditions (temperature,
CO
2
concentrations and different chars) on the carbon conversion; and (ii) to correlate
the reactivity results with the properties of the coals and chars as obtained from the
detailed characaterisation results of Chapter 3.
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

139

5.2 Normalisation of the Experimental Results
Results recorded by the TGA are in a format showing the relative loss in mass, time
and temperature. A typical experimental result obtained from the TGA is shown in
Figure 5.1.


Figure 5.1: Typical mass loss curve for char C2 at 900 C, 100% CO
2
, 0.875 bar.

A rapid decrease in mass was observed during the initial stage of the reaction until
about 50% mass loss. A slower rate of consumption was observed from this point
forward, until an equilibrium stage is reached which indicates that all the carbon
compounds in the char have fully reacted with CO
2
. It was observed that the final
equilibrium mass value corresponded well with the ash content of the chars as
determined from proximate analysis (Table 3.8, Section 3.7.1). The mass of reacted
carbon also corresponded to the fixed carbon value from the proximate analysis.
However, a maximum deviation of 5% from the ash and fixed carbon percentages
were observed in most of the experiments. These slight deviations may be attributed
to the heterogeneity of the individual char particles which was not fully eliminated,
even with the use of a rotary sample splitter to obtain a representative sample.
Furthermore, the unreactive or unreacted highly aromatic parts of the carbon matrix
may as well introduce some anomalies in the results.
0
100
200
300
400
500
600
700
800
900
0
10
20
30
40
50
60
0 200 400 600 800 1000 1200
T
e
m
p
e
r
a
t
u
r
e

(

C
)
C
h
a
r

m
a
s
s
,

m

(
m
g
)
Time, t (min)
Mass Temperature
Ash
Carbon
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

140

Reaction time for the experiments varied from 12 hours to over 48 hours at lower
concentrations of CO
2
. Kaitano (2007) also experienced very long reaction time with
chars prepared from inertinite-rich low grade coals, similar to chars used in this
investigation.
For a quantitative study of the conversion of the fixed carbon in the char only, the
TGA raw data were normalised on an ash free basis on the assumption that negligible
product gas is produced from mineral transformations or mineral reactions with CO
2

according to Equation 5.1 (Dutta et al., 1977; Hampartsoumian et al., 1993; Kyotani
et al., 1993; Ye et al., 1998; Zhang et al., 2006; akal et al., 2007; Kaitano, 2007;
Zhang et al., 2007; Everson et al., 2006 and 2008a; Wu et al., 2008):

ash o
t o
m m
m m
X

= (5.1)

A typical normalized result (carbon conversion versus time plot) from the
experimental result of Figure 5.1, is presented in Figure 5.2. The number of data
points used throughout this investigation was 29 to 31 and is considered accurate
enough for all interpretations as well as kinetic modelling.


Figure 5.2: Conversion-time plot for char C2 at 900 C, 100% CO
2
, 0.875 bar.
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

141

5.3 Reproducibility of the Experimental Results
The experimental results from the TGA were crosschecked for reproducibility. This
was done by conducting three different experiments with each of the respective char
samples at the same operating conditions of temperature and CO
2
concentration. The
characteristic experimental data were analysed for initial reactivity, R, structural
parameter, , time factor t
f
, and time for 50% and 90% conversions (t
0.5
and t
0.9
).
Experimental error in terms of the standard deviation and percentage average deviation
were determined for each set of three experiments. From the results presented in
Appendix B, it was found that the average deviation for all the determined parameters
was < 2% for chars B and C; < 3% for char C2; and < 4% for char D2. Thus,
experimental results from the TGA are reproducible. The distinct plots from the
reproducibility tests for each of the chars are also included in Appendix B.


5.4 Effect of Operating Conditions on Char-CO
2
Gasification
Reactivity
Apart from using four different char samples for the experiments, the isothermal
temperature of reaction and CO
2
concentration were varied. The resultant effects of
these variables are discussed in this section. It should however be noted that only a
few of the results are shown here; more results are presented in Appendix B.


5.4.1 Effect of Isothermal Temperature of Reaction
The relative impact of the isothermal temperature of the char-CO
2
reaction (varied
between 900 C and 950 C) is illustrated using the conversion-time plots in Figure
5.3. From the results, it is clear that that there is a relative increase in carbon
conversion with increasing isothermal reaction temperature. This shows that the char-
CO
2
gasification reaction follow Arrhenius type kinetics. These tendencies were
observed in all the chars. Similar results were also obtained for different CO
2

concentrations in the reaction gas mixture.
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

142

The same trend was observed by other investigators on char-CO
2
gasification
reactions (Dutta et al., 1977; Matsui et al., 1987; Radovic et al., 1985; Harris and
Smith, 1990; Miura et al., 1990; Hampartsoumian et al., 1993; Ochoa et al., 2001; Fu
and Wang, 2001; Sina et al., 2003; Murillo et al., 2006; Zhang et al., 2006; Kajitani
et al., 2006; Zhang et al., 2007; Kaitano, 2007; Everson et al., 2006 and 2008a;
Sangtong-Ngam and Narasingha, 2008; Wu et al., 2008; Hattingh 2009).


(i) Effect of temperature on the CO
2
reactivity of (ii) Effect of temperature on the CO
2
reactivity of
char B at 50% CO
2
, 0.875 bar char C at 75% CO
2
, 0.875 bar

(iii) Effect of temperature on the CO
2
reactivity of (iv) Effect of temperature on the CO
2
reactivity of
char C2 at 25% CO
2
, 0.875 bar char D2 at 100% CO
2
, 0.875 bar
Figure 5.3: Effect of temperature on the CO
2
reactivity of the chars at different constant
CO
2
concentrations, 0.875 bar.
0
0.2
0.4
0.6
0.8
1
0 300 600 900 1200 1500 1800
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 50% CO
950 C 933 C 917 C 900 C
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200
C
o
n
v
e
r
s
i
o
n
,

X


(
-
)
Time, t (min)
Char C, 75% CO
950 C 933 C 917 oC 900 C
0
0.2
0.4
0.6
0.8
1
0 400 800 1200 1600 2000 2400 2800
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 25% CO
950 C 933 C 917 C 900 C
0
0.2
0.4
0.6
0.8
1
0 50 100 150 200 250
C
o
n
v
e
r
s
i
o
n

(
-
)
Time, t (min)
Char D2, 100% CO
950 C 933 C 917 C 900 C
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

143

5.4.2 Effect of CO
2
Concentration in the Reaction Gas
Carbon dioxide gas composition was varied from 25 mole% to 100 mole%. The effect
of changing concentrations of CO
2
in the reaction gas on the reactivity of the chars is
given in Figure 5.4. It is evident from the results that increase in the concentration of
CO
2
in the reaction gas mixture, culminates in a corresponding higher reactivity of the
chars.


(i) Effect of CO
2
concentration on the reactivity of (ii) Effect of CO
2
concentration on the reactivity of
char B at 933 C, 0.875 bar char C at 900 C, 0.875 bar

(i) Effect of CO
2
concentration on the reactivity of (i) Effect of CO
2
concentration on the reactivity of
char C2 at 950 C, 0.875 bar char D2 at 917 C, 0.875 bar

Figure 5.4: Effect of CO
2
concentration on the char reactivity at various constant
isothermal temperatures, 0.875 bar.
0
0.2
0.4
0.6
0.8
1
0 300 600 900 1200 1500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 933 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 400 800 1200 1600 2000 2400 2800
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 900 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200 1400
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 950 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 100 200 300 400 500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 917 C
100% CO 75% CO 50% CO 25% CO
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

144

This can be explained by the fact that an increased CO
2
concentration means that
more CO
2
molecules are available for the gasification reaction. The presence of larger
amounts of CO
2
molecules in the reaction gas stream imply that more active sites in
the char take part in the reaction resulting in a faster reaction rate. All the investigated
chars exhibited similar trends, even at distinct isothermal reaction temperatures and
correlate well with the results of other investigators (Harris and Smith, 1990; Zhang
and Calo, 1996; Fu and Wang, 2001; Murillo et al., 2006; Zhang et al., 2006; Kaitano,
2007, Everson et al., 2006 and 2008a; Hattingh, 2009).
It was also observed that the reactivity at 25% CO
2
concentration was substantially
lower than at the other concentrations. This may be attributed to the substantially
lower amounts of CO
2
molecules in the reaction gas mixture, making it relatively
difficult for the participation of all the char active site in the reaction. Kaitano (2007)
also made a similar finding in char gasification experiments with CO
2
at lower
concentrations of 20 mole% CO
2
. More results are included in Appendix B.


5.5 Determination of the CO
2
Reactivity of the Chars
Reactivity of the chars in this study was confined to the initial reactivity, R, defined as
(Jngten, 1981; Czechowski and Kidawa, 1991; Senneca et al., 1998; Kajitani et al.,
2006; Zhang et al., 2006; Kaitano, 2007; Everson et al., 2008a; Wu et al., 2008;
Hattingh, 2009; Zhang et al., 2010):

0 =
=
t
dt
dX
R
(5.2)

This was achieved by constructing a quantitative plot of the rate of reaction, dt dX
versus the carbon fractional conversion, X. It should be noted that the initial reactivity
is also equivalent to the slope of the carbon conversion versus time plot at time, t = 0.
The characteristic plots of dt dX against X that was used to evaluate the initial
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

145

reactivity of the chars at 25% CO
2
concentration, 0.875 bar is given in Figure 5.5(i) -
(iv); while similar plots at 50, 75 and 100% CO
2
concentrations in the reaction gas
mixture, are included in Appendix B. The summary of determined initial reactivity of
all char samples is presented in Table 5.1.



(i) Rate of reaction versus conversion for char B at (ii) Rate of reaction versus conversion for char C at
25% CO
2
, 0.875 bar. 25% CO
2
, 0.875 bar.


(iii) Rate of reaction versus conversion for char C2 at (iv) Rate of reaction versus conversion for char D2 at
25% CO
2
, 0.875 bar. 25% CO
2
, 0.875 bar.

Figure 5.5: Rate of reaction versus fractional conversion for the chars at 25% CO
2

concentration, 0.875 bar.

0
0.001
0.002
0.003
0.004
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char B
950 C, 25% CO
933 C, 25% CO
917 C, 25% CO
900 C, 25% CO
0
0.0006
0.0012
0.0018
0.0024
0.003
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C
950 C, 25% CO
933 C, 25% CO
917 C, 25% CO
900 C, 25% CO
0.0000
0.0005
0.0010
0.0015
0.0020
0.0025
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C2
950 C, 25% CO
933 C, 25% CO
917 C, 25% CO
900 C, 25% CO
0
0.004
0.008
0.012
0.016
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char D2
950 C, 25% CO
933 C, 25% CO
917 C, 25% CO
900 C, 25% CO
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

146

It is obvious from Figure 5.5 and Table 5.1 that char D2 had the highest initial
reactivity, while char C2 exhibited the lowest reactivity among the four chars. It was
also observed that for chars B, C, and D2, the maximum rate of reaction occurred at
time, t = X = 0; while for char C2, the maximum rate of reaction occurred at time, t >
0 and X > 0. The reactivity of char D2 was found to be greater than the reactivity of
the other chars by a factor > 4. Thus, the increasing order of the reactivity of the chars
is: char C2 < char C < char B < char D2.
The initial reactivity of the chars were correlated with the various characteristic parent
coals and subsequent chars properties to study their contribution and or influence to
the char-CO
2
gasification reaction and kinetics.

Table 5.1: Determined initial gasification reactivity, R of the char at various operating
conditions, 0.875 bar.
Temp. (C)
2
CO
y (mol. %)
Initial Reactivity, R 10
-3
(min
-1
)
Char ID Char B Char C Char C2 Char D2
900 C
25%
1.56 1.02 0.92 6.87
917 C 2.16 1.55 1.25 7.96
933 C 2.51 2.21 1.67 9.57
950 C 3.49 2.69 2.02 12.8
900 C
50%
1.95 1.73 1.29 9.37
917 C 2.86 2.01 1.78 12.0
933 C 3.54 2.93 2.52 13.8
950 C 5.53 4.43 3.37 19.0
900 C
75%
2.24 2.18 1.72 12.3
917 C 3.07 3.19 2.52 16.4
933 C 4.92 4.15 3.06 20.5
950 C 6.72 5.80 4.41 23.3
900 C
100%
3.07 2.37 1.84 15.3
917 C 3.93 3.96 3.09 19.7
933 C 5.92 4.80 3.83 25.6
950 C 8.07 6.50 5.52 27.2
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

147

5.6 Effect of Coal and Char Properties on CO
2
Reactivity of the Chars
One of the objectives of this investigation is to correlate the chemical and physical
properties of both coals and chars with the subsequent char reactivity towards CO
2

gasification. The influence of the parent coals petrographic properties such as total
vitrinite content, total inertinite content, inertinite-vitrinite ratio, coal- TRC and TIC,
MI and RMI*, and rank parameter (vitrinite reflectance (Rr %)), on the reactivity of
the subsequent chars were assessed.
The effects of char properties such as: petrographic properties (char- TRC and TIC);
chemical structural or carbon crystallite properties (aromaticity, fraction of
amorphous carbon, and degree of disorder index) were investigated. The catalytic
influence of the high ash components of the chars were examined using the lumped
parameter, the alkali index (AI). Physical structural properties such as: D-R micropore
surface area, H-K average micropore diameter, and porosity, were also correlated with
the reactivity results.



5.6.1 Effect of Parent Coals Petrographic Properties
It is generally accepted that coal vitrinites are more reactive and possesses better
burnout characteristics than the coal inertinite macerals (Tsai and Scaroni, 1987; du
Cann, 2008). It has also been proven that some inertinite maceral components are
indeed reactive (Cloke and Lester, 1994). These were assessed by qualitative plots
of the initial reactivity of the chars against the petrographic properties of the parent
coals, including the inertinite-vitrinite ratios and coal TRC as shown in Figure 5.6(i)-
(iv). Excluding char D2 that is entirely out of trend with respect to the petrographic
properties of the parent coals, it can be deduced from Figures 5.6, (although the trend
is rather insignificant); that the char reactivity for chars B, C, and C2, increases
slightly with increasing volumes of coal vitrinites; and decreases with increasing
proportions of inertinites in the parent coals.

Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

148


(i) Initial reactivity of the chars versus vitrinite (ii) Initial reactivity of the chars versus inertinite
content of the parent coals content of the parent coals


(iii) Initial reactivity of the chars versus inertinite- (iv) Initial reactivity of the chars versus the parent
vitrinite ratios of the parent coals coals TRC

Figure 5.6: Relationship between the initial reactivity of the chars and the petrographic
properties of the parent coals at 100% and 75% CO
2
, 0.875 bar.

The reactivity of coal vitrinites is enhanced by their propensity to swell, expand and
form vesicles that improve the porosity and increases the surface area of the reacting
chars (Falcon and Snyman, 1986; Cloke and Lester, 1994; Tsai and Scaroni, 1987; du
Cann, 2008). Although some inertinite components are reactive, most of the inertinite
species observed in the four parent coals are the much denser components that do not
easily open up, soften or expand during devolatilisation. This results in denser chars
with little or no significant augmentation of the char porosity or surface area (with the
0.000
0.005
0.010
0.015
0.020
0.025
0.030
0 5 10 15 20 25 30
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y

(
m
i
n
-
1
)
Vitrinite content of parent coals (vol.
%, mmb)
900 C, 100% CO 917 C, 100% CO
933 C, 100% CO 950 C, 100% CO
900 C, 75% CO 917 C, 75% CO
B
C
C2
D2
0.000
0.005
0.010
0.015
0.020
0.025
0.030
50 60 70 80
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y

(
m
i
n
-
1
)

Inertinite content of parent coals (vol. %, mmb)
950 C, 100% CO 933 C, 100% CO 917 C, 100% CO
900 C, 100% CO 950 C, 75% CO 933 C, 75% CO
917 C, 75% CO 900 C, 75% CO
0.000
0.005
0.010
0.015
0.020
0.025
0.030
0 10 20 30
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y

(
m
i
n
-
1
)
Inertinite - vitrinite ratio of parent coals (-)
950 C, 100% CO 933 C, 100% CO 917 C, 100% CO
900 C, 100% CO 950 C, 75% CO 933 C, 75% CO
917 C, 75% CO 900 C, 75% CO
C2
D2
0.000
0.005
0.010
0.015
0.020
0.025
0.030
30 33 36 39 42
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y

(
m
i
n
-
1
)
Parent coals TRC (vol. %, mmb)
950 C, 100% CO 933 C, 100% CO
917 C, 100% CO 900 C, 100% CO
950 C, 75% CO 933 C, 75% CO
D2
C2
C
B
B
C
C2
D2
B
C
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

149

exception of char D2). This is responsible for the observed lower reactivity of the
chars with increasing inertinite volumes in the parent coals (B, C and C2). It was also
observed that the initial reactivity of the chars decrease with increasing inertinite-
vitrinite ratios of the parent coals (though quite insignificant), with the exception of
char D2 which again, was conspicuously off the trend. This is as expected, since
vitrinites are far more reactive than inertinites (Jones et al., 1985; Cai et al., 1998;
Megaritis et al., 1999). Char C2 with the highest original coal inertinite-vitrinite ratio
generally exhibited the lowest reactivity, while char B whose parent coal has the
lowest inertinite-vitrinite ratio was more reactive in all the experiments.
The total reactive components (TRC) of the parent coals earlier discussed in Section
3.7.6, was also assessed for its influence on the initial reactivity of the chars. It was
observed from Figures 5.5(iv) that; for chars B, C and C2, initial reactivity increases
with increasing volume of coal TRC; while the exhibited trend is rather infinitesimal
to be considered significant. The TRC of the precursor coals comprises of reactive
macerals: liptinites, vitrinites, reactive semifusinite and reactive inertodetrinite and its
relative abundance in coal should be a valuable indicator of the expected reactivity of
the chars. It was however observed that char D2 was again out of trend in this regard.
The plot of the initial reactivity of the chars against both the parent coals- and chars-
TIC yielded a considerable scatter without any observed trend and was therefore not
reported.


5.6.1.1 Effect of Maceral Index and Modified Reactive Maceral
Index of the Parent Coals
Due to the inconsequential and unsystematic trends observed in the correlations of the
parent coals petrographic properties to the subsequent chars reactivity, a further step
was taken to confirm the observed negligible trends. It was thus considered that the
use of a lumped predictive parameter that encompasses the combined effects of the
maceral components should be useful. The maceral index (MI), proposed by Su et al.
(2001) and the modified reactive maceral index (RMI*) (Phiri, 2010), modified from
the original reactive maceral index (RMI) proposed by Helle et al. (2003); was
evaluated and assessed. The maceral index (MI) was evaluated from the expression
(Su et al., 2001):
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

150


( )
5 . 2
25 . 1
2
30
|

\
| +
=
GCV
TI
Rr TV TL
MI (5.3)

Helle et al. (2003) modified the MI to RMI by considering the total inertinite
denominator (TI) of Equation 5.3 to consist only of the inert components of the
inertinites adjusted by the subtraction of reactive- semifusinite and inertodetrinite from
the TI term. Thus, the TI was modified to the coal TIC (Table 3.22, Section 3.7.6),
consisting of only inert inertinite maceral components: inert semifusinite,
fusinite/secritinite, micrinite and inert inertodetrinite. The GCV term was also removed
and the RMI takes the form (Helle et al., 2003):

( )
25 . 1
2
coal
TIC
Rr TV TL
RMI
+
= (5.4)

Phiri (2010) went further by modifying Equation 5.4 to reflect the hitherto reactivity of
the reactive semifusinite and reactive inertodetrinite not included in Equation 5.4 by
Helle et al. 2003. By introducing TIC
coal
, from Equation 3.16, Section 3.7.6, the
proposed modified reactive maceral index (RMI*) can be given as:


( )
25 . 1
2
) (
*
coal
TIC
Rr RINT RSF TV TL
RMI
+ + +
= (5.5)

Substituting Equation 3.12 (Section 3.7.6) into Equation 5.5, it can then, be re-written
in terms of the coal- TRC and TIC as follows:

( )
25 . 1
2
) (
*
coal
coal
TIC
Rr TL TRC TL
RMI
+
= (5.6)

The MI and RMI* were evaluated using Equation 5.3 and 5.6 respectively. All
maceral components including coal- TRC and TIC were normalised to mineral matter
free basis (mmfb) as presented in Table 5.2.
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

151



Table 5.2: Results used to evaluate the maceral index (MI) and the modified reactive
maceral index (RMI*) of the parent coals on mineral matter free basis (mmfb).
Sample ID Coal B Coal C Coal C2 Coal D2
Total Vitrinite (vol. %, mmfb) 32.94 24.05 3.529 16.67
Total Liptinite (vol. %, mmfb) 3.529 3.797 3.529 5.128
Total Inertinite (vol. %, mmfb) 63.53 72.15 92.94 78.21
TRC
coal
(vol. %, mmfb) 51.76 48.10 40.00 44.87
TIC
coal
(vol. %, mmfb) 48.24 51.90 60.00 55.13
GCV (MJkg
-1
) 21.40 20.00 20.20 18.10
Vitrinite Reflectance, Rr (%) 0.670 0.720 0.750 0.560
MI (-) 0.184 0.087 0.013 0.071
RMI* (-) 0.873 0.641 0.409 0.878



The determined MI of the coals varied significantly from the reported results of Su et
al. (2001), but correlated well with the findings of Hattingh (2009). This can be
attributed to the fact that Su et al. (2001), investigated vitrinite-rich coals, while
inertinite-rich Highveld coals were used both in this study and that of Hattingh
(2009). The RMI* results were found to be similar to the published results of Phiri
(2010), in his investigation of inertinite-rich South African low grade coals.
The lumped parameters, MI and RMI*, of the parent coals were correlated with the
initial reactivity of the subsequent chars as given in Figure 5.7. Analogous to the
petrographic properties of the coals; a trend of infinitesimally increasing reactivity
with increase in both indices were observed, which look rather insignificant to be
considered as systematic or meaningful. Char D2 was again out of trend, while chars
B, C, and C2 were more consistent.
These findings indicate that, although petrographic properties of parent coals do affect
the quality and quantity of chars produced during pyrolysis, its influence on the char-
CO
2
gasification reaction is quite limited.

Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

152


(i) Initial reactivity of chars versus MI of the (ii) Initial reactivity of the chars versus RMI* of
parent coals the parent coals

Figure 5.7: Relationship between initial reactivity of the chars and the MI and RMI* of the
parent coals at 100% and 75% CO
2
, 0.875 bar.


5.6.1.2 Effect of Rank Parameter of the Parent Coals
Vitrinite reflectance of coal is a valuable tool in the determination of age or extent of
maturation of coal in the coalification process, usually referred to as rank.
The relationship between the mean vitrinite random reflectance (Rr %) of the original
coals and the reactivity of the chars is presented in Figure 5.8. A more meaningful and
very systematic trend was observed. Char reactivity was found to decrease with
increasing vitrinite reflectance vis--vis rank of the parent coal. Although coal B, C,
and C2 were of the same rank (bituminous medium rank C), the slight variation in
vitrinite reflectance was pronounced in the reactivity of the chars. Low reflecting
inertinites have been found to be more reactive than higher reflecting inertinites
(Cloke and Lester, 1994). Char D2 was derived from a lower ranked coal (bituminous
medium rank D), had the lowest vitrinite reflectance (Rr %) value of 0.56 %, and was
generally the most reactive among the four chars.

0.000
0.005
0.010
0.015
0.020
0.025
0.030
0 0.05 0.1 0.15 0.2
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
,

R
(
m
i
n
-
1
)
Maceral index, MI (-)
900 C, 75% CO 917 C, 75% CO 933 C, 75% CO
950 C, 75% CO 900 C, 100% CO 917 C, 100% CO
933 C, 100% CO 950 C, 100% CO
B
C
C2
0.000
0.005
0.010
0.015
0.020
0.025
0.030
0.35 0.45 0.55 0.65 0.75 0.85
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
,

R
(
m
i
n
-
1
)
Modified reactive maceral index, RMI* (-)
900 C, 75% CO 917 C, 75% CO
933 C, 75% CO 950 C, 75% CO
900 C, 100% CO 917 C, 100% CO
D2
C
C
D2
B
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

153


Figure 5.8: Relationship between the initial reactivity of the chars and the vitrinite
reflectance (Rr %) of the parent coals at 100% and 75% CO
2
, 0.875 bar.

However, the higher reactivity of lower ranked coals had long been known to be
positively influenced by the inherent catalytic effects of ash components (Walker and
Hippo, 1975; Laurendeau, 1978; Miura et al., 1989; Huang et al., 1991; Czechowski
and Kidawa, 1991; Kyotani et al., 1993, Ye et al., 1998; Samaras et al., 1996;
Httinger and Nattermann, 1994; Tomita, 2001; Kabe et al., 2004; Lee, 2007). This
holds true for char D2 as it has the highest amount of ash (36.8 wt. %, db) compared
to the other three chars.


5.6.2 Influence of Char Properties on Char-CO
2
Reactivity
5.6.2.1 Influence of Char Petrography (Char- TRC and TIC)
The development of char carbon forms: with frequent tiny gas pores from coal
vitrinites (Category A1 (i)), with fine walled more open network from coal reactives
(Category A2) and thick walled isotropic coke from coal vitrinites (Category B7) in
the transition to char process, improved some properties of the char (du Cann, 2007;
Kaitano, 2007; Everson et al., 2008b). Notable of these enhanced properties are the
0.000
0.004
0.008
0.012
0.016
0.020
0.54 0.59 0.64 0.69 0.74
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
,

R
(
m
i
n
-
1
)
Vitrinite Reflectance, Rr (%)
900 C, 25% CO 917 C, 25% CO 933 C, 25% CO 950 C, 25% CO
900 C, 50% CO 917 C, 50% CO 933 C, 50% CO 950 C, 50% CO
D2
C
C2
B
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

154

surface area and porosity which play a vital role in the gasification reaction. It is
therefore expected that an increase in the volume of total reactive components will
correspondingly give rise to increased char reactivity. This was however, not fully
validated by the insignificant and unsystematic trend observed in the qualitative plot
of the char reactivity against the char TRC (char carbon forms) presented in Figure
5.9. From the result, a relatively infinitesimal increase in char reactivity was seen with
increasing volumes of char TRC for chars B, C, and C2; while char D2 did fit into the
trend.




Figure 5.9: Relationship between the initial reactivity of the chars with the char TRC at
100% and 75% CO
2
, 0.875 bar.

Like the parent coals- TIC, the qualitative correlation of the chars- TIC with the initial
reactivity of the chars, resulted in a disperse scatter without any observed relationship
and was therefore not reported. This further strengthens the fact that, petrographic
properties do not effectively predict the char-CO
2
gasification behaviour of the chars
in this investigation

0.000
0.005
0.010
0.015
0.020
0.025
0.030
5 10 15 20 25 30
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
,

R
(
m
i
n
-
1
)
Char- total reactive components (TRC) (vol. %, mmb)
950 C, 100% CO 933 C, 100% CO 917 C, 100% CO 900 C, 100% CO
950 C, 75% CO 933 C, 75% CO 917 C, 75% CO 900 C, 75% CO
D2
C
C2
B
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

155

5.6.2.2 Influence of Char Carbon Crystallite (Chemical Structural)
Properties
5.6.2.2.1 Influence of Char Aromaticity
Aromaticity of carbonaceous materials refers to the ratio of aromatic carbons to the
total aromatic and aliphatic carbons contained in them (Schoening, 1983; Lu et al.,
2001, 2002a and 2002b; Maity and Mukherjee, 2006; Van Niekerk, 2008). It is the
fraction of aromatic carbon atoms within the char crystalline structure (Lu et al.,
2002b) and can describe the level of structural ordering of the carbons.
A plot of aromaticity of the chars versus the initial reactivity at 50% and 25% CO
2

concentrations is shown on Figure 5.10. It can be deduced from the result, that initial
reactivity decreases with increasing aromaticity of the chars. Char D2 with the lowest
aromaticity of 0.75 exhibited the highest initial reactivity, while char C2 with the
highest aromaticity of 0.95 showed the lowest reactivity. Lu et al. (2002a and 2002b)
reported increasing aromaticity with increasing fractional burnout which is evident
that reactivity is reduced with increasing aromaticity of the chars as expected.



Figure 5.10: Relationship between the initial reactivity of the chars and the aromaticity of
the char samples at 50% and 25% CO
2
, 0.875 bar.
0.000
0.004
0.008
0.012
0.016
0.020
0.84 0.86 0.88 0.9 0.92 0.94 0.96
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
,

R
(
m
i
n
-
1
)
Aromaticity, f
a
(-)
950 C, 50% CO 933 C, 50% CO 917 C, 50% CO 900 C, 50% CO
950 C, 25% CO 933 C, 25% CO 917 C, 25% CO 900 C, 25% CO
C
C2
D2
B
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

156

5.6.2.2.2 Influence of Fraction of Amorphous Carbon in Char

The influence of the amorphous carbon content of the chars on their initial CO
2

reactivity is illustrated in Figure 5.11. It is obvious from the plot, that the initial
reactivity of the char increases almost linearly with increasing content of amorphous
carbon. Amorphous carbon is the fraction of disordered, dis-orientated, non aromatic
carbon contained in these chars (Franklin, 1950; Hirsch, 1954; Ergun, 1968; Lu et al.,
2001, 2002a and 2002b; Kawakami et al., 2006; Wu et al., 2008).
This can be explained by the fact that increasing amorphous fraction of carbon
indicates that the chars are less structurally ordered and may still contain some
aliphatic and or pseudo-aliphatic carbons which are definitely more reactive than the
aromatic carbons. It is also directly related to the aromaticity of the samples, as
samples with higher aromaticity exhibited lower fractions of amorphous carbon as
shown earlier in Figure 3.8(i) in Section 3.7.4.2. Lu et al. (2002a) and more recently,
Wu et al. (2008) reported similar findings.



Figure 5.11: Relationship between the initial reactivity of the chars and the fraction of
amorphous carbon in chars at 50% and 25% CO
2
, 0.875 bar.


0.000
0.004
0.008
0.012
0.016
0.020
0.2 0.25 0.3 0.35 0.4 0.45 0.5
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
,

R
(
m
i
n
-
1
)
Fraction of amorphous carbon, X
A
(-)
950 C, 50% CO 933 C, 50% CO 917 C, 50% CO 900 C, 50% CO
950 C, 25% CO 933 C, 25% CO 917 C, 25% CO 900 C, 25% CO
B
C
C2
D2
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

157

5.6.2.2.3 Influence of Degree of Disorder Index of chars

The fraction of amorphous carbon and aromaticity has been identified as having major
influences on char reactivity. To correlate their lumped contribution to reactivity, Lu
and co-workers (2002a) introduced the degree of disorder index (DOI) to quantify the
disordered carbon in char. They found that reactivity increases with increasing DOI.
This observed trend is not unexpected as the DOI quantitatively refers to the carbon in
amorphous phase and at the aliphatic side chains. An increasing DOI should therefore
suggest an increasing fraction of amorphous carbon and decreasing aromaticity which
will both result to increased char reactivity. The impact of this index on the initial
CO
2
reactivity of the four chars at 50% and 25% CO
2
in reaction gas mixture is
presented in Figure 5.12. It is evident from the plot that reactivity increases linearly
with increasing DOI. Char C2 with the lowest DOI is slowest, while Char D2 with the
highest DOI has the highest reactivity.



Figure 5.12: Relationship between the initial reactivity of chars and the degree of disorder
index, DOI, at 50% and 25% CO
2
, 0.875 bar.

0.000
0.004
0.008
0.012
0.016
0.020
0.25 0.3 0.35 0.4 0.45 0.5 0.55
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
,

R
(
m
i
n
-
1
)
Degree of disorder index, DOI (-)
950 C, 50% CO 933 C, 50% CO 917 C, 50% CO 900 C, 50% CO
950 C, 25% CO 933 C, 25% CO 917 C, 25% CO 900 C, 25% CO
C
C2
D2
B
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

158

5.6.2.3 Inherent Catalytic Effects of Ash Components of Chars
Char D2 has been characterised with some out-of-trend behaviour in most of the
correlations with the petrographic properties of both coals and chars. To have a better
understanding of this, the catalytic activity of the ash components was studied. A
lumped parameter, the alkali index (Sakawa et al., 1982; Zhang et al., 2006) was
evaluated from the XRF results. Alkali index, AI, is a parameter frequently used to
describe the overall influence of catalytically active components in the ash of the
reacting chars. The alkali indices of the chars were plotted against the initial reactivity
at 50% and 25% CO
2
gas compositions. The result is given in Figure 5.13.


Figure 5.13: Influence of the alkali index on the initial reactivity of the chars at 50% and
25% CO
2
, 0.875 bar.

It can be observed from these results that reactivity increases with increasing alkali
index of the chars. Thus, the higher catalytic effect impacted on char D2 by its higher
ash content and a corresponding higher alkali index and the parent coals lower rank,
may well be responsible for its higher initial reactivity even with its parent coals low
vitrinite content and higher inertinite-vitrinite ratio. These may be attributed to the
combined catalytic effects of Ca
2+
, Mg
2+
, Na
+
, Fe
3+
, and K
+
species in the char.
Walker and Hippo (1975); Sakawa et al. (1982); Miura et al. (1989); Tomita (2001);
Sun et al. (2004); Zhang et al. (2006); Lee (2007); Zhang et al. (2010) reported
similar findings.
0.000
0.004
0.008
0.012
0.016
0.020
2 3 4 5 6 7 8
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
,

R
(
m
i
n
-
1
)
Alkali index, AI (-)
900 C, 50% CO 917 C, 50% CO 933 C, 50% CO 950 C, 50% CO
900 C, 25% CO 917 C, 25% CO 933 C, 25% CO 950 C, 25% CO
D2
C2
B
C
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

159

5.6.2.4 Effect of Physical Structural Properties of Chars
5.6.2.4.1 Effect of Micropore Surface Area of Chars

The D-R micropore surface area was correlated with the initial CO
2
reactivity of the
chars. It was observed that the initial reactivity of the chars increases almost linearly
with increasing D-R micropore surface area as shown on Figure 5.14.


Figure 5.14: Influence of the D-R micropore surface area of chars on their initial reactivity
at 100% and 75% CO
2
, 0.875 bar.

This corroborates the result of other investigators that higher surface area promotes
char reactivity (Ng et al., 1988; Liu et al., 2000; Kajitani et al., 2006; akal et al.,
2007). The higher reactivity of Char D2 relative to the other chars was thus promoted,
at least partially, by its higher micropore surface areas, although some contributions
may be expected from its mesopore surface area as the latter acts as feeder pores
(transitional pores) to the micropores (Ng et al., 1988; Liu et al., 2000).
However, due to the large differences usually observed between the micropore surface
area and total pore area (meso- and macropore surface area), Dutta et al. (1977) as well
as Kajitani et al. (2006) had proposed that the total reactive surface area can
effectively be estimated by the micropore surface area.
0.000
0.005
0.010
0.015
0.020
0.025
0.030
110 160 210 260 310
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
,

R
(
m
i
n
-
1
)
Micropore surface area, S (m
2
g
-1
)
950 C, 100% CO 933 C, 100% CO 917 C, 100% CO 900 C, 100% CO
950 C, 75% CO 933 C, 75% CO 917 C, 75% CO 900 C, 75% CO
C
C2
D2
B
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

160

5.6.2.4.2 Influence of Average Micropore Diameter of Chars

The micropore surfaces of chars had been reported as preferred reaction surface area
for gasification reactions (Ng et al., 1988; akal et al., 2007; Liu et al., 2000; Kajitani
et al., 2006). Hence, the average micropore diameter should be able to give a
qualitative description of the microporosity of the chars. Interestingly, char D2 with the
lowest average micropore diameter (3.613 ) has the highest initial reactivity. Char C2
with a higher average micropore diameter exhibited lower reactivity under the same
reaction conditions. The result of this correlation is given in Figure 5.15.


Figure 5.15: Influence of the average micropore diameter of chars on the initial reactivity
at 100% and 75% CO
2
, 0.875 bar.


5.6.2.4.3 Influence of Char Porosity
Char porosity is one of the char properties that undergo significant improvement during
the charring process with respect to the parent coals. The removal of volatiles from coal
and the formation of vesicles, vacuoles and pores in the char enhance this property
(Tsai and Scaroni, 1987; Senneca et al., 1998; du Cann, 2008).
0.000
0.005
0.010
0.015
0.020
0.025
0.030
3.6 3.65 3.7 3.75 3.8 3.85
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y
.

R
(
m
i
n
-
1
)
Average micropore diameter ()
950 C, 100% CO 933 C, 100% CO 917 C, 100% CO 900 C, 100% CO
950 C, 75% CO 933 C, 75% CO 917 C, 75% CO 900 C, 75% CO
D2
B
C
C2
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

161

The relationship between the initial char porosity and reactivity is presented in Figure
5.16. It can be deduced from the plot that the initial reactivity increases with
increasing char porosity. An increase in char porosity has been reported to impact
positively on the gasification reactivity of chars (Laurendeau, 1978; Senneca et al.,
1998) which correlate well with the findings of this investigation.


Figure 5.16: Influence of porosity of char on the initial reactivity at 100% and 75% CO
2
,
0.875 bar.



5.7 Comparison of the CO
2
Reactivity of the four Chars
Reactivity of coal and chars has been used as benchmark for assessing their burn-out
properties and performance during various conversion processes. The four chars used
in this study were thus compared with each other at the same temperature and reaction
gas composition. Only part of the results is presented in Figure 5.17 while the
remainder are included in Appendix B. It is clear from the comparison plots that char
D2 is more reactive than the other three chars: B, C and C2.
Char B generally exhibited a higher initial reactivity with respect to chars C and C2
up to about 90%, and then a very slow reactivity to completion. This has the effect
0.000
0.005
0.010
0.015
0.020
0.025
0.030
4 6 8 10 12 14
I
n
i
t
i
a
l

R
e
a
c
t
i
v
i
t
y

(
m
i
n
-
1
)
Porosity (%)
950 C, 100% CO 933 C, 100% CO 917 C, 100% CO 900 C, 100% CO
950 C, 75% CO 933 C, 75% CO 917 C, 75% CO 900 C, 75% CO
D2
B
C
C2
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

162

that, although Char B has a higher initial reactivity when compared to chars C and C2,
it also has a longer burnout time than for chars C and char C2. This may be attributed
to the structural properties of char B and is an indication that its use in commercial
systems may give rise to the presence of unburnt carbon in the ash or fly ash with
reduced carbon efficiency (Cloke and Lester, 1994; Ouazzane et al., 2002; Helle et
al., 2003; Styszko-Grochowiak et al., 2004)



(i) Comparison CO
2
reactivity of the chars at (i) Comparison CO
2
reactivity of the chars at
900 C, 100% CO
2
, 0.875 bar 917 C, 25% CO
2
, 0.875 bar

(iii) Comparison CO
2
reactivity of the chars at (iv) Comparison CO
2
reactivity of the chars at
933 C, 50% CO
2
, 0.875 bar 950 C, 75% CO
2
, 0.875 bar
Figure 5.17: Comparison of CO
2
reactivity of the chars at various temperatures and CO
2

concentrations, 0.875 bar.
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200 1400
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
900 C, 100% CO
Char D2 Char B Char C Char C2
0
0.2
0.4
0.6
0.8
1
0 400 800 1200 1600 2000
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
917 C, 25% CO
Char D2 Char B Char C Char C2
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
933 C, 50% CO
Char B Char C Char C2 Char D2
0
0.2
0.4
0.6
0.8
1
0 100 200 300 400 500 600 700
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
950 C, 75% CO
Char D2 Char B Char C Char C2
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

163

Generally, reactivity of the chars increases in the order: chars C2 < C < B < D2. The
very low reactivity of char C2 was possibly due to its possession of a very low
fraction of amorphous carbon, low micropore surface area and the high aromaticity
value (Sections 3.7.4 and 3.7.7; Tables 3.15 and 3.23). Char C2 was also found to
contain more structurally ordered carbons than the rest of the chars, with a very low
DOI (Davis et al., 1995; Senecca et al., 1998; Lu et al., 2002a; Kawakami et al.,
2006; Wu et al., 2008). In addition, Char C2 also possessed the highest volumes of
inert char carbon forms amongst the four chars (du Cann, 2008). Furthermore, char
C2 had the lowest alkali index as well as lowest microporosity relative to the other
chars (Sections 3.7.3 and 3.7.7; Tables 3.11 and 3.23).
Char D2 on the other hand exhibited very high micropore surface area, highest
proportion of amorphous carbon, lowest aromaticity value and a relatively higher
alkali index; depicting a higher level of disorderliness which impacts positively on
reactivity (Sections 3.7.3, 3.7.4 and 3.7.7; Tables 3.15 and 3.23). The high AI value of
char D2 should also mean, that the inherent catalytic influence of the ash components
may accelerate its gasification reaction (Laine et al., 1963; Kayembe and Pulsifer,
1976; Ng et al., 1988; Httinger and Natterman, 1994; Miura et al., 1989; Tomita,
2001; Zhang et al., 2006). Moreso, char D2 has the highest microporosity as well as
the lowest average micropore diameter. Microporosity and micropore surface area
have been identified as indicative of the preferred active surface area for CO
2

gasification reactions (Ng et al., 1988; akal et al., 2007; Liu et al., 2000; Kajitani et
al., 2006).
Coal D2 is of lower rank (bituminous medium rank D) with respect to the other three
coals (bituminous medium rank C) (Section 3.7.5; Table 3.16). It is generally
accepted that chars derived from lower ranked coals are more reactive than chars
derived from coals of higher rank (Laurendeau, 1978; Huang et al., 1991; Czechowski
and Kidawa, 1991; Kyotani et al., 1993; Cloke and Lester, 1994; Samaras et al., 1996;
Ye et al., 1998; Kabe et al., 2004; Lee et al., 2007). Thus, the higher reactivity of char
D2 was influenced by the rank parameter.
However, Httinger and Natterman (1994); Miura et al. (1989); Tomita (2001) and
Lee (2007) had attributed the higher reactivity of lower ranked coals partially to the
catalytic activity of their ash components, which also applies to char D2.
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

164


5.8 Summary
With regard to the operational variables used in this study, the char-CO
2
gasification
reactivity was found to follow the Arrhenius type kinetics. Reactivity of the chars were
observed to increase with increasing temperature as well as CO
2
concentration in the
reaction gas mixture.
A comparison between the reactivity of the four chars shows that char reactivity
increases in the order: char C2 < char C < char B < char D2. The reactivity of char D2
was found to be higher than the reactivity of the three other chars by a factor > 4.
The high reactive nature of char D2, although lower in parent coal vitrinite content
and higher in inertinite-vitrinite ratio in comparison to the chars B and C, can
probably be attributed to its higher micropore surface area and microporosity (Ng et
al., 1988; Liu et al., 2000; akal et al., 2007; Zhang et al., 2010), higher alkali index
(Sakawa et al., 1982; Zhang et al., 2006) and higher fraction of amorphous carbon
(Lu et al., 2002a; Wu et al., 2008). Its lower aromaticity also means that there are
more disordered carbons in char D2 with respect to the other chars, hence its higher
reactivity. The lumped parameter of structural disorderliness, the DOI was also found
to have a positive effect on the reactivity of the chars and corroborates the results of
char aromaticity and fraction of amorphous carbon in the chars.
Except for the rank parameter, the vitrinite reflectance, the parent coals petrographic
properties did not exhibit systematic or significant trends when correlated with the
subsequent char reactivity. The only meaningful observation was that, char reactivity
increases with decreasing vitrinite reflectance, which implies that reactivity decreases
with increasing rank. Hence, char C2 which was generally marked with lowest
reactivity had its parent coal possessing a higher vitrinite reflectance (0.75 Rr %), a
very high inertinite content (79 vol. %, mmb) and high inertinite-vitrinite ratio of 26.3.
Char D2 was however, derived from a lower ranked coal (0.56 Rr%).
An overview of the influence of the parent coal and subsequent char properties on the
char-CO
2
gasification reactivity, show that major factors impacting on the reactivity
of the four chars as it pertain to this study can be outlined as follows:
Chapter 5: Gasification with Carbon Dioxide: Results and Discussion

165


Parent Coal properties:
Rank of parent coal (Vitrinite reflectance)
Char properties:
Carbon crystallite (Chemical structural) properties:
Aromaticity
Fraction of amorphous carbon
Degree of disorder index
Catalytic effects of mineral matter (Alkali index)
Physical structural properties:
Micropore surface area
Average micropore diameter
Porosity







166














CHAPTER 6 66 6





6.0 CHAR GASIFICATION WITH CARBON DIOXIDE:
KINETIC MODELLING AND PARAMETERS
EVALUATION

6.1 Introduction
The main objective of kinetic modelling is to use fundamental equations to predict the
gasification reaction process (Murillo et al., 2004; Zhang et al., 2010). The kinetic
modelling as well as the evaluation of various kinetic parameters is presented in this
chapter. A kinetic model- the random pore model (RPM)- was chosen amongst other
models to adequately model the char-CO
2
gasification reaction. The choice was based
Chapter 6: Kinetic Modelling and Parameters Evaluation

167

on fundamental principles and relevance to this type of heterogeneous reaction. The
model description is given in Section 6.2 and 6.3, while the evaluation of the kinetic
parameters, modelling and validation of the experimental results is presented in
Sections 6.5 and 6.6.


6.2 The Random Pore Model
The random pore model and its modifications have been used by several investigators
in the modelling of heterogeneous char-CO
2
gasification reactions (Gavalas, 1980;
Bhatia and Perlmutter, 1980 and 1981; Radovic et al., 1985; Miura et al., 1990;
Ochoa et al., 2001; Murillo et al., 2004 and 2006; Kajitani et al., 2006; Zhang et al.,
2006; Kaitano, 2007; Everson et al., 2008a; Sangtong-Ngam and Narasingha 2008).
The basis assumptions and features of the random pore model are as follows:
The random pore model was developed on the assumption that the reaction
sites occur in cylindrical pores of arbitrary pore size distribution, which may
also overlap, as they grow or collapse during reaction.

The RPM encompasses the intra-particle reaction, which arises due to the
changes imparted to the char (change in porosity, surface area, morphology,
etc) by devolatilisation, thermal annealing and cracking. These cannot be
neglected, as the resultant char physical and chemical properties, impacts on
the overall reaction and overcomes the limitations usually experienced with
the use of the unreacted shrinking core model.

The changes in surface area during reaction are also accounted for in terms of
the initial surface area and other structural properties of the reacting char.
Thus, the variations of the active surface area of the char as gasification
progresses are accounted for by the random pore model.

The model can also be used to identify the different reaction regimes by
considering limiting values for dimensionless groups, characteristic of the
different controlling mechanisms.

Chapter 6: Kinetic Modelling and Parameters Evaluation

168

The random pore model is more flexible than the other models as it is capable
of describing processes that show a maximum rate at certain conversion levels,
as well as processes that do not exhibit this behavior.


6.3 The Random Pore Model Equation
The experimental conditions as given in Table 4.3, Section 4.4 and the very long
experimental burnout times, are such that the char-CO
2
gasification reactions were
considered as under chemical reaction kinetics controlled regime (Regime I).
Furthermore, these conditions are similar to conditions used by various investigators
for Regime I char-CO
2
gasification kinetics (Dutta et al., 1977; Radovic et al., 1985;
Matsui et al., 1987; Hampartsoumian et al., 1993; Fu and Wang, 2001; Ochoa et al.,
2001; Sina et al., 2003; Kajitani et al., 2006; Zhang et al., 2006; Everson et al.,
2006). Thus, random pore model (RPM) evaluations were based on Regime I as
modified by Everson et al. (2008a).
According to Bhatia and Perlmutter (1980), the RPM overall reaction rate can be
described as:

( ) ( )
( )
o
o s
X X S r
dt
dX


=
1
1 ln 1 1
(6.1)

The influence of the operating variables are taken care of by the inclusion of the
intrinsic reaction rate, r
s
, in the overall reaction rate equation, while variations in
structural properties are compensated by the structural parameter, , defined as a
function of the initial char properties:

( )
2
1 4
o
o o
S
L


= (6.2)

By introducing a dimensionless parameter, , defined by Kaitano (2007) and Everson
et al. (2008a) as:

Chapter 6: Kinetic Modelling and Parameters Evaluation

169

( )
o
o s
t S r

=
1
(6.3)

Equation 6.1 can be rewritten in a dimensionless form:

) 1 ln( 1 ) 1 ( X X
d
dX
=

(6.4)

Integration of Equation 6.4 gives the carbon conversion, X, as a function of time, t, or
the dimensionless time, (implicit and explicit).

( ) 1 ) 1 ln( 1
) 1 ( 2

= X
S r
t
o s
o

(6.5)

Substituting the dimensionless parameter, , defined in Equation 6.3, into Equation
6.5 yields the following:

( ) 1 1 ln( 1
2
= X X

(6.6)

Rearranging Equation 6.6 and expressing the carbon conversion, X, as a function of
the dimensionless time, , gives the model equation:

\
|
+ =
4
1 exp 1

X (6.7)

The model equation described in Equation 6.7 can be simplified by incorporating into
it, a time factor, t
f
, defined as (Everson et al., 2008a):

( )
o
o s
f
S r
t

=
1
(6.8)

Expressing Equation 6.7 in terms of the time factor,
f
t , and the real time, t:
Chapter 6: Kinetic Modelling and Parameters Evaluation

170

|
|

\
|
+ =
4
1 exp 1
t t
t t X
f
f
(6.9)

To further simplify the use of the model equations, Kaitano (2007) and Everson et al.
(2008) proposed the use of the reduced time, ) (
x
t t , in place of real time, t , in
Equation 6.5.
) (
x
t t

is the ratio of the reaction time and time for a fractional conversion of X , and
is independent of the parameters before the square root term in Equation 6.5 except
the structural parameter, . This will have the effect that all results for a particular
char gasified should be the same and simplify the determination of the structural
parameter numerically. Equation 6.5 thus takes the form:

1 ) 9 . 0 1 ln( 1
1 ) 1 ln( 1
9 . 0


=

X
t
t
(6.10)

where
9 . 0
t is the time for 90% carbon conversion of the char (carbon). It should
however be noted that:

9 . 0 9 . 0

=
t
t
(6.11)

Since the char-CO
2
gasification reaction follow Arrhenius type kinetics with
dependency on temperature and the power rate law, the intrinsic reaction rate,
s
r ,
equation was therefore based on the n
th
order power rate relationship. This is
expressed as:

m
CO
a
so s
y
RT
E
k r
2
exp |

\
|
= (6.12)
Chapter 6: Kinetic Modelling and Parameters Evaluation

171

6.4 Validation Procedure
The use of the random pore model to determine the kinetic parameters under chemical
reaction controlled regime (Regime 1) requires that the structural parameter , and
the time factor,
f
t , be evaluated. Kaitano (2007) noted that the structural parameter
can be determined either from BET results or from image analysis results as proposed
by Bhatia and Perlmutter (1980), or from experimental data (Liu et al., 2000).
The limitations of applying these methods is that the surface area and pore volume
approximations as using BET and image analyses results of heterogeneous coal
particles with non-uniform pore sizes are not precise. This arises due to the
estimations required in the description of the non-uniform pore sizes especially the
micropores (Kaitano, 2007). The assumptions used in image analysis may introduce
some ambiguities as well. The maximum of the experimental reaction rate curves
obtained from the conversion-time profiles had been used to determine the structural
parameter (Liu et al., 2000). The method of Liu and co-workers (2000), involves the
numerical differentiation of the experimental data, which could be subject to errors as
a result of the chosen differential intervals as noted by Kaitano (2007).
These shortcomings were overcame in this investigation by evaluating the structural
parameter using Equation 6.10 and involves fitting the equation by regression with an
arbitrary structural parameter to corresponding experimental result. The regressed -
value that gives the best fit is accepted as the structural parameter for the experiment
and repeated for all the experiments. The validity of the regressed structural parameter
is further verified by plotting the reduced time against carbon conversion up to 90%
conversion. This plot is supposed to be the same for any of the four chars for all
experiments.
The time factor was determined using Equation 6.13 obtained by rearranging
Equation 6.5:

( ) 1 ) 1 ln( 1
2
= X
t
t
f

(6.13)

Chapter 6: Kinetic Modelling and Parameters Evaluation

172

By plotting the real time, t , against 1 ) 1 ln( 1 X , the time factor can be
calculated from the slope,
t
s , by inserting the regressed structural parameter value
into this simplified relation:

t
f
s
t

2
(6.14)

This was calculated for all the experiments.
The intrinsic reaction rate parameters were evaluated using Equations 6.8 and 6.12.
Rearranging Equation 6.8 gives:

o
o f
s
S
t
r
) 1 (
= (6.15)

Equation 6.12 and Equation 6.15 are equal, since they define the same parameter, the
intrinsic reaction rate
s
r , thus:

m
CO
a
so
o
o f
y
RT
E
k
S
t
2
exp
) 1 (
|

\
|
=

(6.16)

By defining a lumped pre-exponential factor,
'
so
k :

)
'
1 (
o
o so
so
S k
k

= (6.17)

Rearranging Equation 6.16:

m
CO
a
so f
y
RT
E
k t
2
exp
'
|

\
|
= (6.18)

And taking logarithms of both sides gives:
Chapter 6: Kinetic Modelling and Parameters Evaluation

173

) ln( ) ln( ) ln(
'
2
so CO
a
f
k y m
RT
E
t + + = (6.19)

For kinetic parameter evaluations:
Activation energy
a
E , can be calculated using the slope of the plot of ) ln(
f
t
against
1
T at constant CO
2
concentration.
By plotting ) ln(
f
t against ) ln(
2
CO
y at constant temperature, the reaction order
with respect to CO
2
concentration m, can be determined from the slope.
The lumped pre-exponential factor ) ln(
'
so
k , can be evaluated from the
intercepts of ) ln(
f
t against
1
T and ) (
f
t Ln

versus ) ln(
2
CO
y .



6.5 Evaluation of Kinetic Parameters
To effectively quantify the effect of the operational variables of the experiment:
temperature, CO
2
composition in the reaction gas, and the different chars used; kinetic
modelling using the random pore model (RPM) was conducted on all four chars. The
evaluation of the RPM involves three steps (Kaitano, 2007; Everson et al., 2008;):
The determination of the structural parameter , through numerical
regression of the experiment data using Equation 6.10;
The calculation of the time scale parameter (time factor) t
f
, which is one of the
intrinsic reaction rate constants, for each of the experiments as outlined in
Equations 6.13 and 6.14 using the structural parameter , determined using
Equation 6.10.;
The evaluation of the other intrinsic reaction rate constants: activation energy
E
a
, reaction order m, with respect to CO
2
concentration in the reaction gas
mixture, and the lumped pre-exponential factor,
'
so
k .

Chapter 6: Kinetic Modelling and Parameters Evaluation

174

6.5.1 Evaluation of the Structural Factor,
Equation 6.10 was used to determine the structural parameter.

1 ) 9 . 0 1 ln( 1
1 ) 1 ln( 1
9 . 0


=

X
t
t
(6.10)

The dimensionless time
9 . 0
t t was calculated using the experimental real times t, at
different carbon conversions X, and the time to reach 90% conversion,
9 . 0
t . By
choosing an initial arbitrary value for ; the term on the right hand side of Equation
6.10 was regressed numerically. Using the solver add-in function in Microsoft Excel,
the value of was iterated until the error sum of squares (ESS) converged to the
lowest minimum using the equation:

2
9 . 0
1 ) 9 . 0 1 ln( 1
1 ) 1 ln( 1

|
|

\
|

|
|

\
|
=
Cal
Exp
X
t
t
ESS

(6.20)

Note that the subscripts: Exp and Cal refer to reduced time determined from the
experimental data and that calculated using the arbitrary values respectively.
The summary of this evaluation is given in Table 6.1, while the detailed results are
included in Appendix C.

Table 6.1: Dimensionless structural parameters for the char pores.
Char ID Char B Char C Char C2 Char D2
Structural Parameter, (-)
1.11 0.10 1.75 0.18 2.58 0.32 1.57 0.20


Kaitano (2007) had noted that the values of the structural parameter do not change
significantly with variation of operating conditions. This was also observed with the
char samples, as the respective values for each experiment did not show substantial
variations to the average value for each of the chars. This is reflected in their
Chapter 6: Kinetic Modelling and Parameters Evaluation

175

acceptable average deviations which were: 9% for char B; 10% for char C; and 12%
for both char C2 and D2 from their respective average values.
The structural parameter , values for chars B, C and D2 were < 2, which show that
the maximum reaction rate and the maximum reactive surface area occur at the
fractional carbon conversion, 0 = X . Thus pore coalescence was more significant
than pore growth at the initial stage of the reaction. From this point onwards, the
reaction surface area and the reaction rate decreases with increasing conversion
(Bhatia & Perlmutter, 1980).
For char C2 with a structural parameter value of 2.58, ( > 2), maximum reaction rate
and maximum surface area do not occur at fractional carbon conversion 0 = X but at
a latter value of X . This indicates that pore growth rather than pore coalescence is the
initial dominant structural mechanism. Maximum reaction rate will thus occur when
pore coalescence become more significant after the initial pore growth (Bhatia &
Perlmutter, 1980, Liu et al., 2000; Kaitano, 2007). Apart from the values, a good
depiction of the significant structural mechanisms during the gasification reaction is
the rate of reaction versus fractional conversion plots (Figure 5.5) earlier presented in
Section 5.5. It is clear from Figure 5.5, that except for char C2 with maximum rate of
reaction occurring at about X = 0.2, implying that real time, t is > 0; chars B, C and
D2 exhibited maximum rate of reaction at X = 0 and therefore, t = 0. These are
similar to the findings and discussions of Bhatia & Perlmutter (1980); Liu et al.
(2000); Kaitano (2007); Murillo et al. (2004 and 2006) and Everson et al. (2008a).
Other results are given in Figure 6.8 of Section 6.6 and Appendices B and D. It is
important to note that the char-CO
2
gasification reaction is sufficiently within Regime
1 (chemical reaction controlled), according to the RPM, not-withstanding the low
porosity of the chars.
A comparison between the experimental results and the model predictions using a
dimensionless plot of conversion X, against reduced time
9 . 0
t t , for char B is
presented in Figure 6.1. The result confirms that the evaluated structural parameter
value is appropriate to describe the experimental data of char B. The dotted lines in
the figure indicate the error interval. This was also validated for chars C, C2 and D2
with similar results and is included in Appendix C.
Chapter 6: Kinetic Modelling and Parameters Evaluation

176


Figure 6.1: Comparison of the gasification experimental and model results for char B,
0.875 bar.



6.5.2 Determination of the Time Factor, t
f

The time scale parameter, time factor (t
f
), for all the experiments were determined
using Equation 6.13 and 6.14.

( ) 1 ) 1 ln( 1
2

= X
t
t
f

(6.13)

t
f
s
t

=

2
(6.14)

This was achieved by using the slope
t
s , of the plot of the experimental real times
(t) versus the linearised function of X, ( ) 1 ) 1 ln( 1 X and the evaluated
structural parameter characteristic of the experiment. The time factor calculated at
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Reduced time, t/t
0.9
(-)
Char B
Model 917 C, 100% CO 900 C, 100% CO
933 C, 100% CO 950 C, 100% CO 900 C, 75% CO
917 C, 75% CO 933 C, 75% CO 950 C, 75% CO
900 C, 50% CO 917 C, 50% CO 933 C, 50% CO
950 C, 50% CO 900 C, 25% CO 917 C, 25% CO
933 C, 25% CO
Chapter 6: Kinetic Modelling and Parameters Evaluation

177

various operating conditions, compared with the experimentally determined initial
reactivity for char B as well as time taken to reach 50% and 90% conversion (t
0.5
and
t
0.9
), are summarised in Table 6.2. Similar results were obtained for chars C, C2 and
D2 from corresponding evaluations and are presented in Appendix C.


Table 6.2: Summary of the structural parameter, time factors and initial reactivity for
char B, 0.875 bar.
Temp.
(C)
2
CO
y
(mol. %)
t
0.5

(min)
t
0.9

(min)
R 10
-3

(min
-1
)
t
f
10
-3

(min
-1
)

(-)
900 C
25%
319.91 1099.2 1.56 1.51
1.06
917 C 231.10 866.13 2.16 1.97
0.91
933 C 198.93 693.10 2.51 2.38
1.03
950 C 143.47 484.65 3.49 3.35
1.11
900 C
50%
256.60 813.10 1.95 1.92
1.27
917 C 174.66 588.21 2.86 3.06
1.10
933 C 141.22 443.05 3.54 3.48
1.31
950 C 90.467 337.17 5.53 5.12
0.93
900 C
75%
223.57 735.65 2.24 2.13
1.21
917 C 162.97 549.93 3.07 2.87
1.17
933 C 101.53 364.33 4.92 4.57
1.02
950 C 74.417 237.13 6.72 6.59
1.26
900 C
100%
162.85 605.93 3.07 2.77
0.96
917 C 127.25 433.30 3.93 3.71
1.12
933 C 84.467 274.87 5.92 5.83
1.19
950 C 61.983 209.22 8.07 7.96
1.08


From the table, it is clear that the time factor increases with increasing temperature
and CO
2
concentration, with values approximately equal to the experimentally
determined initial reactivity of the chars. Similar results were reported by Kaitano
(2007), for char gasification with CO
2
. The time factor values obtained for char: B, C
Chapter 6: Kinetic Modelling and Parameters Evaluation

178

and C2; as well as initial reactivity determined from the experimental data are close to
each other, which suffixes that their respective reactivity did not vary significantly as
seen in Figure 5.17 of Section 5.7. Char D2 generally exhibited higher initial
reactivity with a similar higher RPM determined time factor, t
f
.
This confirms the correlation of char properties earlier given in Section 5.6. A closer
look at Table 6.2 and comparing it with similar results provided in Appendix C for
chars C, C2 and D2; reveals and confirms further that the reactivity of char D2 is
higher than the reactivity of the other three chars by a factor > 4, evident from their
respective time factor and initial reactivity. This can also be deduced from the char
comparison plots of Figure 5.17, Section 5.7.


6.5.3 Determination of Activation Energy, E
a

The activation energy of the chars was calculated using the linearised form of
Equation 6.18 given in Equation 6.19:

) ln( ) ln( ) ln(
'
2
so CO
a
f
k y m
RT
E
t + + = (6.19)

This was achieved by plotting ) ln(
f
t against the reciprocal of the isothermal
experimental temperature (in Kelvin), at constant CO
2
concentration and the apparent
activation energy
a
E , was evaluated from the slope
a
E
s , according to the relation:

R
E
s
a
E
a

= (6.21)

A detailed result of the determined activation energy at different reaction gas mixtures
for all four samples is presented in Table 6.3; while their respective Arrhenius plots,
with correlation coefficients > 0.95, are shown in Figure 6.2.

Chapter 6: Kinetic Modelling and Parameters Evaluation

179

Table 6.3: Details of results of activation energy for the char samples at different CO
2
concentrations in the reaction gas, 0.875 bar.
Char B Char C2
2
CO
y (mol.%)
E
a
(kJ mol
-1
) R
2
2
CO
y (mol.%)
E
a
(kJ mol
-1
) R
2

25% 185.4 0.987 25% 194.6 0.990
50% 221.5 0.966 50% 236.8 0.998
75% 276.2 0.991 75% 224.8 0.987
100% 259.5 0.990 100% 252.9 0.980
Average E
a
, char B = 235.7 32.2 Average E
a
, char C2 = 227.3 17.6
Char C Char D2
25% 229.9 0.987 25% 152.3 0.974
50% 232.0 0.957 50% 168.1 0.990
75% 229.2 0.996 75% 165.0 0.973
100% 235.8 0.974 100% 167.7 0.995
Average E
a
, char C = 231.7 2.2 Average E
a
, char D2 = 163.3 5.5


From the results it is clear that there are no significant variations in the activation
energy of the chars at the various CO
2
gas mixtures. This is further corroborated by
the small average deviations of 0.95%, 7.7%, and 3.3% for chars C, C2 and D2
apiece. However, char B exhibited a relatively low activation energy value at 25%
CO
2
concentration culminating in a rather higher average deviation of 13.6%.
The average activation energy values obtained for the chars are: 235.7 32.2; 231.7
2.2; 227.3 17.6; and 163.3 5.5 kJmol
-1
for char B, C, C2 and D2 respectively.
Thus, the activation energy for chars B, C and C2 are close to each other; while char
D2 has the lowest E
a
value. This may not be unconnected with the catalytic effect of
the higher ash components of char D2 (Miura et al., 1989 and 1990; Sun et al., 2004;
Zhang et al., 2006). The higher activation energy values for the other chars can be
attributed to the high inertinite maceral content of the precursor coals (Sun et al.,
2004). This leads to formation of higher volumes of dense char forms after pyrolysis
with poor burnout properties and longer residence time in gasification systems and
other conversion processes (Kaitano, 2007; du Cann, 2008; Everson et al., 2008b).
Chapter 6: Kinetic Modelling and Parameters Evaluation

180


(i) Arrhenius plot for char B, 0.875 bar (i) Arrhenius plot for char C, 0.875 bar

(iii) Arrhenius plot for char C2 at 0.875 bar (iv) Arrhenius plot for char D2 at 0.875 bar
Figure 6.2: Arrhenius plots of the char-CO
2
gasification reaction at 100, 75, 50 and 25%
CO
2
concentrations, 0.875 bar.


Activation energy values for CO
2
gasification of inertinite-rich chars of 199 - 266
kJmol
-1
was reported by Kaitano (2007); while Everson et al. (2008a) observed E
a

values of 192 - 247 kJmol
-1
on similar chars (produced from inertinite-rich low grade
coals). Comparison with other results from literature is presented in Table 6.7.
-6.6
-6.2
-5.8
-5.4
-5
8.1 8.2 8.3 8.4 8.5 8.6
L
n

(
t
f
)
T
-1
10
-4
(K
-1
)
Char B
100% CO 75% CO 50% CO 25% CO
-7
-6.6
-6.2
-5.8
-5.4
-5
8.1 8.2 8.3 8.4 8.5 8.6
L
n

(
t
f
)
T
-1
10
-4
(K
-1
)
Char C
100% CO 75% CO 50% CO 25% CO
-7
-6.6
-6.2
-5.8
-5.4
-5
8.1 8.2 8.3 8.4 8.5 8.6
L
n

(
t
f
)
T
-1
10
-4
(K
-1
)
Char C2
100% CO 75% CO 50% CO 25% CO
-5.1
-4.8
-4.5
-4.2
-3.9
-3.6
-3.3
8.1 8.2 8.3 8.4 8.5 8.6
L
n

(
t
f
)
T
-1
10
-4
(K
-1
)
Char D2
100% CO 75% CO 50% CO 25% CO
Chapter 6: Kinetic Modelling and Parameters Evaluation

181

6.5.4 Determination of Order of Reaction, m
The reaction order m, with respect to CO
2
concentration in the reactant gas
composition was determined using the linearised Equation 6.19. It was obtained
directly from the slope of the plot of ) (
f
t Ln versus ) (
2
CO
y Ln at constant temperature
and pressure. This was determined for all the char samples.

) ln( ) ln( ) ln(
'
2
so CO
a
f
k y m
RT
E
t + + = (6.19)

These plots are shown in Figures 6.3 and 6.4; while the evaluated order of reaction for
all four chars including the correlation coefficient is presented in Table 6.4.
The results show that the orders of reaction of the char-CO
2
gasification with respect
to CO
2
concentration are: 0.52 0.11; 0.67 0.03; 0.61 0.07; and 0.63 0.05
respectively for chars B, C, C2 and D2. The correlation coefficients for the plots are >
0.95. The average deviations for the determined order of reaction m, of the chars are:
4.5% for char C; 11.5% for char C2; and 7.9% for chars D2; while char B exhibited a
higher average deviation of 21%.


(i) Determination of gasification reaction (i) Determination of gasification reaction
order for char B order for char C

Figure 6.3: Determination of char-CO
2
gasification reaction order for chars B and C at
various temperature and constant CO
2
concentrations, 0.875 bar.
-6.6
-6.1
-5.6
-5.1
-4.6
3 3.4 3.8 4.2 4.6
L
n

(
t
f
)
Ln (y
CO
)
Char B
950 C 933 C 917 C 900 C
-7.0
-6.5
-6.0
-5.5
-5.0
3 3.4 3.8 4.2 4.6
L
n

(
t
f
)
Ln (y
CO
)
Char C
900 oC 917 oC 933 oC 950 oC
Chapter 6: Kinetic Modelling and Parameters Evaluation

182


(i) Determination of gasification reaction (i) Determination of gasification reaction
order for char C2 order for char D2

Figure 6.4: Determination of char-CO
2
gasification reaction order for chars C2 and D2 at
various temperature and constant CO
2
concentrations, 0.875 bar.



Table 6.4: Details of result for reaction order with respect to CO
2
concentration at
different temperatures and constant CO
2
gas composition at 0.875 bar.
Char B Char C2
Temp. (C)
m ) (
2
m
CO
y
R
2
Temp. (C)
m ) (
2
m
CO
y
R
2

900 C 0.41 0.938 900 C 0.52 0.991
917 C 0.41 0.865 917 C 0.65 0.986
933 C 0.64 0.991 933 C 0.56 0.996
950 C 0.62 0.999 950 C 0.71 0.999
Average m: char B = 0.52 0.11 Average m: char C2 = 0.61 0.07
Char C Char D2
900 C 0.64 0.980 900 C 0.57 0.989
917 C 0.71 0.956 917 C 0.59 0.982
933 C 0.63 0.979 933 C 0.72 0.987
950 C 0.68 0.990 950 C 0.62 0.991
Average m: char C = 0.67 0.03 Average m: char D2 = 0.63 0.05
-7.0
-6.5
-6.0
-5.5
-5.0
3 3.4 3.8 4.2 4.6
L
n

(
t
f
)
Ln (y
CO
)
Char C2
900 oC 917 oC 933 oC 950 oC
-5.1
-4.7
-4.3
-3.9
-3.5
3 3.4 3.8 4.2 4.6
L
n

(
t
f
)
Ln (y
CO
)
Char D2
900 oC 917 oC 933 oC 950 oC
Chapter 6: Kinetic Modelling and Parameters Evaluation

183

The order of char-CO
2
gasification reaction obtained in this study correlates well with
the results of other investigators (Harris and Smith, 1990; Hampartsoumian et al.,
1993; Fu and Wang, 2001; Ochoa et al., 2001; Kajitani et al., 2006). Comparison with
published results from literature is also given in Table 6.7.


6.5.5 Determination of Lumped Pre-exponential Factor,
'
so
k
The lumped pre-exponential factor was determined from the intercepts of Figures 6.2
and 6.3 and 6.4. Details of the evaluation of
'
so
k at different CO
2
concentration in the
reaction gas mixture is given in Table 6.5. From the results, average lumped pre-
exponential factor values of: 5.22 2.0 10
8
; 4.36 1.9 10
7
; 1.12 1 10
8
; and 2.69
1 10
5
min
-1
bar
-1
were obtained for char B, C, C2 and D2 respectively. These range
of results are in good agreement with results reported by: Miura et al., (1990); Fu and
Wang, (2001); Kaitano, (2007); Everson et al., (2008a); Sangtong-Ngam and
Narasingha, (2008).

Table 6.5: Determination of lumped pre-exponential factor for the four char samples,
0.875 bar.
Char B Char C2
2
CO
y
'
so
k (min
-1
bar
-1
)
2
CO
y
'
so
k (min
-1
bar
-1
)
25% 1.07 10
8
25% 9.25 10
6

50% 4.62 10
8
50% 4.77 10
7

75% 5.53 10
8
75% 1.86 10
7

100% 9.67 10
8
100% 3.74 10
8

Average
'
so
k = 5.22 2.0 10
8
Average
'
so
k = 1.12 1.0 10
8

Char C Char D2
25% 1.86 10
7
25% 4.01 10
4

50% 3.53 10
7
50% 2.88 10
5

75% 3.71 10
7
75% 2.85 10
5

100% 8.35 10
7
100% 4.65 10
5

Average
'
so
k = 4.36 1.9 10
7


Average
'
so
k = 2.69 1.0 10
5

Chapter 6: Kinetic Modelling and Parameters Evaluation

184

A summary of the structural and kinetic parameters derived from this study is
presented in Table 6.6; while a comparison with published results from other
investigators is presented in Table 6.7.

Table 6.6: Summary of structural and kinetic parameters for chars B, C, C2 and D2
Kinetic Parameter Char B Char C Char C2 Char D2

(-)

1.11 0.10 1.75 0.18 2.58 0.32 1.57 0.20
m (-) 0.52 0.11 0.67 0.03 0.61 0.07 0.63 0.05
E
a
, (kJ mol
-1
) 235.7 32.2 231.7 2.2 227.3 17.6 163.3 5.5
'
so
k , (min
-1
bar
-1
)
5.22 2.0 10
8
4.36 1.9 10
7
1.12 1.0 10
8
2.69 1 10
5



6.6 Validation of Kinetic Model and Associated Parameters
The validation of the RPM was done by a plot of experimental or actual initial
reactivity, R versus the RPM predicted initial reactivity designated by the time scale
parameter (time factor), t
f
. The relationship between the actual- and predicted t
f
for all
four chars is shown in the parity plot of Figure 6.5.


Figure 6.5: Parity plot of predicted versus actual t
f
values for all four char samples.
0.000
0.005
0.010
0.015
0.020
0.025
0.030
0.000 0.005 0.010 0.015 0.020 0.025 0.030
P
r
e
d
i
c
t
e
d

R
e
a
c
t
i
v
i
t
y
,

t
f
(
m
i
n
-
1
)
Initial Reactivity, R (min
-1
)
Char B
Char C
Char C2
Char D2
Linear trend
Chapter 6: Kinetic Modelling and Parameters Evaluation

185



Table 6.7: Models used, structural parameter, and kinetic parameters obtained for char-CO
2
gasification reaction by other investigators.
Parent coal
type
Reactor
used
Temperature
range (C)
Model
used
E
a

m (-)
so k
'

Reference
(-) (kJ mol
-1
) (min
-1
bar
-1
)

Bituminous TGA 900-950 RPM 1.11-2.58 163-236 0.52-0.67 2.69 10
5
-5.22 10
8
This investigation
Lignite TGA 840-1100 VRM - 248 - - Dutta et al., 1977
Sub bituminous TGA 800-950 SCM - 109-137 - - Everson et al., 2006
Coal discard TGA 850-900 RPM 1.04 192-247 - 8.69 10
7
-1.03 10
8
Everson et al., 2008a
Coal TGA 927-1127 SUCM - 245 0.85 1.18 10
7
Fu and Wang, 2001
UK coals EFR 627-960 - - 239 0.5-0.6 - Hampartsoumian et al., 1993
Brown coal VTR 750-900 - - 214-230 0.57 - Harris and Smith, 1990
Bituminous HPTGA 900-1000 RPM 1.23-1.43 171-202 0.49-0.77 - Hattingh, 2009
Coal discard HPTGA 850-900 RPM 1.04 192-266 0.5 9.60 10
8
Kaitano, 2007
Bituminous PDTF/TGA 750-1400 RPM 1.0-26 240-280 0.43-0.56 2.54 10
7
-1.19 10
9
Kajitani et al., 2006
Sub bituminous TGA 885-985 VRM - 268 - - Matsui et al., 1987
Lignite TGA 650-900 Various 5.0
a
204-212 0.71-1.0 1.45 10
3
-6.17 10
6
Murillo et al., 2006
Sub bituminous TGA 900-1160 RPM 4.7-7.0 156-165 0.50-0.58 - Ochoa et al., 2001
UK coal TGA 827-1102 - - 220 - - Radovic et al., 1985
Lignite HTR 900-980 - - 128-146 - - Sina et al., 2003
Anthracites TGA 920-1050 SCM - 146-202 - 9.2 10
3
-9.8 10
5
Zhang et al., 2006
a
- Determined using RPM
Chapter 6: Kinetic Modelling and Parameters Evaluation

186

From the parity plot, it is obvious that the predicted initial reactivity of the chars by
the RPM time factor, (t
f
) is a good approximation of the experimentally determined
initial reactivity. This is good evidence to uphold the validity of the RPM to describe
the char-CO
2
gasification reaction with respect to this investigation under Regime I.
The validity of the model was also checked by constructing a plot of the
experimentally determined conversions with the RPM predicted conversions. This is
demonstrated in Figure 6.6 for char B at all the experimental conditions at 0.875 bar
pressure. The results obtained for char C, C2 and D2 compare very well with the
results of char B and can be found in Appendix C.



Figure 6.6: Comparison between experimental and model gasification results for
char B.




0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
M
o
d
e
l

X
,

t
i
m
e

f
a
c
t
o
r
,

(
t
f
)

(
m
i
n
-
1
)
Experimental X, intial reactivity, R (min
-1
)
Char B
900 C, 100% CO 917 C, 100% CO 933 C, 100% CO
950 C, 100% CO 900 C, 75% CO 917 C, 75% CO
933 C, 75% CO 950 C, 75% CO 900 C, 50% CO
917 C, 50% CO 933 C, 50% CO 950 C, 50% CO
900 C, 25% CO 917 C, 25% CO 933 C, 25% CO
950 C, 25% CO Linear Trend
Chapter 6: Kinetic Modelling and Parameters Evaluation

187

The fitting of the RPM to the experimental conversion data for all four char samples
were conducted as a further validation of the model. The results of this fitting at
different operating conditions are given in Figure 6.7(i) - (iv). It can be observed from
the result that the RPM adequately described the experimental data (conversion) of all
four chars under Regime I. Only part of the results is presented in Figure 6.7; while
the remainder are included in Appendix D.



(i) RPM fitting of char B conversion at (ii) RPM fitting of char C conversion at 75%
50% CO
2
, 0.875 bar CO
2
, 0.875 bar


(iii) RPM fitting of char C2 conversion (iv) RPM fitting of char C2 conversion at
at 25% CO
2
, 0.875 bar 100% CO
2
, 0.875 bar

Figure 6.7: RPM fitting of char conversion at different CO
2
concentrations.
0
0.2
0.4
0.6
0.8
1
0 300 600 900 1200 1500 1800
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 900 C, 50% CO
Char B, 917 C, 50% CO
Char B, 933 C, 50% CO
Char B, 950 C, 50% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 900 C, 75% CO
Char C, 917 C, 75% CO
Char C, 933 C, 75% CO
Char C, 950 C, 75% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 450 900 1350 1800 2250 2700
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 900 C, 25% CO
Char C2, 917 C, 25% CO
Char C2, 933 C, 25% CO
Char C2, 950 C, 25% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 50 100 150 200 250
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 900 C, 100% CO
Char D2, 917 C, 100% CO
Char D2, 933 C, 100% CO
Char D2, 950 C, 100% CO
RPM
Chapter 6: Kinetic Modelling and Parameters Evaluation

188

It was also possible to use the RPM to predict the rate of reaction. Fitting of the RPM
predicted conversion rate (dX/dt) versus conversion to the experimental reaction rate
change against conversion were constructed and presented in Figure 6.8. Again, a
good fit was obtained. More of these plots are provided in Appendix D.
The good fit to the experimental data exhibited by RPM is a good confirmation of the
validity of the model and also shows that the gasification reaction is under Regime I
(chemical reaction controlled).




Figure 6.8: RPM fitting of char conversion rate for chars B, C, C2 and D2 at 25% CO
2

concentration and different temperatures.

0
0.001
0.002
0.003
0.004
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char B, =1.11
950 C, 25% CO
933 C, 25% CO
917 C, 25% CO
900 C, 25% CO
RPM
0
0.001
0.002
0.003
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C, =1.75
950 C, 25% CO
933 C, 25% CO
917 C, 25% CO
900 C, 25% CO
RPM
0.0000
0.0005
0.0010
0.0015
0.0020
0.0025
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C2, =2.58
950 C, 25% CO
933 C, 25% CO
917 C, 25% CO
900 C, 25% CO
RPM
0
0.004
0.008
0.012
0.016
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char D2: =1.57
950 C, 25% CO
933 C, 25% CO
917 C, 25% CO
900 C, 25% CO
RPM
Chapter 6: Kinetic Modelling and Parameters Evaluation

189

6.7 Summary
Structural parameters, , determined for the respective chars: B, C, C2 and D2 are:
1.11, 1.75, 2.58 and 1.57. This shows that the maximum reaction rate and maximum
reaction surface area occur at: X = 0 for chars B, C, and D2; while C2 exhibited a
maximum at: X > 0 (Bhatia and Perlmutter, 1980; Liu et al., 2000; Kaitano, 2007).
Pore coalescence was therefore the dominant structural mechanism during the initial
stage of gasification for chars B, C and C2. On the other hand, pore growth was more
significant during the initial period of char C2 gasification reaction. This was
corroborated by the conversion plots of the chars and validated by the plot of
experimental conversion against the reduced time (t/t
0.9
) for each experiment, with a
satisfactory correlation.
Time scale parameter t
f
, determined for each experiment were similar in value to the
initial reactivity evaluated from the experimental result. The RPM predicted reactivity
also confirmed the higher reactivity of the chars, with the reactivity of char D2 higher
than that of the other chars by a factor > 4. The time factor was validated by a parity
plot of model predicted initial reactivity (t
f
) versus the experimentally determined
reactivity R; which showed a consistent linear trend for all four chars.
The activation energy obtained for the gasification reactions are: 235.7 kJmol
-1
for
char B; 231.7 kJmol
-1
for char C; 227.3 kJmol
-1
for char C2 and 163.3 kJmol
-1
for
char D. The order of reaction with respect to CO
2
concentration in the reaction gas
composition were found to be between 0.52 to 0.67 for the chars. Lumped pre-
exponential factors stood at: 5.22 10
8
, 4.36 10
7
, 1.12 10
8
and 2.69 10
5
min
-1
bar
-1

for chars B, C, C2, and D2 respectively. The lower activation energy of char D2 was
attributed to its lower rank with respect to the other chars and the significant catalytic
influence of inorganic components of the ash.
The random pore model (RPM) was found to satisfactorily describe the experimental
results of chars B, C, C2 and D2 in the reaction controlled regime (Regime I). The
RPM was validated by a parity plot of the experimental conversion against the model
predicted conversion which followed a linear trend for the four chars. The fitting of
the RPM conversions and conversion rates data to the experimental conversions and
conversion rates also gave adequate fits.
Chapter 6: Kinetic Modelling and Parameters Evaluation

190

The kinetic parameters and the reaction rate constants determined for each char
corresponded well with published results in the open literature.
The intrinsic reaction rate of the chars can thus be written in these empirical
Arrhenius forms:
Char B: ( )
52 . 0 8
2
235700
exp 10 22 . 5
CO s
y
RT
r
|

\
|
= (6.22)
Char C: ( )
67 . 0 7
2
231700
exp 10 36 . 4
CO s
y
RT
r
|

\
|
= (6.23)
Char C2: ( )
61 . 0 8
2
227300
exp 10 12 . 1
CO s
y
RT
r
|

\
|
= (6.24)
Char D2: ( )
63 . 0 5
2
163300
exp 10 69 . 2
CO s
y
RT
r
|

\
|
= (6.25)















191













CHAPTER 7 77 7





7.0 CONCLUSION AND RECOMMENDATIONS

7.1 Introduction
The effects of the characteristic chemical and physical properties of both the parent
coals and the subsequent chars on the char-CO
2
gasification reaction kinetics have
been elucidated in this dissertation. In this chapter, a summary of the major findings
and conclusions reached during the investigation are presented. Prospects and relevant
recommendations for future studies are also provided.
Chapter 7: Conclusion and Recommendations

192

7.2 General Conclusions
The general conclusions derived from this investigation are as follows:
Petrographically, the four parent coals were characterised as inertinite-rich and
have high ash contents. The vitrinite reflectance are within a small range of 0.56 -
0.75% Rr in all the coal samples. These coal samples are thus classified as
bituminous medium rank C to D. Considerable shift to higher values were
observed in the total maceral reflectance scans of the chars, relative to Rsc% of the
parent coals. The pure mono-maceral, inertite was observed to be substantially
most abundant in all four coals. The transition from coal to char led to the
formation of various char carbon forms in the chars. Greater volumes of dense and
thicker walled char networks were formed from coal inertinites, while coal
vitrinites yielded less dense chars with tiny gas pores and minor proportions of
isotropic coke. Coal reactive macerals formed fine walled, more open char
networks with a rather low abundance in all four chars. The total reactive
components (TRC) generally decreased from coals to the chars with losses >36.5%.
The total inert components (TIC) exhibited gains in the transition from coals to
chars, with gains of 12.2 to 24.4%, except char C2 with neither gain nor loss (0%).
These losses and gains may be attributed to the conversion of some of the
reactive macerals in the coals to inert char carbon forms during the pyrolysis
reactions, while the 0% loss observed in char C2 may be due to its content of
higher volume of partially reacted macerals.

Except for the maceral contents and inertinite-vitrinite ratios, significant changes in
chemical, physical and structural properties were not observed in the four original
coals. The subsequent chars prepared at 900 C with a holding time of 70 minutes,
however exhibited differences, both amongst the chars and from the parent coals.
Chemical properties such as: fixed carbon, elemental carbon, and ash contents
increased, while elemental oxygen and hydrogen content decreased from coal to
the chars. Volatile matter and moisture contents were substantially reduced from
the coals to the chars. There were however, some traces of both moisture and
volatiles in the chars. These significant changes resulted in an increase in skeletal
density, aromaticity and a corresponding reduction in the fraction of amorphous
carbon and the degree of disorder index of the chars.
Chapter 7: Conclusion and Recommendations

193

On the physical and structural properties; both the micropore surface area and
microporosity increase, while average micropore diameter decrease from coals to
chars due to the evacuation of volatiles and the opening of hitherto blind or
inaccessible fine pores. The increasing order of magnitude of the micropore
surface area and microporosity of the original coals did however change in the
chars. Carbon crystallite lattice parameters determined for the parent coals and the
subsequent chars revealed that the chars are more structurally ordered, more
compact and condensed (smaller in size) than the original coals evident from the
decreased lattice parameters: L
a
, L
c
and N
ave
and increased d
002
from coals to chars.
Although, these changes have been largely attributed to the heat treatment during
char production, the mechanism of structural ordering is not well understood. The
increasing orders of magnitude of both the fraction of amorphous carbon and
structural disorderliness was found to change from coals (coals: B < C < C2 < D2)
to the chars (chars: C2 < C < B < D2). XRD mineral analysis confirmed the
predominance of clay and quartz mineral in both coal and char samples which was
corroborated by the XRF ash component analysis of the chars. Alkali index,
evaluated for the char samples reveals that char D2 reactivity was enhanced by
inherent catalysis of the inorganic components than the other chars.
Gasification reactivity of the four char samples (1000 m < d
p
< 1120 m) was
determined on a Thermax 500 TGA using various concentrations (25 - 100 mol. %)
of CO
2
in the CO
2
-N
2
reaction gas mixture in the temperature range of 900 - 950
C. Reactivity of the chars was found to increase with increasing temperature as
well as CO
2
composition in the reaction gas mixture. Comparison of the reactivity
of the four chars shows that, the reactivity of the chars generally increases in the
order: char C2 < char C < char B < char D2. The reactivity of char D2 was found
to be higher than the reactivity of the other three chars by a factor > 4.
Correlation of the parent coal petrographic properties to the reactivity of the
respective chars gave insignificant trends except for the rank parameter, the
vitrinite reflectance. Lower reflecting inertinite (Char D2) was found to be more
reactive that the other chars (chars B, C and C2). Char C2 with the highest vitrinite
reflectance was least reactive. Even the maceral index and modified reactive
maceral index, coal- and char- TRC and TIC failed to give a meaningful trend.
Chapter 7: Conclusion and Recommendations

194

Correlation of char properties with char gasification reactivity show systematic and
significant trends. The reactivity of the chars was found to increases with
increasing fraction of amorphous carbon, degree of disorder index and alkali index;
and decreases with increasing aromaticity of the chars. Thus, the extent of
structural disorderliness is a good predictor of the reactivity of the chars. On the
influence of the physical properties of the chars, it was observed that reactivity
increases with increasing micropore surface area, porosity; and decreases with
increasing average micropore diameter of the chars.
Structural parameters, , determined for the respective chars: B, C, C2 and D2,
using the random pore model (RPM) is: 1.11; 1.75; 2.58; and 1.57. The time scale
parameter, the time factor, t
f
, determined for each experiment was approximately
the same as the initial reactivity evaluated from the experimental data. The RPM
predicted reactivity t
f
, also confirmed the higher reactivity of char D2 relative to
the reactivity of the other chars by a factor > 4. The activation energy obtained for
the char-CO
2
gasification reactions were between 163.3 kJmol
-1
to 235.7 kJmol
-1
;
while the order of reaction with respect to CO
2
concentration ranged from 0.52 to
0.67. Lumped pre-exponential factor determined for the four chars were: 5.22 10
8
,
4.36 10
7
, 1.12 10
8
and 2.69 10
5
min
-1
bar
-1
for chars B, C, C2, and D2
respectively. The gasification reaction with CO
2
of all four chars studied in this
investigation within the specified operating conditions were found to be kinetically
chemical reaction controlled and were satisfactorily described by the RPM
(Regime I). The RPM was validated by a parity plot of the experimental
conversions against the model predicted conversions which followed a linear trend
for the four chars. The fitting of the RPM conversions and conversion rate data to
the experimental conversions and conversion rates also gave adequate fits.





Chapter 7: Conclusion and Recommendations

195

7.3 Contributions to Knowledge Base of Coal Science and Technology
The following results derived from this study are considered important contributions
to the knowledge base of coal science and technology:
The detailed characterisation of the chemical and physical properties of both
the parent coals and the subsequent chars has revealed that the properties of
the chars are different from that of the coals. Detailed petrographic analysis
has shown that there is a considerable reduction in the total reactive
components (TRC); while the total inert components (TIC) increase from coal
to char.

Carbon crystallite analysis using XRD (which has not been previously studied
for these four samples) was applied to study the carbon crystallite evolution
from coals to chars from which the carbon aromaticity, fraction of amorphous
carbon and the degree of disorder in the samples were evaluated. Results from
this analysis shows that the charring process changes the carbon crystallite
such that the increasing order of structural ordering in the coals change when
compared with the increasing order of structural ordering in the chars.
Furthermore, the crystallite lattice was elucidated by the analysis result.

The use of the characteristic lumped parameters such as the alkali index and
the degree of disorder index to adequately predict the reactivity and the
gasification behaviour of the chars is regarded as an important contribution to
coal science and technology. Lumped parameters determined from the
petrographic results of the coals (MI and RMI*) did not give meaningful
correlations with the char reactivity results.

The significant changes in chemical and physical properties of chars has had
the effect that the correlation of char reactivity with the parent coal properties
(including the lumped parameters- MI and RMI*) did not give meaningful or
systematic trends. Thus, a better prediction of the reactivity of chars should be
based on the char properties, because the subsequent chars are different from
the parent coals. This is one of the major contributions of this study to the
knowledge base of coal science and technology.
Chapter 7: Conclusion and Recommendations

196

7.4 Recommendations for Future Studies
Based on the findings of this study, the following recommendations are proposed for
further investigations:

A detailed characterisation of the carbon crystallite of the coals and
subsequent chars using
13
C NMR and HRTEM. This will enable the study of
other properties of the carbon BSU not covered by XRD as well as validate the
results obtained from the XRD analysis.
Investigation of chars produced from coals of different ranks and maceral
concentrates should be undertaken with focus on detailed characterisation
(chemical and physical properties) of the parent coals and chars and
gasification reactivity tests. This will give a better understanding of changes in
properties including carbon crystallite evolution from coals to chars and
broader correlations to the reactivity of the chars.
The investigation of char gasification characteristics with mixtures of two or
more of gasifying agents: O
2
, H
2
O, CO
2
, CO, H
2
, and Air; in order to study the
combined effect of the gasifying agents and products as well as any inhibition
and or promotion arising from the combination(s), during the gasification
process. This will give a more pragmatic application to gasification studies
and will yield results that can describe the practical gasification process better.
The investigation of changes in properties impacted to chars prepared at
different charring temperature (up to 1400 C) and the subsequent reactivity of
the chars. This will immensely help in a better understanding of changes in the
carbon BSU. This will also imply that some reactivity experiments will have
to be carried out at higher temperatures (> 1200 C) which may shift the
kinetic control to diffusion controlled (Regime II). A study of this type will be
ideal as most practical gasification processes occur in this regime.
Pilot plant studies of different combustion and gasification systems should
also be undertaken. This will give a more realistic result that can easily be
applied or retrofitted to practical gasification systems.

References

197

REFERENCES

ALEXANDER, L. E. AND SOMMER, E. C. (1959). Systematic analysis of carbon
black structures. Journal of Physical Chemistry, 60:1646-1649.
ALONSO, M. J. G., BORREGO, A. G., ALVAREZ, D. AND MENENDEZ, R.
(1999). Pyrolysis behaviour of pulverised coals at different temperatures. Fuel, 78:
1501-1513.
ANDERSON, R. B., BAYER, J. AND HOFER, L. J. E. (1965). Determining surface
areas from CO
2
isotherms. Fuel, 44: 443-452.
ARENILLAS, A., RUBIERA, F., PEVIDA, C., ANIA, C. O. AND PIS, J. J. (2004).
Relationship between structure and reactivity of carbonaceous materials. Journal of
Thermal Analysis and Calorimetry, 76: 593-602
ASO, H., MATSUOKA, K., SHARMA, A. AND TOMITA, A. (2004). Evaluation of
size of graphene sheet in anthracite by a temperature-programmed oxidation method.
Energy & Fuels, 18:1309- 314.
BAILEY, J. G., TATE, A., DIESSEL, C. F. K. AND WALL, T. F. (1990). A char
morphology system with applications to coal combustion. Fuel, 69: 225-239.
BENFELL, K. E., LIU, G-S., ROBERTS, D. G., HARRIS, D.J., LUCAS, J. A.,
BAILEY, J. G. AND WALL, T. F. (2000). Modelling char combustion: The influence
of parent coal petrography and pyrolysis pressure on the structure and intrinsic
reactivity of its chars. Proceedings of the Combustion Institute, 28: 2233-2241.
BHATIA, S.K., AND D. D. PERLMUTTER, D. D. (1980). A random pore model for
fluid-solid reactions: I. isothermal, kinetic control, AIChE Journal, 26: 379-386.
BHATIA, S.K., AND D. D. PERLMUTTER, D. D. (1981). A random pore model for
fluid-solid reactions: II. Diffusion and transport effects, AIChE Journal, 27: 247-254.
BLACKWOOD, J. D. AND INGEME, A, J. (1960). The reaction of carbon with
carbon dioxide at high pressure. Australian Journal of Chemistry, 13: 194-209.
References

198

BLANC, P., VALISOLALAO, J., ALBRECHT, P., KOHUT, J.P., MULLER, J.F.,
AND DUCHENE, J.M. (1991). Comparative geochemical study of three maceral
groups from a high-volatile bituminous coal. Energy & Fuels, 5: 875884.
BOUHADDA, Y., BORMANN, D., SHEU, E., BENDEDOUCH, D., KRALLAFA,
A. AND DAAOU, M. (2007). Characterization of Algerian Hassi-Messaoud
asphaltene structure using Raman spectrometry and X-ray diffraction. Fuel, 86: 1855-
1864.
BUSKIES, U. (1996). The efficiency of coal-fired combined-cycle power plants.
Applied Thermal Engineering, 16: 959-974.
CAI, H-Y., GELL, A.J., CHATZAKIS, I.N., LIM, J-Y. AND DUGWELL, D.R.
KANDIYOTI, R. (1996). Combustion reactivity and morphological change in coal
chars: Effect of pyrolysis temperature, heating rate and pressure. Fuel, 75: 15-24.
CAI, H-Y., MEGARITIS, A., MESSENBOCK, R., DIX, M., DUGWELL, D.R. AND
KANDIYOTI, R. (1998). Pyrolysis of coal maceral concentrates under pf combustion
conditions I: Changes in volatile release and char combustibility as a function of rank.
Fuel, 77: 1273-1282.
CAIRNCROSS, B. (2001). An overview of the Permian (Karoo) coal deposits of
Southern Africa. African Earth Sciences, 33: 529-562.
AKAL, G.., YCEL, H. AND GRZ, A.G. (2007). Physical and chemical
properties of selected Turkish lignites and their pyrolysis and gasification rates
determined by thermogravimetric analysis. J. Anal. Appl. Pyrolysis, 80: 262-268.
CAMERON, M.A. AND HUNT, J.W. (1985). A model for the statistical distribution
of microlithotypes in coal. Mathematical Geology, 17:267-285.
CAMPBELL, P.E., McMULLAN, J.T. AND WILLIAMS, B.C. (2000). Concept for
a competitive coal fired integrated gasification combined cycle power plant. Fuel, 79:
1031-1040.
CARLSON, G.A. (1992). Computer simulation of the molecular structure of
bituminous coal. Energy & Fuels, 6: 771-778.
References

199

CHEN, J. C., CASTAGNOLI, C. AND NIKSA, S. (1992). Coal devolatilization
during rapid transient heating: II. Secondary pyrolysis. Energy & Fuels, 6: 264271.
CHITSORA, C.T., MHLEN, H-J., VAN HEEK, K.H. AND JUNTGEN, H. (1987).
The Influence of Pyrolysis Conditions on the Reactivity of Char in H
2
O. Fuel
Processing Technology, 15: 17-29.
CHOI, C. Y., MUNTEAN, J. V., THOMPSON, R. A., AND BOTTO, R. E. (1989).
Characterization of coal macerals using combined chemical and NMR spectroscopic
methods. Energy & Fuels, 3: 528533.
Claudius Peters technologies. (2001). Pulverised fuel Injection: Technology you can
trust. http://www.claudiuspeters.com/_apps/dynamic/library/videos/309%20PCI.pdf
(Accessed: 24-03-2009).
CLOKE, M. AND LESTER, E. (1994). Characterisation of coals for combustion
using petrographic analysis: a review. Fuel, 73: 315-320.
CLOKE, M., WU, T., BARRANCO, R. AND LESTER, E. (2003). Char
characterisation and its application in a coal burnout model, Fuel, 82: 1989-2000.
CZECHOWSKI, F. AND KIDAWA, H. (1991). Reactivity and susceptibility to
porosity development of coal maceral chars on steam and carbon dioxide gasification
Fuel Processing Technology, 29: 57-73.
DAVIS, K. A., HURT, R. H., YANG, N. Y. C. AND HEADLEY, T. J. (1995).
Evolution of char chemistry, crystallinity, and ultrafine structure during pulverized-
coal combustion. Combustion & Flame, 100:31-40.
de LA PUENTE, G., FUENTE, E. AND PIS, J.J. (2000). Reactivity of pyrolysis
chars related to precursor coal chemistry. J. Anal. Appl. Pyrol. 53: 81-93.
de LA ROSA, L., PRUSKI, M., LANG, D., GERSTEIN, B. AND SOLOMON, P.
(1992). Characterization of the Argonne premium coals by using
1
H and
13
C NMR
and FT-IR spectroscopies. Energy & Fuels, 6: 460-468.
DME Coal Statistics. (2006). http:/www.dme.gov.za/energy/coal.stm (Accessed: 23-
03-2009).
References

200

DME: DIGEST OF SOUTH AFRICAN ENERGY STATISTICS. (2006).
http://www.dme.gov.za/pdfs/energy/planning/2006%20Digest.pdf (Accessed: 23-03-
2009).
du Cann, V.M. (2007). Test Report- PSA 2007-016, Petrographics SA, Pretoria,
South Africa.
du Cann, V.M. (2008). Test Report- PSA 2008-040, Petrographics SA, Pretoria,
South Africa.
DUTTA, S., WEN, C.Y. AND BELT, R.J. (1977). Reactivity of coal and char: I. In
carbon dioxide atmosphere. Ind. Eng. Chem. Proc. Des. Dev., 16: 20-30.
ERGUN, S AND TIENSUU, V. H. (1959). Interpretation of the intensities of X-rays
scattered by coals. Fuel, 38: 64-78.
ERGUN, S. AND MENTSER, M. (1968). Reactions of Carbon with Carbon Dioxide
and Steam. In: WALKER, P. L. Jr. (Editor), Chemistry and physics of carbon,
Volume 1, Marcel Dekker, Inc., NY: 203-263.
ERGUN. S. (1968). Structure of carbon. Carbon, 6: 141-159.
ESKOM ANNUAL REPORT, (2008). http://www.eskom.co.za/annreport08/ar_
2008/downloads.htm. (Accessed: 21-03-2009)
ESKOM FACT SHEET. (2007). Medupi power station. http://www.eskom.co.za/
content/CO_0013MedupiPSRev0.pdf. Pp: 1-6 (Accessed: 31-03-2009)
ESSENHIGH, R. H. (1981) Fundamentals of coal combustion. In: ELLIOTT, M. A.
(Editor), Chemistry of coal utilisation, 2
nd
Supplementary Volume, John Wiley and
Sons, New York: 1153-1311.
ESSENHIGH, R. H. (1991). Rate equations for the carbon-oxygen reaction: an
evaluation of the Langmuir adsorption isotherm at atmospheric pressure. Energy &
Fuels, 5: 41-46.
References

201

EVERSON, R.C., NEOMAGUS, H.W.J.P., KAITANO, R., FALCON, R. AND du
CANN, V. M. (2008a). Properties of high ash coal-char particles derived from
inertinite-rich coal: II. Gasification kinetics with carbon dioxide. Fuel, 87: 3403-3408.
EVERSON, R.C., NEOMAGUS, H.W.J.P., KAITANO, R., FALCON, R., van
ALPHEN C. AND du CANN, V. M. (2008b). Properties of high ash coal-char
particles derived from inertinite-rich coal: I. Chemical, structural and petrographic
characteristics. Fuel, 87: 3082-3090.
EVERSON, R.C., NEOMAGUS, H.W.J.P., KASAINI, H. AND NJAPHA, D. (2006).
Reaction kinetics of pulverised coal-chars derived from inertinite-rich coal discards:
Gasification with carbon dioxide and steam. Fuel, 85:1076-1082.
FALCON, R.M.S. (1989). Macro- and micro-factors affecting coal-seam quality and
distribution in Southern Africa with particular reference to the No. 2 seam, Witbank
coalfield, South Africa. International Journal of Coal Geology, 12: 681-731.
FALCON, R.M.S. AND SNYMAN, C.P. (1986). An introduction to coal
petrography: atlas of petrographic constituents in the bituminous coals of Southern
Africa. The Geological Society of South Africa, Johannesburg: Pp 1-27.
FALCON, R.M.S., VAINIKKA, P., van ALPHEN, C. AND du CANN, V.M. (2010).
Case study of four South African high-ash power station coals in a CFBC boiler.
Presentation at the FFF 15TH Coal Science and Technology Conference,
Johannesburg. 17-18 November, 2010.
FENG, B. AND BHATIA, S.K. (2002). On the validity of thermogravimetric
determination of carbon gasification kinetics. Chemical Engineering Science, 57:
2907-2920.
FENG, B., BHATIA, S.K. AND BARRY, J.C. (2003). Variation of the crystalline
structure of coal char during gasification. Energy & Fuels, 17: 744-754.
FLETCHER, T. H. (1993). Swelling properties of coal chars during rapid pyrolysis
and combustion. Fuel, 72: 1485-1495.
References

202

FRANKLIN, R.E. (1950). The interpretation of diffuse X-ray diagrams of carbon.
Acta Crystallographica., 3: 107-121.
FRANKLIN, R.E. (1951). The structure of graphitic carbons. Acta
Crystallographica, 4: 253-261.
FRIESEN, W.I. AND OGUNSOLA O.I. (1995). Mercury porosimetry of upgraded
western Canadian coals. Fuel, 74:604-609.
FU, W-B. AND WANG, Q-H. (2001). A general relationship between the kinetic
parameters for the gasification of coal chars with CO
2
and coal type. Fuel Processing
Technology, 72: 63-77.
GALE, T. K., BARTHOLOMEW, C. H. AND FLETCHER, T. H. (1996). Effects of
pyrolysis heating rates on intrinsic reactivities of coal chars. Energy & Fuels, 10: 766-
775.
GAN, H., NANDI, P.S. AND WALKER, P. L. Jr. (1972). Nature of the porosity in
American coals. Fuel, 51: 272-277.
GARRIDO, J., LINARES-SOLANO, A., MARTN-MARTNEZ, J.M., MOLINA-
SABIO, M., RODRGUEZ-REINOSO, F. AND TORREGROSA, R. (1987). Use of
N
2
vs. CO
2
in characterisation of activated carbons, Langmuir 3: 76-81.
GAVALAS, G. R. (1980). A random capillary model with application to char
gasification at chemically controlled rates. AIChE Journal, 26: 577-585.
GIVEN, P.H. (1960). The distribution of hydrogen in coals and its relations to coal
structure. Fuel, 39: 147-153.
GRAINGER, L. AND GIBSON, J. (1981). Coal Utilisation: Technology, Economics
& Policy. Graham & Trotman. London: 31-36.
GTC, (2008). Gasification: Redefining clean energy. http://www.gasification.org/
Docs/Final_whitepaper.pdf. (Accessed: 21-03-2009).
GUPTA, R. (2007). Advanced coal characterisation: A review. Energy & Fuels, 21:
451-460.
References

203

HAMPARTSOUMIAN, E., MURDOCH, P. L., POURKASHANIAN, M. AND
TRANGMAR, D. T. (1993). The reactivity of coal chars gasified in a carbon dioxide
environment. Combustion Science and Technology, 92: 105-121.
HANNA, J. V., VASSALLO, A. M. AND WILSON, M. A. (1992) CRAMPS
determined proton aromaticities of Australian coals: A comparison with dipolar
dephasing. Energy & Fuels, 6: 28-34.
HARRIS, D.J., AND SMITH, I.W. (1990). Intrinsic reactivity of petroleum coke and
brown coal char to carbon dioxide, steam and oxygen. Symposium (International) on
Combustion, 23: 1185-1190.
HASHIMOTO, K., MIURA, K., YOSHIKAWA, F. AND IMAI , I. (1979). Change
in pore structure of carbonaceous materials during activation and adsorption
performance of activated carbon. Ind. Eng. Chem., Proc. Des. & Dev., l8: 72-80.
HATCHER, P. G. (1988) Dipolar-dephasing 13C NMR studies of decomposed wood
and coalified xylem tissue: evidence for chemical structural changes associated with
defunctionalization of lignin structural units during coalification. Energy & Fuels, 2:
48-58.
HATTINGH, B. (2009). The determination of the reaction mechanisms involved in
the CO
2
gasification of inertinite-rich, high ash coal. M.Eng dissertation, North-West
University, Potchefstroom Campus.
HELLE, S., GORDON, A., ALFARO, G., GARCIA, X. AND ULLOA, C. (2003).
Coal blend combustion: link between unburnt carbon in fly ashes and maceral
composition. Fuel Processing Technology, 80: 209-233.
HINDMARSH, C. J., THOMAS, K. M., WANG, W. X., CAI, H-Y., GELL, A. J.
DUGWELL, D. R. AND KANDIYOTI, R. (1995). A comparison of the pyrolysis of
coal in wire-mesh and entrained-flow reactors. Fuel, 74: 1185-1190.
HIRSCH, P. B. (1954). X-ray scattering from coals. Proc. of the Royal Society - A,
226: 143-169.
References

204

HORVATH, G. AND KAWAZOE, K. (1983). Method for the calculation of
effective pore size distribution in molecular sieve carbon. Journal of Chemical
Engineering of Japan, 16: 470-475.
HU, H., ZHOU, Q., ZHU, S., MEYER, B., KRZACK, S. AND CHEN, G. (2004).
Product distribution and sulphur behaviour in coal pyrolysis. Fuel Processing
Technology, 85: 849-861.
HUANG, Y.H., YAMASHITA, H. AND TOMITA, A. (1991). Gasification
reactivities of coal macerals. Fuel Processing Technology, 29: 75-84.
HUBBELL, J.H. AND SELTZER, S.M. (1996). Tables of x-ray mass attenuation
coefficients and mass energy-absorption coefficients. NIST IR 5632. Ionizing
Radiation Division, Physics Laboratory, NIST.
(http://www.nist.gov/physlab/data/xraycoef/index.cfm). Pg: 1-21 Accessed: 24-05-
2010.
HTTINGER, K. J. (1990). A method for the determination of active sites and true
activation energies in carbon gasification: I. Theoretical treatment. Carbon, 28: 453-
456.
HTTINGER, K.J. AND NATTERMANN C. (1994). Correlations between coal
reactivity and inorganic matter content for pressure gasification with steam and
carbon dioxide. Fuel, 73: 1682-1684.
HTTINGER, K.J. AND NILL, J.S. (1990). A method for the determination of
active sites and true activation energies in carbon gasification: II. Experimental
results. Carbon, 28: 457-465.
JARONIEC, M., GADKAREE, K.P. AND CHOMA, J. (1996). Relation between
adsorption potential distribution and pore volume distribution for microporous
carbons. Colloids and Surfaces A: Physiochemical and Engineering Aspects, 111:
203-210.
JENKINS, R. G., NANDI, S. P. AND WALKER, P. L., Jr. (1973). Reactivity of
heat-treated coals in air at 500C. Fuel, 52: 288-293.
References

205

JONES, J.M., POURKASHANIAN, M., RENA, C.D. AND WILLIAMS, A. (1999).
Modelling the relationship of coal structure to char porosity. Fuel, 78: 1737-1744.
JONES, R.B., McCOURT, C.B., MORLEY, C. AND KING, K. (1985). Maceral and
rank influences on the morphology of coal char. Fuel, 64: 1460-1467.
JNTGEN, H. (1981). Reactivities of carbon to steam and hydrogen and applications
to technical gasification processes- A review. Carbon, 19: 167-173.
JNTGEN, H., KLEIN, J., KNOBLAUCH, K., SCHRTER, H-J. AND SCHULZE,
J. (1981) Conversion of coal and gases produced from coal into fuels, chemicals and
other products. In: ELLIOTT, M. A. (Editor), Chemistry of coal utilisation, 2
nd

Supplementary Volume, John Wiley and Sons, New York: 2071-2158.
KABE, T., ISHIHARA, A., QIAN, E. W., SUTRISNA, I.P. AND KABE, Y. (2004).
Coal and coal-related compounds: structures, reactivity and catalytic reactions.
Studies in surface science and catalysis series, Volume 150. Elsevier, Amsterdam: 1-
112; 269-309.
KAITANO, R. (2007). Characterisation and reaction Kinetics of high ash chars
derived from inertinite-rich coal discards. Doctoral thesis. North-West University.
South Africa.
KAJITANI, S., SUZUKI, N., ASHIZAWA, M. AND HARA, S. (2006) CO
2

gasification rate analysis of coal char in entrained flow coal gasifier. Fuel, 85: 163-
169.
KASHIWAYA, Y. AND ISHII, K. (1991). Kinetic analysis of coke gasification
based on non-crystal/crystal ratio of carbon. ISIJ International, 31: 440-448.
KAWAKAMI, M., KANBA, H., SATO, K., TAKENAKA, T., GUPTA, S.,
CHANDRATILLEKE, R. AND SAHAJWALLA, V. (2006). Characterisation of
thermal annealing effects on the evolution of coke carbon structure using Raman
spectroscopy and X-ray diffraction. ISIJ International, 46: 1165-1170.
KAYEMBE, N. AND PULSIFER, A.H. (1976). Kinetics and catalysis of coal char
and steam. Fuel, 55: 211-216.
References

206

KEATON ENERGY. (2009). About SA coalfields. (web)
http:/www.keatonenergy.co.za/cm/sa_coal.asp (Accessed: 21-06-2009).
KHAIRIL, KAMIHASHIRA, D., NAKAYAMA, K. AND NARUSE, I. (2001).
Fundamental Reaction Characteristics of Pulverized Coal at High Temperature. ISIJ
International, 41:136-141.
KOORNNEEF, J., JUNGINGER, M. AND FAAIJ, A. (2007). Development of
fluidised bed combustion - An overview of trends, performance and cost. Progress in
Energy and combustion Science, 33: 19-55.
KOWALCZYK, P., TERZYK, A.P., GAUDEN, P.A. AND SOLARZ, L. (2002).
Numerical analysis of Horvath-Kawazoe equation. Computers and Chemistry, 26:
125-130.
KRISTIANSEN, A. (1996). Understanding coal gasification, IEACR/ 86. IEA Coal
Research, London: 12-50.
KRUSZEWSKA, K. J. (2003). Fluorescing macerals in South African coals.
International Journal of Coal Geology, 54: 79-94.
KHL, H., KASHANI-MOTLAGH, M. M., MHLEN, H-J. AND van HEEK, K. H.
(1992). Controlled gasification of different carbon materials and development of pore
structure. Fuel, 71: 879-882.
KULASEKARAN, S., LINJEWILE, T. M., AGARWAL, P. K. AND BIGGS, M. J.
(1998) Combustion of a porous char particle in an incipiently fluidized bed. Fuel, 77:
1549-1560.
KUMAR, M. AND GUPTA, R. C. (1995). Graphitization study of Indian Assam
coking coal. Fuel Processing Technology, 43: 169-176.
KURODA, H. AND AKAMATU, H. (1959). Studies on the graphitization II.
Substructure and crystallite growth of carbon black. Bulletin of the Chem. Soc. of
Japan, 32: 142-149.
References

207

KYOTANI, T., KUBOTA, K., CAO, J., YAMASHITA, H. AND TOMITA, A.
(1993) Combustion and CO
2
gasification of coals in a wide temperature range. Fuel
Processing Technology, 36: 209-217.
KYOTANI, T., ZHANG, Z-G., HAYASHI, S. AND TOMITA, A. (1988). TPD
study on H
2
O-gasified and O
2
-chemisorbed coal chars. Energy & Fuels, 2: 136-141.
LAHAYE, J. (1998). The Chemistry of Carbon Surfaces. Fuel, 77: 543-547.
LAINE, N. R., VASTOLA, F. J. AND WALKER, P. L. Jr. (1963). The importance of
active surface area in the carbon-oxygen reaction. J. Phys. Chem., 67: 2030-2034.
LAURENDEAU, N.M. (1978). Heterogeneous kinetics of coal char gasification and
combustion. Progress in Energy & Combustion Science. 4: 221-270.
LEE, C. W., JENKINS, R. G. AND SCHOBERT, H. H. (1992). Structure and
Reactivity of Char from Elevated Pressure Pyrolysis of Illinois No. 6 Bituminous
Coal. Energy & Fuels, 6: 40-47.
LEE, S. (2007). Gasification of coal. In: LEE, S., SPEIGHT, J.G. AND LOYALKA,
S. K. (Editors). Handbook of alternative fuel technologies. CRC Press, New York. Pp:
25-79.
LEVENSPIEL, O. (1972). Chemical reaction engineering. 2
nd
Edition, Wiley Eastern
Ltd. New Delhi, India. Pp: 357-408.
LITTLEWOOD, K. (1977). Gasification: Theory and application. Progress in
Energy & Combustion Science. 3: 35-71.
LIU, G., BENYON, P., BENFELL, K. E., BRYANT, G. W., TATE, A. G., BOYD,
R. K., HARRIS, D. J. AND WALL, T. F. (2000). The porous structure of bituminous
coal chars and its influence on combustion and gasification under chemically
controlled conditions. Fuel, 79: 617-626.
LIZZIO, A. A., JUANG, H. AND RADOVIC, L. R. (1990). On the Kinetics of
Carbon (Char) Gasification: Reconciling Models with Experiments. Carbon, 28: 7-19.
References

208

LOUBSER, M. AND VERRYN, S. (2008). Combining XRF and XRD analyses and
sample preparation to solve mineralogical problems. South African Journal of
Geology, 111: 229-238.
LOVE, G.D., LAW, R.V. AND SNAPE, C.E. (1993). Determination of non-
protonated aromatic carbon concentrations in coals by single pulse excitation
13
C
NMR. Energy & Fuels, 7: 639644.
LU, L., KONG, C., SAHAJWALLA, V. AND HARRIS, D. (2002a). Char structural
ordering during pyrolysis and combustion and its influence on char reactivity. Fuel,
81: 1215-1225.
LU, L., SAHAJWALLA, V. KONG, C. AND MCLEAN, A. (2002b). Char structural
ordering during pyrolysis and combustion and its influence on char reactivity. ISIJ
International, 42: 816-825.
LU, L., SAHAJWALLA, V., KONG, C. AND HARRIS, D. (2001). Quantitative X-
ray diffraction analysis and its application to various coals. Carbon, 39: 1821-1833.
MAITY, S. AND MUKHERJEE, P. (2006). X-ray structural parameters of some
Indian coals. Current Science, 91: 337-340.
MARBAN, G., PIS, J.J. AND FUERTES, A.B. (1995). Characterising fuels for
atmospheric fluidized bed combustion. Combustion and Flame, 103:41-58.
MAROTO-VALER, M.M., LOVE, G.D. AND SNAPE C.E. (1994). Relationship
between carbon aromaticities and H/C ratios for bituminous coals. Fuel, 73: 1926-
1928.
MARSH, H. (1989). Introduction to carbon science. Butterworth & Co. (Publishers)
Ltd, London. Pp: 1-31, 107-145, 259-283.
MASTALERZ, M. AND MARC BUSTIN, R. (1993). Variation in elemental
composition of macerals; an example of the application of electron microprobe to coal
studies. International Journal of Coal Geology, 22: 83-99.
References

209

MATHEWS, J.P., HATCHER, P.G. AND SCARONI, A.W. (2001). Proposed Model
Structures for Upper Freeport and Lewiston-Stockton vitrinites. Energy & Fuels, 15:
863-873.
MATJIE, R.H. (2008). Sintering and slagging of mineral matter in South African
coals during the coal gasification process. PhD Thesis, University of Pretoria.
(http://www.upetd.up.ac.za/thesis/available/etd-11112008-125913/, Accessed: 30-06-
2010).
MATSUI, I., KUNII, D. AND FURUSAWA, T. (1987). Study of char gasification by
carbon dioxide: I. Kinetic study by thermogravimetric analysis. Ind. Eng. Chemistry
Research, 26: 91-95.
MEGARITIS, A., MESSENBCK, R. C., CHATZAKIS, I. N., DUGWELL, D. R.
AND KANDIYOTI, R. (1999). High-pressure pyrolysis and CO2-gasification of coal
maceral concentrates: conversions and char combustion reactivities. Fuel, 78: 871-
882.
MICROMERITICS INSTRUMENT CORPORATION. 2006. Accelerated surface
area and porosimetry system (ASAP2020), Operators manual V3.01, Pg: C1-44.
MIURA, K., HASHIMOTO, K. AND SILVESTON, P. L. (1989). Factors affecting
the reactivity of coal chars during gasification, and indices representing reactivity.
Fuel, 68: 1461-1474.
MIURA, K., MAKINO, M. AND SILVESTON, P.L. (1990). Correlation of
gasification reactivities with char properties and pyrolysis conditions using low rank
Canadian coals. Fuel, 69: 580- 89.
MHLEN, H-J. AND SULIMMA, A. (1987). Thermogravimetric apparatus for
characterisation of coal with regard to pyrolysis and gasification under pressures up to
100 bar. Fuel Processing Technology. 15: 145-455.
MHLEN, H-J., VAN HEEK, K. H. AND JNTGEN, H. (1985). Kinetic studies of
steam gasification of char in the presence of H
2
, CO
2
and CO. Fuel, 64: 944-949.
References

210

MURILLO, R., NAVARRO, M.V., LPEZ, J.M., GARCA, T., CALLN, M.S.,
AYLN, E. AND MASTRAL, A.M. (2006). Activation of pyrolytic lignite char with
CO
2
- kinetic study. Energy & Fuels, 20: 11-16.
MURILLO, R., NAVARRO, M.V., LPEZ, J.M., GARCA, T., CALLN, M.S.,
AYLN, E. AND MASTRAL, A.M. (2004). Kinetic model comparison for waste tire
char reaction with CO
2
. Ind. Eng. Chem. Res., 43: 7768-7773.
NARKIEWICZ, M. R. AND MATHEWS, P.J. (2008). Improved low-volatile
bituminous coal representation: Incorporating the molecular-weight distribution.
Energy & Fuels, 22: 31043111.
NG, S.H., FUNG, D.P.C. AND KIM, S.D. (1988). Study of the pore structure and
reactivity of Canadian coal-derived chars. Fuel, 67: 700-706.
NSAKALA, N. Y., ESSENHIGH, R.H AND WALKER, P. L. Jr. (1978).
Characteristics of chars produced from lignites by pyrolysis at 808 C following rapid
heating. Fuel, 57: 605-611.
OCHOA, J., CASSANELLO, M. C., BONELLI, P. R. AND CUKIERMAN, A. L.
(2001). CO
2
gasification of Argentinean coal chars: a kinetic characterization. Fuel
Processing Technology,74: 161-176.
OSBORNE, D.G., GRAHAM, J.M. AND ELLIOT, L.K. (1996). New coal utilisation
technologies. Minerals Engineering, 9: 215-233.
OUAZZANE, A. K., CASTAGNER, J. L., A.R. JONES, A. R. AND ELLAH, S.
(2002). Design of an optical instrument to measure the carbon content of fly ash.
Fuel, 81: 1907-1911.
PALMER, A. D., CHENG, M., GOULET, J. C. AND FURIMSKY, E. (1990).
Relation between particle size and properties of some bituminous coals. Fuel, 69:
183-188.
PATNAIK, P. (2008). Deans Analytical Chemistry Handbook, 2
nd
Edition.
http://www.accessengineeringlibrary.com/mghpdf/0071455981_ar015.pdf. The
McGraw-Hill Companies: Section 15: 1-18 (Accessed: 21-03-2009).
References

211

PHIRI, Z. (2010). Nitric oxide reduction by chars derived from high ash inertinite-
rich discard coals. M.Sc dissertation, North-West University, Potchefstroom Campus.
PODOLSKI, W. F., DAVID K. SCHMALZER, D. K., CONRAD, V., et al., (2008).
Energy Resources, Conversion, and Utilization. In: Green, D. W., Editor, Perrys
Chemical Engineers Handbook, 8
th
Edition. Downloaded from Digital Engineering
Library @ McGraw-Hill. http://www.nwu.ac.za/opencms/export/system/galleries/
externallinks/nwu-bib/digitalengineeringlibrary. Chapter 24: 9-31 (Accessed: 2-04-
2009).
PUGMIRE, R.J., SOLUM, M.S., GRANT, D.M., CRITCHFIELD, S. AND
FLETCHER, T.H. (1991). Structural evolution of matched tar-char pairs in rapid
pyrolysis experiments. Fuel, 70: 414-423.
Quantachrome Instruments. (2009). Stereopycnometer operating manual (Part No.:
05006 Rev D SPY-D160-E. Pp 1-14.
RADOVIC, L. R., STECZKO, K., WALKER, P. L. Jr. AND JENKINS, R. G.
(1985). Combined effects of inorganic constituents and pyrolysis conditions on the
gasification reactivity of coal chars. Fuel Processing Technology. 10: 311-326.
RADOVIC, L. R., WALKER, P. L. Jr. AND JENKINS, R. G. (1983). Importance of
carbon active sites in the gasification of coal chars. Fuel, 62: 849-856.
ROUZAUD, J-N. (1990). Contribution of transmission electron microscopy to the
study of the coal carbonization processes. Fuel Processing Technology, 24: 55-69.
ROUZAUD, J-N., VOGT, D. AND OBERLIN, A. (1988). Coke properties and their
microtexture. Part I: Microtextural analysis: a guide for cokemaking. Fuel Processing
Technology, 20: 143-154.
RUSSELL, N. V., GIBBINS, J. R. AND WILLIAMSON, J. (1999). Structural
ordering in high temperature coal chars and the effect on reactivity. Fuel, 78: 803
807.
SAKAWA, M., SAKURAI, Y. AND HARA, Y. (1982). Influence of coal
characteristics on CO2 gasification. Fuel, 61:717-720.
References

212

SAMARAS, P., DIAMADOPOULOS, E. AND SAKELLAROPOULOS, G.P.
(1996). The effect of mineral matter and pyrolysis conditions on the gasification of
Greek lignite by carbon dioxide. Fuel, 75: 1108-1114.
SANGTONG-NGAM, K. AND NARASINGHA, M.H. (2008). Kinetic study of
Thai-lignite char gasification using random pore model. Thammasat International
Journal of Science & Technology, 13: 16-26.
SCHOENING, F.R.L. (1982). X-ray structural parameter of coal. Fuel, 61: 695- 699.
SCHOENING, F.R.L. (1983). X-ray structure of some South African coals before
and after heat treatment at 500 and 1000 C. Fuel, 62: 1315- 1320.
SENNECA, O., SALATINO, P. AND MASI, S. (1998). Microstructural changes and
loss of gasification reactivity of chars upon heat treatment. Fuel, 77: 1483-1493.
SHA, X-Z., CHEN, Y-G., CAO, J., YANG, Y-M. AND REN, D-Q. (1990a). Effects
of operating pressure on coal gasification. Fuel, 69: 656-659.
SHA, X-Z., KYOTANI, T. AND TOMITA, A. (1990b). Rate retardation
phenomenon during gasification of Wandoan coal char. Fuel, 69: 1564-1567.
SHARMA, A., KADOOKA, H., KYOTANI, T. AND TOMITA, A. (2002). Effect of
microstructural changes on gasification reactivity of coal chars during low
temperature gasification. Energy & Fuels, 16: 54-61.
SHARMA, A., KYOTANI, T. AND TOMITA, A. (1999). A new quantitative
approach for microstructural analysis of coal char using HRTEM images. Fuel, 78:
1203-1212.
SHIM, H-S. AND HURT, R.H. (2000). Thermal annealing of chars from diverse
organic precursors under combustion-like conditions. Energy & Fuels, 14: 340-348.
SHINN, J. H. (1984). From coal to single-stage and two-stage products: a reactive
model of coal structure. Fuel, 63: 1187-1196.
SHIRAISHI, M. AND KOBAYASHI, K. (1973). An X-ray study of coal tar pitch.
Bulletin of the Chemical Society of Japan. 46: 2575-2578.
References

213

SHORT, M.A. AND WALKER P.L. Jr.. (1963). Measurement of the interlayer
spacing and crystal sizes in turbostratic carbons. Carbon, 1: 3-9.
SINA, A., SINEK, K., TEKE, A.T., MISIRLIOLU, Z., CANEL, M. AND
WANG, L. (2003). Study on CO
2
gasification reactivity of chars obtained from Soma-
Isklar lignite (Turkey) at various coking temperatures. Chemical Engineering and
Processing, 42: 1027-1031.
SMITH, I.W. (1978). The intrinsic reactivity of carbons to oxygen. Fuel, 57: 409-
414.
SMITH, I.W., HARRIS, D.J., VALIX, M.G. AND TRIM, D.L. (1991). Mass
Transport and carbon reactivity at high temperature. In: LAHAYE, J. AND
EHRBURGER P. (Editors), Fundamental issues in control of carbon gasification
reactivity. NATO ASI Series, Vol. 192. Kluwer Academic Publishers, The
Netherlands. Pp: 49-77.
SMITH, W. H., ROUX, H. J. AND STEYN, J. G. H. (1983). The classification of
coal macerals and their relation to certain chemical parameters of coal. Special
publication of the Geological Society of South Africa. 7: 111-115.
SNYMAN, C.P. (1989). The role of coal petrography in understanding the properties
of South African coal. International Journal of Coal Geology, 14: 83-101.
SNYMAN, C.P. AND BOTHA, J.W. (1993). Coal in South Africa. Journal of
African Earth Sciences, 16: 171-180.
SOLUM, M.S., PUGMIRE, R.J. AND GRANT, D.M. (1989).
13
C Solid-state NMR
of Argonne-premium coals. Energy & Fuels, 3: 187-193.
SPALDING-FECHER, R., WILLIAMS, A. AND van HERON, C. (2000). Energy
and environment in South Africa: charting a course to sustainability. Energy for
Sustainable Development, 4: 8-17.
SPEARS, D. A. (2000). Role of clay minerals in UK coal combustion. Applied Clay
Science, 16: 87-95.
References

214

STEEL, K.M. AND PATRICK, J. W. (2001). The production of ultra clean coal by
chemical demineralisation. Fuel, 80: 2019-2023.
STRUGALA, A. (1994). Empirical formulae for calculation of real density and total
pore volume of hard coals. Fuel, 73: 1781-1785.
STYSZKO-GROCHOWIAK, K., GOA, J., JANKOWSKI, H. AND KOZISKI,
S. (2004). Characterization of the coal fly ash for the purpose of improvement of
industrial on-line measurement of unburned carbon content. Fuel, 83: 1847-1853.
SU, S., POHL, J.H., HOLCOMBE, D. AND HART, J.A. (2001). A proposed maceral
index to predict combustion behaviour of coal. Fuel, 80: 699-706.
SUN, Q., LI, W., CHEN, H. AND LI, B. (2004). The CO
2
-gasification and kinetics of
Shenmu maceral chars with and without catalyst Fuel, 83: 1787-1793.
TAKAGI, H., MARUYAMA, K., YOSHIZAWA, N., YAMADA, Y. AND SATO,
Y. (2004). XRD analysis of carbon stacking structure in coal during heat treatment.
Fuel, 83: 2427-2433.
TAMARI, T. AND AGUILAR-CHVEZ, A. (2004). Optimum design of the
variable-volume gas pycnometer for determining the volume of solid particles.
Measurement Science Technology, 15: 11461152.
TAULBEE, D., POE, S.H., ROBL, T. AND KEOGH, B. (1989). Density gradient
centrifugation separation and characterisation of maceral groups from a mixed
maceral bituminous coal. Energy & Fuels, 3: 662-670.
THOMAS, J.M. AND THOMAS, W.J. (1967). Introduction to the principles of
heterogeneous catalysis. Academy Press, London. Pp: 365-450.
TOMITA, A. (2001). Catalysis of carbon-gas reactions. Catalysis Surveys from
Japan, 5: 17-24.
TREJO, F., ANCHEYTA, J., MORGAN, T.J., HEROD, A.A. AND KANDIYOTI,
R. (2007). Characterization of asphaltenes from hydrotreated products by SEC,
LDMS, MALDI, NMR, and XRD. Energy & Fuels, 21:2121-2128.
References

215

TSAI, C-Y AND SCARONI, A.W. (1987). The structural changes of bituminous coal
particles during the initial stages of pulverized-coal combustion. Fuel, 66: 200-206.
TSAI, S.C. (1982). Coal science and technology 2: Fundamentals of coal
beneficiation and utilization. Amsterdam: Elsevier Scientific Publishing Company. Pp
: 375.
TURKDOGAN, E.T. AND VINTERS, J.V. (1969). kinetics of oxidation of graphite
and charcoal in carbon dioxide. Carbon, 7: 101-117.
UNFCC Website. http://unfcc.int/kyoto_protocol/mechanisms/clean_development_
mechanism/items/2718.php (Accessed: 31-03-2009).
van ALPHEN, C. (2009). Advanced coal and ash analysis. Oral presentation to Coal
Research Group, School of Chemical & Minerals Engineering, NWU. 15
th
July, 2009.
van de VENTER, E. (2005). Sasol-Lurgi coal gasification technology and low rank
coal. Presentation to the Gasification Technologies Council Conference, 10-12
th
Oct.
2005, San Francisco.
van HEEK, K.H. AND MHLEN, H-J. (1985). Aspects of coal properties and
constitution important for gasification. Fuel, 64: 1405-1414.
van HEEK, K.H. AND MHLEN, H-J. (1987). Effect of Coal and Char Properties
on Gasification. Fuel Processing Technology, 15: 113-133.
van KREVELEN, D.W. (1981). Coal: Typology - Physics, Chemistry, Constitution.
Coal Science and Technology Series, Volume 3. ANDERSON, L. L. (Editor),
Elsevier, Amsterdam. Pp: 58-88, 111-126, 160-174, 309-341.
VAN NIEKERK, D. (2008). Structural elucidation, molecular representation and
solvent interactions of vitrinite-rich and inertinite-rich South African coals. Doctoral
dissertation. Penn State University, U.S.A.
VASSALLO, A.M., HANNA, J.V., WILSON, M.A. AND LOCKHART, C. (1991).
High-resolution solid-state proton NMR spectroscopy of density fractions from
Callide coal. Energy & Fuels, 5: 643647.
References

216

WALKER, P.L. Jr. AND HIPPO, E.J. (1975). Factors affecting reactivity of coal
chars. Am. Chem. Soc. Div. Fuel Chem. Prepr., 20: 45-51.
WALKER, P.L. Jr. AND KINI, K.A. (1965). Measurement of the ultrafine surface
area of coals. Fuel, 44: 453-459.
WALKER, P.L. Jr. VERMA, S.K., UTRILLA, J.R. & DAVIS, A. (1988). Densities,
porosities and surface areas of coal macerals as measured by their interaction with
gases, vapours and liquids. Fuel, 67:1615-1623.
WALKER, P.L. Jr., SHELEF, M. AND ANDERSON, R.A. (1968) Catalysis of
carbon gasification. In: WALKER, P.L. Jr. (Editor), Chemistry and Physics of
Carbon Volume 1, Marcel Dekker Inc. New York. Pp: 287-383.
WARREN, B. E. (1941). X-ray diffraction in random layer lattices. Physical Review,
59: 693-698.
WATKINSON, A.P., LUCAS, J.P. AND LIM, C.J. (1991). A prediction of
performance of commercial gasifiers. Fuel, 70: 519-527.
WCI COAL STATISTICS. (2007) World Coal Institute. http://www.world coal
.org/resources /coal-statistics/index.php. (Accessed: 23-03-2009).
WCI ECOAL. (2003). WCI Ecoal, Volume 45, March, 2003
http://www.worldcoal.org/resources/ecoal/ecoal---back-issues/. The Quarterly
Newsletter of the World Coal Institute. (Accessed: 23-03-2009).
WEB MINERAL DATA. (2009). http://webmineral.com/data. (Accessed: 4-08-
2010).
WEBB, P. A. (2001). Volume and density determination for particle technologist.
Micromeritics Instrument Corp., Norcross, Georgia, USA.
http://www.micromeritics.com/Repository/files/density-determination.pdf. (Accessed
22-08-2010).
WIGLEY, F., WILLIAMSON, J., AND GIBB, W.H. (1997). The distribution of
mineral matter in pulverised coal particles in relation to burnout behaviour. Fuel, 76:
1283-1288.
References

217

WU, H., BRYANT, G., BENFELL, K. AND WALL, T. (2000). An Experimental
Study on the Effect of System Pressure on Char Structure of an Australian Bituminous
Coal. Energy & Fuels, 14: 282-290.
WU, S., GU, J., ZHANG, X., WU Y. AND GAO, J. (2008). Variation of Carbon
Crystalline Structures and CO2 Gasification reactivity of Shenfu coal chars at
elevated temperatures. Energy & Fuels, 22: 199-206.
YE, D.P., AGNEW, J.B. AND ZHANG, D.K. (1998). Gasification of a South
Australian low rank coal with carbon dioxide and steam- kinetics and reactivity
studies. Fuel, 77: 1209-1219.
YU, J-L., LUCAS J., WALL, T., LIU, G., AND SHENG, C. (2004). Modelling the
development of char structure during rapid heating of pulverised coal. Combustion
and Flame, 136: 519-532.
ZHANG, J-W., ZONG, Z-M., WANG, T-X., XIE, R-L., DING, M-J., CAI, K-Y.,
HUANG, Y-G., GAO J-S., WU, Y-Q. AND WEI, X-Y. (2007). Reactivities of
Shenfu chars towards gasification with carbon dioxide. Journal of China University of
Mining & Technology, 17: 197-200.
ZHANG, L. AND CALO, J.M. (1996). The effect of CO
2
partial pressure on
gasification reactivity. Am. Chem. Soc., Div. Fuel Chem. Prep. 41:138-142.
ZHANG, L., HUANG, J., FANG, Y. AND WANG, Y. (2006). Gasification
Reactivity and Kinetics of Typical Chinese Anthracite Chars with Steam and CO
2
.
Energy & Fuels, 20: 1201-1210.
ZHANG, Y., HARA, S., KAJITANI, S. AND ASHIZAWA, M. (2010) Modelling of
catalytic gasification kinetics of coal char and carbon. Fuel, 89: 152-157.

15TH



218










APPENDICES








Appendix A-1

219

APPENDIX A
Coal and Char Characterisation and Results
A-1 Description of Standard Methods Used for Characterisation
ASTM 3682: Standard test method for major and minor elements in combustion
residues from coal utilization processes.
CKS 561 (1982): Specification for anthracitic and bituminous coals.
ISO 7404-2 (1994): Methods for the petrographic analysis of bituminous coal and
anthracite-Part 2: Preparation of coal samples.
ISO 7404-3 (1994): Methods for the petrographic analysis of bituminous coal and
anthracite-Part 3: Method of determining maceral group composition.
ISO 7404-4 (1994): Methods for the petrographic analysis of bituminous coal and
anthracite-Part 4: Method of determining microlithotype-, carbominerite- and
minerite- composition.
ISO 7404-5 (1994): Methods for the petrographic analysis of bituminous coal and
anthracite-Part 5: Method of determining microscopically the reflectance of vitrinite.
ISO 11760 (2005): Classification of coals (ISO, 2009). ISO 12902 : Determination of
total carbon, hydrogen and nitrogen-instrumental methods.
ISO 19759: Determination of total sulphur through IR spectroscopy.
SABS ISO 562 (1998): Hard coal and coke: Determination of volatile matter.
SABS 925 (1978): Moisture content of coal samples intended for general analysis:
vacuum oven method.
SABS ISO 1171 (1997): Solid mineral fuels: Determination of ash content.

Appendix A-2

220

SABS ISO 1928 (1995): Determination of gross calorific value by the bomb
calorimetric method, and calculation of net calorific value.

A-2 Vitrinite Reflectance Scan Histograms of Coal Samples


Figure A-1: Vitrinite reflectance scans histogram of the coal samples.
0
10
20
30
40
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Vitrinite reflectance, Rr %
Coal B Vitrinite Rr% 0.67
0
10
20
30
40
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
R
e
l
a
t
i
v
e

F
r
e
q
u
n
c
y

%
Vitrinite reflectance, Rr %
Coal C Vitrinite Rr% 0.72
0
10
20
30
40
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Vitrinite reflectance, Rr %
Coal C2 Vitrinite Rr% 0.75
0
10
20
30
40
50
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Vitrinite reflectance, Rr %
Coal D2 Vitrinite Rr% 0.56
Appendix A-3

221

A-3 Total Maceral Reflectance Scan Histograms of Coal and Char
samples

Figure A-2: Total maceral reflectance scans histogram for coals and chars: B and C.



0
2
4
6
8
10
12
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Total maceral reflectance, Rsc %
Coal B Scan Rsc % 1.27
0
2
4
6
8
10
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 9.5 10.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Total maceral reflectance, Rsc %
Char B Scan Rsc % 5.28
0
2
4
6
8
10
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Total maceral reflectance, Rsc %
Coal C Scan Rsc % 1.30
0
2
4
6
8
10
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 9.5 10.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Total maceral reflectance, Rsc %
Char C Scan Rsc % 5.19
Appendix A-3

222


Figure A-3: Total maceral reflectance scans histogram for coals and chars: C2 and D2.




0
2
4
6
8
10
12
14
16
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Total maceral reflectance, Rsc %
Coal C2 Scan Rsc % 1.63
0
2
4
6
8
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 9.5 10.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Total maceral reflectance, Rsc %
Char C2 Scan Rsc % 4.43
0
2
4
6
8
10
12
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Total maceral reflectance, Rsc %
Coal D2 Scan Rsc % 1.15
0
2
4
6
8
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5 9.0 9.5 10.0
R
e
l
a
t
i
v
e

F
r
e
q
u
e
n
c
y

%
Total maceral reflectance, Rsc %
Char D2 Scan Rsc % 5.07
Appendix A-4

223


A-4 Outline of Classification System for Char Carbon Forms (du Cann,
2008).
Group A Char carbon forms
Category A1(i): Dense char with frequent tiny pores from vitrinite.
Category A1(ii): Dense char with frequent tiny pores from inertinites.
Category A2: Char networks: fine walled more open networks from coal
reactives.
Category A3: Char networks: thicker walled networks form inerts.
Group B Char carbon forms
Category B4: Coke- circular anisotropic - fine.
Category B5: Coke- circular anisotropic - medium.
Category B6: Coke- incipient anisotropic.
Category B7: Thick walled coke- isotropic (from coal vitrinite).
Group C Char carbon forms
Category C8: Oxidised- mainly from vitrinite.
Category C9: Oxidised- mainly from inertinite.
Category C10: Oxidised- Low reflecting network.
Group D Char carbon forms
Category D11(i): Original coal - unaffected from vitrinite
Category D11(i): Original coal - unaffected from inertinite
Category D12: Original coal - partially reacted macerals
Group E Char carbon forms
Category E13: Inorganic matter derived from minerals in the original coal
Category E14: Inorganic matter derived from other sources related to the
parent coal
Category E15: Process derived depositional carbons (pyrolytic and spherulitic
carbons)

Appendix B-1

224


APPENDIX B

Char-CO
2
Gasification Reactivity Results

B-1 Reproducibility of Experimental Results and Reactivity of the
Chars

Figure B-1: Reproducibility results for chars B, C, C2 and D2 at different experimental
conditions, 0.875 bar.
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 917 C, 75% CO
Run 1 Run 2 Run 3
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 900 C, 100% CO
Run 1 Run 2 Run 3
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200 1400 1600
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 933 C, 25% CO
Run 1 Run 2 Run 3
0
0.2
0.4
0.6
0.8
1
0 100 200 300 400 500 600
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 900 C, 25% CO
Run 1 Run 2 Run 3
Appendix B-1

225



Table B-1: Analysis of reproducibility and experimental error.
Char B, 917 C, 75% CO
2

Run 1 Run 2 Run 3 Ave. Ave. Dev % Ave. Dev
R 10
-2
(min
-1
) 3.070 3.060 3.110 3.08 2.47 10
-5
1.88 10
-5
0.609
t
f
10
-2
(min
-1
)

2.870 2.720 2.760 2.78 7.78 10
-5
5.78 10
-5
2.070
(-) 1.1696 1.1595 1.1768 1.1686 0.0087 0.0061 0.520
t
0.5
(min) 162.97 158.10 160.47 160.51 2.430 1.640 1.020
t
0.9
(min) 549.93 545.65 553.83 549.80 4.090 2.770 0.500
Char C, 900 C, 100% CO
2

R 10
-2
(min
-1
) 2.370 2.290 2.260 2.310 5.73 10
-5
4.21 10
-5
1.820
t
f
10
-2
(min
-1
)

2.490 2.490 2.460 2.480 2.07 10
-5
1.59 10
-5
0.640
(-) 2.1216 2.2469 2.2029 2.1870 0.060 0.044 1.990
t
0.5
(min) 211.20 218.75 215.47 215.14 3.790 2.630 1.220
t
0.9
(min) 543.13 555.43 547.10 548.56 6.280 4.590 0.840
Char C2, 933 C, 25% CO
2

R 10
-3
(min
-1
) 1.670 1.680 1.710 1.680 2.11 10
-5
1.61 10
-5
0.960
t
f
10
-3
(min
-1
)

1.820 1.780 1.810 1.800 2.25 10
-5
1.72 10
-5
0.950
(-) 2.0101 2.191 2.2348 2.1453 0.120 0.090 4.200
t
0.5
(min) 300.08 294.77 300.66 298.50 3.250 2.490 0.830
t
0.9
(min) 781.03 772.80 788.26 780.70 7.730 5.260 0.670
Char D2, 900 C, 25% CO
2

R 10
-2
(min
-1
) 6.870 6.790 6.900 6.850 5.30 10
-5
3.95 10
-5
0.58
t
f
10
-2
(min
-1
)

6.900 6.810 6.910 6.870 5.60 10
-5
4.29 10
-5
0.62
(-) 1.2117 1.2859 1.3051 1.2676 0.049 0.037 2.940
t
0.5
(min) 72.82 70.55 71.61 71.66 1.130 0.770 1.080
t
0.9
(min) 232.47 219.17 222.45 224.70 6.930 5.180 2.310


Appendix B-2

226

B-2 Determination of CO
2
Reactivity of the Chars

Figure B-2: Rate of reaction versus fractional conversion for chars B and C at different
experimental conditions, 0.875 bar.
0
0.002
0.004
0.006
0.008
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char B
950 C, 50% CO
933 C, 50% CO
917 C, 50% CO
900 C, 50% CO
0
0.002
0.004
0.006
0.008
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char B
950 C, 75% CO
933 C, 75% CO
917 C, 75% CO
900 C, 75% CO
0
0.002
0.004
0.006
0.008
0.01
0 0.2 0.4 0.6 0.8 1
R
e
a
c
t
i
o
n

r
a
t
e

c
h
a
n
g
e
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char B
900 C, 100% CO
917 C, 100% CO
933 C, 100% CO
950 C, 100% CO
0
0.001
0.002
0.003
0.004
0.005
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C
950 C, 50% CO
933 C, 50% CO
917 C, 50% CO
900 C, 50% CO
0
0.002
0.004
0.006
0.008
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C
950 C, 75% CO
933 C, 75% CO
917 C, 75% CO
900 C, 75% CO
0
0.002
0.004
0.006
0.008
0 0.2 0.4 0.6 0.8 1
R
e
a
c
t
i
o
n

r
a
t
e

c
h
a
n
g
e
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C
950 C, 100% CO
933 C, 100% CO
917 C, 100% CO
900 C, 100% CO
Appendix B-2

227


Figure B-3: Rate of reaction versus fractional conversion for chars C2 and D2 at different
experimental conditions, 0.875 bar.

0
0.001
0.002
0.003
0.004
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C2
950 C, 50% CO
933 C, 50% CO
917 C, 50% CO
900 C, 50% CO
0
0.001
0.002
0.003
0.004
0.005
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C2
950 C, 75% CO
933 C, 75% CO
917 C, 75% CO
900 C, 75% CO
0
0.001
0.002
0.003
0.004
0.005
0.006
0.007
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C2
950 C, 100% CO
933 C, 100% CO
917 C, 100% CO
900 C, 100% CO
0.000
0.005
0.010
0.015
0.020
0.025
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char D2
950 C, 50% CO
933 C, 50% CO
917 C, 50% CO
900 C, 50% CO
0.000
0.005
0.010
0.015
0.020
0.025
0.030
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char D2
950 C, 75% CO
933 C, 75% CO
917 C, 75% CO
900 C, 75% CO
0.000
0.006
0.012
0.018
0.024
0.030
0.036
0 0.2 0.4 0.6 0.8 1
R
e
a
c
t
i
o
n

r
a
t
e

c
h
a
n
g
e
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char D2
950 C, 100% CO
933 C, 100% CO
917 C, 100% CO
900 C, 100% CO
Appendix B-3

228

B-3 Effect of Isothermal Temperature of Reaction on the Char
Reactivity

Figure B-4: Effect of temperature on the CO
2
reactivity of chars B and C, 0.875 bar.
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200 1400
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 100% CO
2
950 C 933 C 917 C 900 C
0
0.2
0.4
0.6
0.8
1
0 250 500 750 1000 1250 1500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 75% CO
950 C 933 C 917 C 900 C
0
0.2
0.4
0.6
0.8
1
0 500 1000 1500 2000 2500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B 25% CO
2
900 C 917 C 950 C 933 C
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 100% CO
950 C 933 C 917 C 900 C
0
0.2
0.4
0.6
0.8
1
0 250 500 750 1000 1250 1500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 50% CO
950 C 933 C 917 C 900 C
0
0.2
0.4
0.6
0.8
1
0 500 1000 1500 2000 2500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 25% CO
950 C 933 C 917 C 900 C
Appendix B-3

229


Figure B-5: Effect of temperature on the CO
2
reactivity of chars C2 and D2, 0.875 bar.




0
0.2
0.4
0.6
0.8
1
0 150 300 450 600 750 900 1050
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 100% CO
933 C 917 C 900 C 950 C
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200 1400
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 75% CO
950 C 933 C 917 C 900 C
0
0.2
0.4
0.6
0.8
1
0 250 500 750 1000 1250 1500 1750
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 50% CO
917 C 900 C 933 C 950 C
0
0.2
0.4
0.6
0.8
1
0 60 120 180 240 300 360
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Chart D2, 75% CO
900 C 917 C 933 C 950 C
0
0.2
0.4
0.6
0.8
1
0 50 100 150 200 250 300 350 400
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 50% CO
950 C 933 C 917 C 900 C
0
0.2
0.4
0.6
0.8
1
0 100 200 300 400 500 600
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 25% CO
950 C 933 C 917 C 900 C
Appendix B-4

230

B-4 Effect of CO
2
Concentration in the Reaction Gas on Char
Reactivity

Figure B-6: Effect of CO
2
concentration on the reactivity of chars B and C, 0.875 bar.
0
0.2
0.4
0.6
0.8
1
0 400 800 1200 1600 2000 2400
C
o
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 900 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 250 500 750 1000 1250 1500 1750
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 917 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 150 300 450 600 750 900 1050
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 950 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 300 600 900 1200 1500 1800
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 917 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200 1400 1600
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 933 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 950 C
75% CO 50% CO 100% CO 25% CO
Appendix B-4

231


Figure B-7: Effect of CO
2
concentration on the reactivity of chars C2 and D2, 0.875 bar.

0
0.2
0.4
0.6
0.8
1
0 450 900 1350 1800 2250 2700
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 900 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
1.2
0 400 800 1200 1600 2000
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 917 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
1.2
0 400 800 1200 1600
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 933 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 100 200 300 400 500 600
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 900 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 100 200 300 400
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 933 C
100% CO 75% CO 50% CO 25% CO
0
0.2
0.4
0.6
0.8
1
0 50 100 150 200 250 300
C
o
n
v
e
r
s
i
o
n
,

X
(
m
i
n
)
Time, t (min)
Char D2, 950 C
100% CO 75% CO 50% CO 25% CO
Appendix B-5

232

B-5 Comparison of CO
2
Reactivity of the Four Chars

Figure B-8: Comparison of the CO
2
reactivity of chars at 100 and 75% CO
2
concentration,
0.875 bar.
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
917 oC, 100% CO
Char C Char C2 Char B Char D2
0
0.2
0.4
0.6
0.8
1
0 100 200 300 400 500 600 700
C
o
m
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
933 C, 100% CO
Char D2 CharB Char C Char C2
0
0.2
0.4
0.6
0.8
1
0 100 200 300 400 500 600 700
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
950 C, 100% CO
Char D2 Char B Char C Char C2
0
0.2
0.4
0.6
0.8
1
0 250 500 750 1000 1250 1500 1750
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
900 C, 75% CO
Char B Char C Char C2 Char D2
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
917 C, 75% CO
Char B Char C Char C2 Char D2
0
0.2
0.4
0.6
0.8
1
0 150 300 450 600 750 900
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
933 C, 75% CO
Char B Char C Char C2 Char D2
Appendix B-5

233


Figure B-9: Comparison of the CO
2
reactivity of chars at 50 and 25% CO
2
concentration,
0.875 bar.

0
0.2
0.4
0.6
0.8
1
0 250 500 750 1000 1250 1500 1750
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
900 C, 50% CO
Char D2 Char B Char C Char C2
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200 1400
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
917 C, 50% CO
Char D2 Char B Char C Char C2
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
950 C, 50% CO
Char D2 Char B Char C Char C2
0
0.2
0.4
0.6
0.8
1
0 500 1000 1500 2000 2500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
900 C, 25% CO
Char D2 Char B Char C Char C2
0
0.2
0.4
0.6
0.8
1
0 250 500 750 1000 1250 1500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
933C, 25% CO
Char B Char C Char C2 Char D2
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
950 C, 25% CO
Char B Char C Char C2 Char D2
Appendix C-1

234

APPENDIX C
Evaluation of Kinetic Parameters and Gasification Modelling
C-1 Summary of Structural Parameter, Time Factor and Initial
Reactivity of the Chars

Table C-1: Summary of the structural parameter, time factor and initial reactivity for
char C, 0.875 bar.
Char C
Temp.
(C)
2
CO
y
(mol. %)
t
0.5

(min)
t
0.9

(min)
R 10
-3

(min
-1
)
t
f
10
-3

(min
-1
)

(-)
900 C
25%
489.05 1379.7 1.02 1.05 1.6500
917 C 322.87 920.50 1.55 1.56 1.6389
933 C 226.13 737.43 2.21 2.18 1.1829
950 C 186.08 544.00 2.69 2.73 1.5553
900 C
50%
289.27 773.87 1.73 1.79 1.9181
917 C 249.07 637.02 2.01 2.11 2.1413
933 C 170.80 487.05 2.93 3.01 1.4762
950 C 113.00 310.12 4.43 4.69 1.7483
900 C
75%
229.78 608.77 2.18 2.27 1.9451
917 C 157.02 435.15 3.19 3.37 1.6877
933 C 120.58 339.62 4.15 4.35 1.6422
950 C 86.433 237.20 5.80 6.04 1.7835
900 C
100%
211.20 543.13 2.37 2.49 2.1216
917 C 126.28 339.37 3.96 4.13 1.8508
933 C 104.25 278.52 4.80 5.14 1.8274
950 C 76.983 207.07 6.50 6.91 1.8255




Appendix C-1

235




Table C-2: Summary of the structural parameter, time factor and initial reactivity for
char C2, 0.875 bar.
Char C2
Temp.
(C)
2
CO
y
(mol. %)
t
0.5

(min)
t
0.9

(min)
R 10
-3

(min
-1
)
t
f
10
-3

(min
-1
)

(-)
900 C
25%
545.65 1362.07 0.92 0.99 2.3040
917 C 401.20 1005.68 1.25 1.33 2.3005
933 C 300.08 781.03 1.67 1.82 2.0101
950 C 247.97 620.60 2.02 2.19 2.2519
900 C
50%
388.72 918.52 1.29 1.38 2.7982
917 C 281.40 663.42 1.78 1.90 2.8881
933 C 198.67 485.90 2.52 2.70 2.5192
950 C 148.62 379.07 3.37 3.68 2.1373
900 C
75%
290.90 679.28 1.72 1.81 3.0163
917 C 198.42 478.63 2.52 2.69 2.6895
933 C 163.48 385.98 3.06 3.27 2.8335
950 C 113.33 296.83 4.41 4.80 1.9614
900 C
100%
271.38 605.47 1.84 1.98 3.2980
917 C 161.77 390.38 3.09 3.29 2.6703
933 C 130.62 305.62 3.83 4.04 2.9251
950 C 90.650 215.15 5.52 5.97 2.7396











Appendix C-1

236




Table C-3: Summary of the structural parameter, time factor and initial reactivity for
char D2, 0.875 bar.
Char D2
Temp.
(C)
2
CO
y
(mol. %)
t
0.5

(min)
t
0.9

(min)
R 10
-3

(min
-1
)
t
f
10
-3

(min
-1
)

(-)
900 C
25%
72.817 232.47 6.87 6.90 1.2117
917 C 62.783 191.10 7.96 8.02 1.3815
933 C 52.250 155.62 9.57 9.77 1.4348
950 C 39.017 115.58 12.8 13.1 1.4437
900 C
50%
53.350 159.43 9.37 9.55 1.4414
917 C 41.533 130.28 12.0 12.0 1.2854
933 C 36.133 98.217 13.8 14.5 1.7935
950 C 26.317 70.217 19.0 19.6 1.9396
900 C
75%
40.733 123.70 12.3 12.4 1.3633
917 C 30.450 90.033 16.4 16.9 1.4376
933 C 24.350 69.633 20.5 21.4 1.5567
950 C 21.450 56.717 23.3 24.5 1.9311
900 C
100%
32.633 95.183 15.3 15.8 1.4939
917 C 25.433 71.433 19.7 20.4 1.6372
933 C 19.500 55.417 25.6 26.2 1.6033
950 C 18.382 46.683 27.2 31.6 2.1335


Appendix C-2

237

C-2 Dimensionless Plots for Chars C, C2 and D2.


Figure C-1: Dimensionless plots of conversion versus reduced time for chars C and C2.


0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Reduced time, t/t
0.9
(-)
Char C
Model 900 C, 100% CO 917 C, 100% CO 933 C, 100% CO
950 C, 100% CO 900 C, 75% CO 917 C, 75% CO 933 C, 75% CO
950 C, 75% CO 900 C, 50% CO 917 C, 50% CO 933 C, 50% CO
950 C, 50% CO 900 C, 25% CO 917 C, 25% CO 933 C, 25% CO
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Reduced time, t/t
0.9
(-)
Char C2
Model 900 C, 100% CO 917 C, 100% CO 933 C, 100% CO
950 C, 100% CO 900 C, 75% CO 917 C, 75% CO 933 C, 75% CO
950 C, 75% CO 900 C, 50% CO 917 C, 50% CO 950 C, 50% CO
900 C, 25% CO 917 C, 25% CO 933 C, 25% CO 950 C, 25% CO
Appendix C-3

238


Figure C-2: Dimensionless plot of conversion versus reduced time for char D2.

C-3 Comparison of Experimental and Model Gasification Results for
Chars C, C2 and D2

Figure C-3: Comparison between the experimental and model gasification results of char
C.
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Reduced time, t/t
0.9
(-)
Char D2
Model 900 C, 100% CO 917 C, 100% CO 933 C, 100% CO
950 C, 100% CO 900 C, 75% CO 917 C, 75% CO 933 C, 75% CO
950 C, 75% CO 900 C, 50% CO 917 C, 50% CO 933 C, 50% CO
950 C, 50% CO 900 C, 25% CO 917 C, 25% CO 933 C, 25% CO
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
M
o
d
e
l

X
,

t
i
m
e

f
a
c
t
o
r
,


(
t
f
)

(
m
i
n
-
1
)
Experimetal X, initial Reactivity, R (min
-1
)
Char C
900 C, 100% CO 917 C, 100% CO 933 C, 100% CO 950 C, 100% CO
900 C, 75% CO 917 C, 75% CO 933 C, 75% CO 950 C, 75% CO
900 C, 50% CO 917 C, 50% CO 933 C, 50% CO 950 C, 50% CO
900 C, 25% CO 917 C, 25% CO 933 C, 25% CO 950 C, 25% CO
Appendix C-3

239



Figure C-4: Comparison between the experimental and the model gasification results of
chars C2 and D2.


0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
M
o
d
e
l

X
,

t
i
m
e

f
a
c
t
o
r
,


(
t
f
)

(
m
i
n
-
1
)
Experimental X, initial reactivity R, (min
-1
)
Char C2
900 C, 100% CO 917 C, 100% CO 933 C, 100% CO 950 C, 100% CO
900 C, 75% CO 917 C, 75% CO 933 C, 75% CO 950 C, 75% CO
900 C, 50% CO 917 C, 50% CO 933 C, 50% CO 950 C, 50% CO
900 C, 25% CO 917 C, 25% CO 933 C, 25% CO 950 C, 25% CO
Linear trend
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
M
o
d
e
l

X
,

t
i
m
e

f
a
c
t
o
r
,

(
t
f
)

(
m
i
n
-
1
)
Experimental X, initial reactivity R, (min
-1
)
Char D2
900 C, 100% CO 917 C, 100% CO 933 C, 100% CO 950 C, 100% CO
900 C, 75% CO 917 C, 75% CO 933 C, 75% CO 950 C, 75% CO
900 C, 50% CO 917 C, 50% CO 933 C, 50% CO 950 C, 50% CO
900 C, 25% CO 917 C, 25% CO 933 C, 25% CO 950 C, 25% CO
Linear Trend
Appendix D-1

240

APPENDIX D

Model Validation: Random Pore Model (RPM)
D-1 RPM fitting to the Experimental Data of Chars B, C, C2 and D2.
In the following figures the solid lines represent the model predicted conversions,
while the points are the experimental conversions.



Figure D-1: RPM fitting of the experimental results of chars B and C at 0.875 bar.

0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200 1400
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 900 C, 100% CO
Char B, 917 C, 100% CO
Char B, 933 C, 100% CO
Char B, 950 C, 100% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000 1200 1400 1600
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 900 C, 75% CO
Char B, 917 C, 75% CO
Char B, 933 C, 75% CO
Char B, 950 C, 75% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 400 800 1200 1600 2000 2400
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char B, 900 C, 25% CO
Char B, 917 C, 25% CO
Char B, 933 C, 25% CO
Char B, 950 C, 25% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 150 300 450 600 750 900 1050
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 900 C, 100% CO
Char C, 917 C, 100% CO
Char C, 933 C, 100% CO
Char C, 950 C, 100% CO
RPM
Appendix D-1

241


Figure D-2: RPM fitting of the experimental results of chars C, C2 and D2 at 0.875 bar.



0
0.2
0.4
0.6
0.8
1
0 300 600 900 1200 1500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 900 C, 50% CO
Char C, 917 C, 50% CO
Char C, 933 C, 50% CO
Char C, 950 C, 50% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 450 900 1350 1800 2250 2700
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C, 900 C, 25% CO
Char C, 917 C, 25% CO
Char C, 933 C, 25% CO
Char C, 950 C, 25% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 200 400 600 800 1000
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 900 C, 100% CO
Char C2, 917 C, 100% CO
Char C2, 933 C, 100% CO
Char C2, 950 C, 100% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 250 500 750 1000 1250
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 900 C, 75% CO
Char C2, 917 C, 75% CO
Char C2, 933 C, 75% CO
Char C2, 950 C, 75% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 250 500 750 1000 1250 1500 1750
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char C2, 900 C, 50% CO
Char C2, 917 C, 50% CO
Char C2, 933 C, 50% CO
Char C2, 950 C, 50% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 50 100 150 200 250 300 350
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 900 C, 75% CO
Char D2, 917 C, 75% CO
Char D2, 933 C, 75% CO
Char D2, 950 C, 75% CO
RPM
Appendix D-2

242


Figure D-3: RPM fitting of the experimental results of char D2 at 0.875 bar.





D-2 RPM Fitting of Char Conversion Rate to Experimental Results for
the Chars


Figure D-4: RPM fitting of the char conversion rates for char B at 25% and 50% CO
2

concentration and different temperatures.

0
0.2
0.4
0.6
0.8
1
0 100 200 300 400 500
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 900 C, 50% CO
Char D2, 917 C, 50% CO
Char D2, 933 C, 50% CO
Char D2, 950 C, 50% CO
RPM
0
0.2
0.4
0.6
0.8
1
0 100 200 300 400 500 600
C
o
n
v
e
r
s
i
o
n
,

X
(
-
)
Time, t (min)
Char D2, 900 C, 25% CO
Char D2, 917 C, 25% CO
Char D2, 933 C, 25% CO
Char D2, 950 C, 25% CO
RPM
0
0.002
0.004
0.006
0.008
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char B: =1.11
950 C, 50% CO
933 C, 50% CO
917 C, 50% CO
900 C, 50% CO
RPM
0
0.002
0.004
0.006
0.008
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char B: =1.11
950 C, 75% CO
933 C, 75% CO
917 C, 75% CO
900 C, 75% CO
RPM
Appendix D-2

243


Figure D-5: RPM fitting of the char conversion rates for chars B, C, C2 at different
experimental conditions, 0875 bar.

0.000
0.002
0.004
0.006
0.008
0.010
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)

Conversion, X (-)
Char B: =1.11
950 C, 100% CO
933 C, 100% CO
917 C, 100% CO
900 C, 100% CO
RPM
0
0.001
0.002
0.003
0.004
0.005
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C: =1.75
950 C, 50% CO
933 C, 50% CO
917 C, 50% CO
900 C, 50% CO
RPM
0
0.002
0.004
0.006
0.008
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C: =1.75
950 C, 75% CO
933 C, 75% CO
917 C, 75% CO
900 C, 75% CO
RPM
0
0.002
0.004
0.006
0.008
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C: =1.75
950 C, 100% CO
933 C, 100% CO
917 C, 100% CO
900 C, 100% CO
RPM
0.000
0.001
0.002
0.003
0.004
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C2: =2.58
950 C, 50% CO
933 C, 50% CO
917 C, 50% CO
900 C, 50% CO
RPM
0
0.001
0.002
0.003
0.004
0.005
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C2: =2.58
950 C, 75% CO
933 C, 75% CO
917 C, 75% CO
900 C, 75% CO
RPM
Appendix D-2

244


Figure D-6: RPM fitting of the char conversion rates for chars C and D2 at different
experimental conditions, 0.875 bar.


0
0.001
0.002
0.003
0.004
0.005
0.006
0.007
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char C2: =2.58
950 C, 100% CO
933 C, 100% CO
917 C, 100% CO
900 C, 100% CO
RPM
0.000
0.005
0.010
0.015
0.020
0.025
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char D2: =1.57
950 C, 50% CO
933 C, 50% CO
917 C, 50% CO
900 C, 50% CO
RPM
0.000
0.005
0.010
0.015
0.020
0.025
0.030
0 0.2 0.4 0.6 0.8 1
R
a
t
e

o
f

r
e
a
c
t
i
o
n
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char D2: =1.57
950 C, 75% CO
933 C, 75% CO
917 C, 75% CO
900 C, 75% CO
RPM
0.000
0.006
0.012
0.018
0.024
0.030
0.036
0 0.2 0.4 0.6 0.8 1
R
e
a
c
t
i
o
n

r
a
t
e

c
h
a
n
g
e
,

d
X
/
d
t

(
m
i
n
-
1
)
Conversion, X (-)
Char D2: = 1.57
950 C, 100% CO
933 C, 100% CO
917 C, 100% CO
900 C, 100% CO
RPM

Das könnte Ihnen auch gefallen