You are on page 1of 289

ELECTRICAL RESISTIVITY MEASUREMENTS IN

COAL: ASSESSMENT OF COAL-BED METHANE

CONTENT, RESERVES AND COAL

PERMEABILITY

Thesis submitted for the degree of

Doctor of Philosophy

at the University of Leicester

by

Joo Mealha Sequeira Afonso, MGeol.

Department of Geology

University of Leicester

2014
ABSTRACT

Electrical Resistivity Measurements in Coal: assessment of coal-bed


methane content, reserves and coal permeability
by
Joo Mealha Sequeira Afonso

Coal-bed methane (CBM), also referred to as Coal seam gas (CSG), relates to the
production of methane from coal beds by drilling wells, hence lowering formation
pressure, and triggering methane release. While the potential of this resource is
significant, the assessment of the quantity and the producibility of methane from coal
seams is highly variable. For this reason the objective of this work is to investigate the
assessment of gas content, gas-in-place and coal permeability through petrophysical of
analysis and by gaining a better understanding of coal bulk properties,.
In this study 17 cored production wells were analysed from the Walloon Sub-group
coal seams fairway in the Surat Basin in Queensland Australia, which is today the most
ambitious investment in CSG worldwide. A total of 2374 coal beds were investigated to
understand how the nature of the different coal lithotypes are reflected in core analysis,
wireline logs measurements and DST test results, and how they affect coal quality, and
control gas content, fracture development and reservoir permeability.
High-resolution studies involving fine scale are required to estimate volumes and
CSG formation evaluation turns to the interpretation of standard wireline tools readings
in hundreds of coal seam wells. Nevertheless, the heterogeneous thin-bedded nature of
coal seams, together with the fact that methane within coal is mainly stored by
adsorption, create several difficulties in wireline log petrophysical analysis.
Consequently core description is used to validate the combination of the density log
with the shallow focused electric and induction resistivity measurements, benefitting
the recognition and thickness estimation of thin coal beds and coal laminae rich
mudstones. This observation, and a refined coal quality and gas content estimation
methodology, are presented and tested against previously published workflows and
provide an improved and tested strategy for petrophysical analysis of CSG.

ii
ACKNOWLEDGEMENTS
I would like to thank my supervisors Professor Mike Lovell, Professor Sarah Davies,

Mr. Roger Samworth and Dr. Darren Chaney for their guidance, support, patience,

insightful thoughts in different fields of expertise and giving me the opportunity to do

this PhD research project. I would like also to thank Weatherford UK, for financially

supporting this project and the East Leake team in particular for their kindness and

advice.

I would like to express my appreciation to Mr. Brent Glassborow, for his

fundamental assistance in compiling the dataset used in this project and knowledge

shared of coal seam gas operations. To Mr. Tim Pritchard, for giving me the placement

opportunity in BGs TVP office and Dr. Stefan Calvert for the opportunity to visit and

discuss my work in BGs Brisbane office.

Would like to show my appreciation for the discussions, advices and help throughout

the different stages of the project from my colleagues Sam Matthews and Sven

Koenitzer.

I would like send warm welcome to my friends Louise Anderson, Sally Morgan and

Dave Hartigan for helping me improving my manuscripts, but especially for their

friendship, encouragement and good times shared.

Finally, my biggest thanks to my family, my mother, father, sister and my

grandmother Sameirinho whose amor and unconditional support I could not be without.

iii
LIST OF CONTENTS
ABSTRACT .................................................................................................................ii
ACKNOWLEDGEMENTS ....................................................................................... iii
LIST OF CONTENTS ................................................................................................ iv
LIST OF TABLES .....................................................................................................vii
LIST OF FIGURES .................................................................................................... ix
CHAPTER 1 Introduction ......................................................................................... 1
1.1. Preamble ......................................................................................................... 1
1.2. Structure and Aims of the thesis ..................................................................... 3
CHAPTER 2 CSG Geological and technical background ........................................ 7
2.1. Coal Seam Gas ............................................................................................... 7
2.2. From Peat to Coal ........................................................................................... 9
2.3. Coal Petrographical Characteristics ............................................................. 12
2.3.1. Lithotypes .............................................................................................. 12
2.3.2. Macerals ................................................................................................ 15
2.3.3. Cleat system & Permeability................................................................. 18
2.4. Coal as a Gas reservoir ................................................................................. 22
2.4.1. Coal gas origin ...................................................................................... 22
2.4.2. Gas Storage ........................................................................................... 23
2.4.3. Controls on gas content ......................................................................... 24
2.4.4. Coal Gas Saturation............................................................................... 30
2.4.5. Gas flow ................................................................................................ 33
2.5. CSG reserve & production ........................................................................... 36
2.5.1. Gas-in-place estimation ......................................................................... 36
2.5.2. Production ............................................................................................. 39
2.6. CSG formation evaluation ............................................................................ 43
2.6.1. Core analysis ......................................................................................... 44
2.6.2. Downhole wireline measurements ........................................................ 55
2.6.3. Drill stem tests (DST) ........................................................................... 62
2.7. Study area ..................................................................................................... 64
2.8. Studied Dataset ............................................................................................. 71
CHAPTER 3 Coal lithotype properties of the Walloon Sub-group, in Queensland,
Australia........................... ............................................................................................... 75
3.1. Introduction .................................................................................................. 75
3.2. Methodology................................................................................................. 76
3.2.1. Original data set .................................................................................... 76
3.2.2. Macroscopic description ....................................................................... 77
3.2.3. Sample selection.................................................................................... 78
3.3. Coal bed Thickness....................................................................................... 79
3.4. Lithotype distribution ................................................................................... 82
3.5. Petrography and Geochemistry .................................................................... 83
3.5.1. Rank ...................................................................................................... 83
3.5.2. Proximate Analysis ............................................................................... 85
3.5.3. Petrographical Analysis ........................................................................ 86
3.6. Cleat system.................................................................................................. 87
3.6.1. Lithotype cleat attributes ....................................................................... 88
3.7. Methane producibility .................................................................................. 91
3.7.1. Gas content and saturation .................................................................... 91
3.7.2. Coal lithotype and composition effect .................................................. 96
3.8. Wireline tool responses .............................................................................. 102
3.8.1. Standard logging reading .................................................................... 102
3.8.2. Imager and scanner readings ............................................................... 110
3.8.3. Coal lithotype and permeability effect on wireline measurements ..... 113
3.8.4. Reservoir permeability ........................................................................ 114
3.9. Discussion and Conclusions ....................................................................... 118
CHAPTER 4 Estimating coal thickness in Thinly Bedded Coal Seam Gas
(CSG).............................. .............................................................................................. 121
4.1. Introduction ................................................................................................ 121
4.2. Core and wireline data correlation ............................................................. 123
4.3. Coal Properties ........................................................................................... 125
4.4. Wireline logging for Coal Seam Gas.......................................................... 126
4.4.1. Net coal estimation using density log ................................................. 129
4.5. Coal borehole resistivity ............................................................................. 134
4.5.1. Standard downhole resistivity measurements ..................................... 134
4.5.2. MFE Response to coal thickness......................................................... 136
4.5.3. Effect of permeability on resistivity response ..................................... 139
4.6. Net coal thickness estimation in washout zones......................................... 142
4.7. Using Resistivity in net coal estimation ..................................................... 146

v
4.7.1. Results in cored wells.......................................................................... 148
4.7.2. Results in production test wells........................................................... 153
4.8. Conclusions ................................................................................................ 157
CHAPTER 5 Evaluation of in situ Coal Seam Gas content using wireline logs .. 158
5.1. Introduction ................................................................................................ 158
5.2. Estimation of Coal quality .......................................................................... 159
5.2.1. Density of organic content .................................................................. 160
5.2.2. Density of inorganic content ............................................................... 165
5.2.3. Wireline density readings .................................................................... 167
5.3. Gas content and saturation.......................................................................... 168
5.4. Testing gas content estimation methodologies ........................................... 171
5.4.1. Mavor and Nelson (1997) ................................................................... 175
5.4.2. Rogers (2007) ...................................................................................... 177
5.4.3. Calvert et al. (2011)............................................................................. 179
5.4.4. New methods tested ............................................................................ 180
5.5. Conclusions ................................................................................................ 208
CHAPTER 6 Conclusions ..................................................................................... 209
6.1. Conclusions to initial questions posed ....................................................... 209
6.2. General suggestions for further work ......................................................... 216
APPENDIX A Seam and net coal thickness ......................................................... 218
APPENDIX B proximate and desorption analysis ............................................... 219
APPENDIX C wireline density to organic content x-plots................................... 221
APPENDIX D wireline responses ........................................................................ 223
APPENDIX E Permeability and coal bed characteristics ..................................... 244
APPENDIX F organic and inorganic content density estimation ......................... 245
APPENDIX G Error estimation ............................................................................ 247
APPENDIX H density and gas content equations ................................................ 250
APPENDIX I tools and logs acronyms ................................................................. 256
REFERENCES......................................................................................................... 257

vi
LIST OF TABLES
TABLE 2.1 MODIFIED TABLE FROM GAURAV ET AL. ( 2012) AND EHRENBERG AND NADEAU (2005)........... 9

TABLE 2.2 MACROSCOPIC DESCRIPTION OF COALS. ................................................................................... 13

TABLE 2.3 MACERAL ANALYSIS CLASSIFICATION FOLLOWING THE AUSTRALIAN STANDARDS (AS 2856.1-

2000, AS 2856.2-1998, AS 2856.3-2000, AS 2856-1986). ............................................................... 18

TABLE 2.4 ASTM RANK CLASSIFICATION OF COAL APPROXIMATE BOUNDARY VALUES FROM STACH ET AL.

(1982) AND WARD (1981). ................................................................................................................ 51

TABLE 2.5 NUMBER OF CORE SAMPLES PER COAL SEAM FROM EACH WELL. .............................................. 73

TABLE 2.6 SUMMARY OF THE CORE ANALYSIS AND DST TEST RESULTS DATASET FROM EACH WELL. ....... 73

TABLE 2.7 SUMMARY WIRELINE LOG DATA USED IN THIS STUDY. .............................................................. 74

TABLE 3.1 MACROSCOPIC DESCRIPTION OF COALS BASED ON THE AUSTRALIAN STANDARD. .................... 78

TABLE 3.2 COAL BED THICKNESS FOUND IN THE 10 CORED WELL IN THE WALLOON SUB-GROUP. ............. 80

TABLE 3.3 VITRINITE REFLECTANCE IN 12 WELLS IN THE WALLOON SUB-GROUP COAL SEAMS PROVIDED.

.......................................................................................................................................................... 84

TABLE 3.4 PROXIMATE ANALYSIS DATA FROM SELECTED COAL SAMPLES DOMINATED BY ONE SINGLE

LITHOTYPE AND < 50% ASH YIELD FROM WALLOON SUB-GROUP COAL BEDS. .................................. 85

TABLE 3.5 MACERAL ANALYSIS DATA FROM SELECTED COAL SAMPLES DOMINATED BY ONE SINGLE

LITHOTYPE FROM WALLOON SUB-GROUP COAL BEDS. ...................................................................... 87

TABLE 3.6 VARIATION OF LITHOTYPE CLEAT MEAN FREQUENCY, LENGTH AND FILLING FROM 2374

WALLOON SUB-GROUP COAL BEDS. .................................................................................................. 89

TABLE 3.7 TOTAL DESORBED GAS (D.A.F.) AND GAS SATURATION FROM 12 CORED WELLS........................ 94

TABLES 3.8 MACERAL PETROGRAPHIC COMPOSITION OF THE WALLOON SUB-GROUP (IN PERCENTAGE

MINERAL MATER FREE BASIS). ........................................................................................................... 99

TABLE 3.9 LITHOTYPE MEAN LOG RESPONSES ......................................................................................... 115

TABLE 3.10 DRILL STEM TEST RESULTS COMPARED TO COAL BED CHARACTERISTICS IN WELLS 1 TO 8 AND

10 TO 14. ......................................................................................................................................... 118

TABLE 4.1 TOTAL WELL DEPTH AND WALLOONS SUB-GROUP DEPTH INTERVALS IDENTIFIED IN CORE AND

WIRELINE LOGS. FROM CORED WELLS 1, 2 AND 3 SELECTED SAMPLES WERE TESTED FOR SPECIAL

CORE ANALYSIS. .............................................................................................................................. 124

vii
TABLE 4.2 COAL FACIES IDENTIFIED IN THE CORE DESCRIPTION .............................................................. 126

TABLE 4.3 EXAMPLES OF ACCURACY ESTIMATION OF 80% (IN WELL 1) AND 90% (WELLS 2 AND 3) OF

TOTAL RESERVOIR THICKNESS USING DIFFERENT CUT-OFFS. ........................................................... 149

TABLE 5.1 CORE MACERAL GROUP ANALYSIS MINERAL MATTER FREE RESULTS FROM 12 CORED WELLS (IN

PERCENTAGE MINERAL MATER FREE BASIS).. .................................................................................. 161

TABLE 5.2 ESTIMATED CORE ORGANIC MATTER DENSITY IN EACH COAL SEAM UNIT IN 490 SAMPLES FROM

12 WELLS. FOR MEAN ABSOLUTE ERROR OF THE ESTIMATE AND SAMPLE COUNT SEE APPENDIX G

ERROR ESTIMATION. ....................................................................................................................... 164

TABLE 5.3 FIXED CARBON TO VOLATILE MATTER RATIO IN EACH COAL SEAM UNIT IN 490 SAMPLES FROM

12 WELLS. FOR STANDARD DEVIATION OF THE ESTIMATE AND SAMPLE COUNT SEE APPENDIX G

ERROR ESTIMATION ........................................................................................................................ 164

TABLE 5.4 EXAMPLES OF THE ESTIMATED ASH DENSITY AND MEAN ABSOLUTE ERROR OF THE ESTIMATE

FROM 57 CORE SAMPLES IN EACH COAL SEAM UNIT IN WELLS 2 AND 3. ........................................... 167

TABLE 5.5 ESTIMATED GAS SATURATION FROM CORE DATA FROM 452 FROM 11 WELLS. FOR STANDARD

DEVIATION OF THE ESTIMATE AND SAMPLE COUNT SEE APPENDIX G ERROR ESTIMATION. ....... 169

TABLE 5.6 EQUILIBRIUM OF MOISTURE FROM ADSORPTION LANGMUIR ISOTHERM ANALYSIS CORE DATA.

........................................................................................................................................................ 170

TABLE 5.7 ESTIMATED MEAN ABSOLUTE PERCENTAGE ERROR FOR EACH METHOD FOR COAL QUALITY,

TOTAL GAS CONTENT AND COAL GAS CONTENT FOR WELL 2. .......................................................... 186

TABLE 5.8 ESTIMATED MEAN ABSOLUTE PERCENTAGE ERROR FOR EACH METHOD FOR COAL QUALITY,

TOTAL GAS CONTENT AND COAL GAS CONTENT FOR WELL 3. .......................................................... 186

TABLE 5.9 398.8 METERS OF CORE DESCRIPTION AGAINST THE RESULTS OF RESERVOIR ROCK CUT-OFF FOR

EACH METHODOLOGY IN WELL 2. .................................................................................................... 188

TABLE 5.10 CORE DESCRIPTION AGAINST THE RESULTS OF RESERVOIR ROCK CUT-OFF FOR EACH

METHODOLOGY IN WELL 3. ............................................................................................................. 188

TABLE 5.11 RESUME OF THE DIFFERENT COAL SEAM GAS WELL INTERPRETATION WORKFLOWS. ............ 205

TABLE 5.12 ROGERSS, MAVORS, CALVERTS AND THE NEW WORKFLOW RESULTS FROM WELL 2. ....... 206

TABLE 5.13 ROGERSS, MAVORS, CALVERTS AND THE NEW WORKFLOW RESULTS FROM IN WELL 3. ... 206

viii
LIST OF FIGURES
FIGURE 2.1 WORLDWIDE COAL BED METHANE RESERVES............................................................................ 8

FIGURE 2.2 COALIFICATION STAGES DURING COAL GENERATION. ............................................................. 10

FIGURE 2.3 CHANGES IN THE CHEMICAL AND PHYSICAL PROPERTIES OF COAL WITH INCREASING RANK.

MODIFIED FROM TEICHMLLER AND TEICHMLLER (1968). ............................................................ 11

FIGURE 2.4 EXAMPLES OF SPECIAL CORE ANALYSIS TESTED SAMPLES DESCRIPTION FOLLOWED IN THIS

RESEARCH. PICTURES PROVIDED BY THE OPERATOR. ........................................................................ 14

FIGURE 2.5 COAL MATRIX AND NATURAL FRACTURE NETWORK................................................................ 20

FIGURE 2.6 COAL RANK AND METHANE GENERATION FROM AHMED ET AL. (2009). .................................. 23

FIGURE 2.7 ADSORPTION CURVES AGAINST PRESSURE AT A GIVEN RANK FROM KIM (1977). .................... 25

FIGURE 2.8 TREND IN GAS STORAGE CAPACITY AGAINST RANK (%, RO MAX) IN DIFFERENT COAL

LITHOTYPES FROM LAXMINARAYANA AND CROSDALE (1999). ......................................................... 25

FIGURE 2.9 ADSORPTION CURVES AGAINST TEMPERATURE AT A GIVEN RANK FROM KIM (1977). ............. 30

FIGURE 2.10 EXAMPLES OF VERTICAL PROFILES OF DESORBED GAS CONTENT. .......................................... 31

FIGURE 2.11 SCHEMATIC GAS GENERATION AGAINST GAS STORAGE CAPACITY RELATED WITH DEPTH AND

TIME FROM MOORE (2012). ............................................................................................................... 32

FIGURE 2.12 METHANE FLOW MODEL THROUGH COAL FROM (GAMSON ET AL., 1996). ............................. 34

FIGURE 2.13 ESTIMATES OF PERCENT VOLUME DECREASE VERSUS RANK AND TEMPERATURE FROM RYAN

(2003). .............................................................................................................................................. 36

FIGURE 2.14 CSG RESERVOIR HETEROGENEITY......................................................................................... 38

FIGURE 2.15 CONCEPTUAL FREE GAS AND WATER FLOW MODELS. ............................................................ 40

FIGURE 2.16 TYPICAL COAL SEAM GAS AND WATER PRODUCTION CHARACTERISTICS FROM ANDERSON ET

AL.(2003). ......................................................................................................................................... 42

FIGURE 2.17 APPARATUS REQUIRED FOR DESORBED GAS (Q2) ESTIMATION. ............................................. 47

FIGURE 2.18 THERMOGRAVIMETRIC ANALYZER USED FOR THE PROXIMATE ANALYSIS ESTIMATION. ....... 50

FIGURE 2.19 RETENTION OF GAS IN COAL SEAMS REDRAWN FROM SUREZ-RUIZ AND CRELLING (2008). 53

FIGURE 2.20 EXAMPLE OF ADSORPTION PERFORMANCES OF DIFFERENT GASES PRODUCED FROM LOW

BITUMINOUS COAL SAMPLE FROM WESTERN CANADA FROM SUREZ-RUIZ AND CRELLING (2008). . 54

ix
FIGURE 2.21 TYPICAL TOOL READINGS FROM WIRELINE LOG MEASUREMENTS AGAINST CORE DESCRIPTION

IN COAL SEAM GAS EXPLORATION. EXAMPLE FROM A COAL SEAM IN WELL 3. .................................. 61

FIGURE 2.22 LOCATION OF THE WELL DATA IN THE SURAT BASIN IN QUEENSLAND AUSTRALIA. ............ 66

FIGURE 2.23 WELL LOCATION IN WALLOON SUBGROUP DEPTH (MSS) STRUCTURE MAP. MODIFIED FROM

QGC PTY LIMITED (2012A) REPORT. ................................................................................................ 67

FIGURE 2.24 MAJOR TECTONIC ELEMENTS OF THE SURAT BASIN AND WELL LOCATION. MODIFIED FROM

QGC PTY LIMITED (2012A) REPORT. ................................................................................................ 68

FIGURE 2.25 SEISMIC SECTION LOCATED AS INDICATED IN FIGURE X. FROM (QGC PTY LIMITED, 2012B) 69

FIGURE 2.26 LITHO-STRATIGRAPHY OF THE WALLOON SUB-GROUP FROM (MARTIN ET AL., 2013)........... 70

FIGURE 2.27 QUEENSLAND AUSTRALIA PROVED AND PROBABLE RESERVES ESTIMATED IN 2013 (SOURCE

QUEENSLAND GOVERNMENT, DEPARTMENT OF NATURAL RESOURCES AND MINES). ...................... 71

FIGURE 3.1 COAL LITHOTYPE THICKNESS FREQUENCY AND ACCUMULATED FREQUENCY.......................... 80

FIGURE 3.2 LITHOTYPE THICKNESS HISTOGRAMS DISTRIBUTIONS FROM THE DESCRIPTION OF CORE FROM

10 CORED WELLS. .............................................................................................................................. 81

FIGURE.3.3 VARIATION OF LITHOTYPE THICKNESS WITH DEPTH FOR WALLOON SUB-GROUP. ................... 83

FIGURE 3.4 MEAN VITRINITE REFLECTANCE IN 12 WELL IN THE WALLOON SUB-GROUP COAL SEAMS. ..... 84

FIGURE 3.5 TRENDS BETWEEN BED THICKNESS AND CLEAT FREQUENCY IN EACH LITHOTYPE FROM 2374

WALLOON SUB-GROUP COAL BEDS. .................................................................................................. 89

FIGURE 3.6 CROSS-PLOT OF DEPTH VERSUS MEAN CLEATS PER METRE FROM 2374 COAL BEDS IN WALLOON

SUB-GROUP COAL SEAMS. ................................................................................................................. 90

FIGURE 3.7 MEAN PERCENTAGE OF LOST, DESORBED AND RESIDUAL OF THE TOTAL GAS YIELD IN ALL 491

SAMPLES. .......................................................................................................................................... 92

FIGURE 3.8 GAS CONTENT (D.A.F.) VERSUS DEPTH. ................................................................................... 92

FIGURE 3.9 MEAN TOTAL DESORBED GAS VOLUME AND GAS SATURATION FROM 12 CORED WELLS. ......... 95

FIGURE 3.10 WELL 2 EXAMPLE OF VARIATION OF ESTIMATED SATURATION ON A SINGLE COAL SEAM. ..... 95

FIGURE 3.11 MEASURED (RAW) GAS CONTENT VERSUS ORGANIC CONTENT MEASURED IN 491 CORE

SAMPLES FROM 12 CORED WELLS. ..................................................................................................... 97

FIGURE 3.12 MEASURED (RAW) GAS CONTENT VERSUS LIPTINITE CONTENT IN 454 SAMPLES FROM 12

CORED WELLS. ................................................................................................................................ 100

x
FIGURE 3.13 MEASURED (RAW) GAS CONTENT VERSUS VITRINITE CONTENT IN 454 SAMPLES FROM 12

CORED WELLS. ................................................................................................................................ 101

FIGURE 3.14 MEAN CLEAT DEVELOPMENT AGAINST MEAN MACERAL CONTENT IN THE WALLOON SUB-

GROUP. ............................................................................................................................................ 101

FIGURE 3.15 MACERAL PETROGRAPHIC COMPOSITION OF THE MOST COMMON LITHOTYPES OF THE

WALLOON SUB-GROUP. .................................................................................................................. 102

FIGURE 3.16 EXAMPLE OF WIRELINE LOGS SUPPLIED TO THE OPERATOR FROM WELL 3. .......................... 103

FIGURE 3.17 AN EXAMPLE OF WIRELINE DATA INTERPRETATION IN WELL 3 USING UNFILTERED DENSITY,

PE, AND FEFE LOGS. ....................................................................................................................... 104

FIGURE 3.18 WIRELINE DENSITY RESPONSES FOUND IN COAL BED ZONES IN 10 CORED WELLS................ 106

FIGURE 3.19 WIRELINE NEUTRON-POROSITY RESPONSES FOUND IN COAL BED ZONES IN 10 CORED WELLS.

........................................................................................................................................................ 106

FIGURE 3.20 WIRELINE NATURAL GAMMA-RAY RESPONSES FOUND IN COAL BED ZONES IN 10 CORED

WELLS. ............................................................................................................................................ 107

FIGURE 3.21 WIRELINE PHOTOELECTRIC RESPONSES FOUND IN COAL BED ZONES IN 10 CORED WELLS. .. 107

FIGURE 3.22 WIRELINE SONIC LOG RESPONSES FOUND IN COAL BED ZONES IN 10 CORED WELLS. ........... 108

FIGURE 3.23 WIRELINE INDUCTION RESISTIVITY RESPONSES FOUND IN 3 CORED WELLS IN COAL BED

ZONES. ............................................................................................................................................ 108

FIGURE 3.24 WIRELINE SHALLOW FOCUSED ELECTRIC RESISTIVITY RESPONSES FOUND IN 3 CORED WELLS

IN COAL BED ZONES. ........................................................................................................................ 109

FIGURE 3.25 WIRELINE LATERALOG RESISTIVITY RESPONSES FOUND IN 8 CORED WELLS IN COAL BED

ZONES. ............................................................................................................................................ 109

FIGURE 3.26 EXAMPLE TAKEN FROM WELL 2 OF IMAGER AND SCANNER RESPONSES IN COAL. ................ 111

FIGURE 3.27 (LEFT) EXAMPLE OF A SCANNING ELECTRON MICROSCOPE PICTURE SHOWING A CLEAT WITH

CALCITE FILLING. ............................................................................................................................ 112

FIGURE 3.28 (RIGHT) COAL HAND SPECIMEN SHOWING FACE AND BUTT CLEATS. ................................... 112

FIGURE 4.1 CORRELATION BETWEEN WELL 1, 2 AND 3 BASED ON CORE DESCRIPTION. FOR LOCATION SEE

FIGURE 2.22 IN CHAPTER 2. ............................................................................................................ 124

FIGURE 4.2 AN EXAMPLE OF WIRELINE DATA CORRELATED WITH CORE ANALYSIS FROM WELL 2. .......... 129

xi
FIGURE 4.3 COAL ASH AND ORGANIC CONTENT AGAINST CORE DENSITY CROSS PLOT OF WELLS 1, 2 AND 3.

........................................................................................................................................................ 130

FIGURE 4.4 WIRELINE AGAINST CORE DENSITY WITH COLOUR DISTRIBUTION OF BOREHOLE ENLARGEMENT

OF WELLS 1, 2 AND 3. ...................................................................................................................... 132

FIGURE 4.5 STUDY OF DENSITY LOG CUT-OFF VALUE. ............................................................................. 134

FIGURE 4.6 SHALLOW FOCUSED ELECTRIC RESISTIVITY AGAINST ORGANIC CONTENT ESTIMATED FROM

PROXIMATE ANALYSIS AS A FUNCTION OF COAL SEAM THICKNESS. ................................................. 137

FIGURE 4.7 AN EXAMPLE OF WIRELINE DATA CORRELATED WITH CORE ANALYSIS FROM WELL 3. .......... 138

FIGURE 4.8 CLEAT DEVELOPMENT AGAINST DEPTH IN ALL THE COAL BED FOUND IN WELL 2................ 141

FIGURE 4.9 HIGH-RESOLUTION WIRELINE FEFE SHALLOW RESISTIVITY AGAINST HIGH-RESOLUTION

WIRELINE DENSITY SHOWING COLOURED VARIATION ON CLEAT DEVELOPMENT IN EACH COAL BED.

........................................................................................................................................................ 142

FIGURE 4.10 AN EXAMPLE OF WIRELINE DATA CORRELATED WITH CORE ANALYSIS FROM WELL 1. ........ 144

FIGURE 4.11 STUDY OF GAMMA RAY LOG CUT-OFF VALUE. ..................................................................... 145

FIGURE 4.12 STUDY OF NEUTRON POROSITY LOG CUT-OFF VALUE. ......................................................... 145

FIGURE 4.13 STUDY OF SONIC LOG CUT-OFF VALUE. ............................................................................... 146

FIGURE 4.14 STUDY OF WORKFLOW CUT-OFF VALUE. ............................................................................. 149

FIGURE 4.15 COMPARISON BETWEEN THE ENTIRE RESERVOIR ROCK SEQUENCES IN WELL 1, 2 AND 3 FROM

CORE DESCRIPTION AGAINST THE ESTIMATED FROM 1.8 AND 2.0G/CM3 DENSITY CUT-OFF WITHOUT

AND WITH RESISTIVITY FILTERS. ..................................................................................................... 152

FIGURE 4.16 THE FEFE RESISTIVITY LOG AGAINST THE HIGH-RESOLUTION DENSITY LOG RESPONSES

THROUGHOUT THE WALLOON COAL MEASURES SEQUENCE IDENTIFIED IN CORE DESCRIPTION IN

WELL 3. ........................................................................................................................................... 155

FIGURE 4.17 THE FEFE RESISTIVITY LOG AGAINST THE HIGH-RESOLUTION DENSITY LOG WITH ESTIMATED

NET COAL THROUGHOUT THE WALLOON COAL MEASURES SEQUENCE BY 0.025M SAMPLE

INCREMENTS IN WELL 3. .................................................................................................................. 155

FIGURE 4.18 THE FEFE RESISTIVITY LOG AGAINST THE HIGH-RESOLUTION DENSITY LOG RESPONSES

THROUGHOUT THE WALLOON COAL MEASURES SEQUENCE BY 0.025M SAMPLE INCREMENTS IN

WELLS 3A AND 3B RESPECTIVELY. ................................................................................................. 156

xii
FIGURE 5.1 AN EXAMPLE OF THE EXTRAPOLATION AND MEAN ABSOLUTE ERROR OF THE ESTIMATE BARS

(WILLMOTT AND MATSUURA, 2005) OF DRY ASH AND INHERENT MOISTURE PERCENTAGE BY WEIGHT

AGAINST RECIPROCAL DENSITY ONE WELL IN ARGYLE AND CONDAMINE COAL SEAMS FROM 24

SAMPLES FROM WELL 14. ................................................................................................................ 163

FIGURE 5.2 DRY ORGANIC MATTER, MOISTURE AND ASH FROM PROXIMATE CORE ANALYSIS AGAINST

RECIPROCAL DENSITY FROM 491 CORE SAMPLES TAKEN FROM 12 CORED WELLS. ........................... 166

FIGURE 5.3 MOISTURE CONTENT ESTIMATED IN PROXIMATE ANALYSIS AGAINST DEPTH FROM 491 CORE

SAMPLES TAKEN FROM 12 CORED WELLS. ....................................................................................... 166

FIGURES 5.4 AND 5.5 RELATIONSHIP BETWEEN CORE RELATIVE DENSITY AND CORE PROXIMATE ANALYSIS

IN ALL SAMPLES FROM WELL 2 AND 3. ............................................................................................. 178

FIGURE 5.6 COMPARISON BETWEEN INORGANIC CONTENT FROM CORE PROXIMATE ANALYSIS AGAINST

MAVOR AND NELSON (1997) WIRELINE LOG ESTIMATION............................................................... 189

FIGURE 5.7 COMPARISON BETWEEN CORE PROXIMATE ANALYSIS AGAINST ROGERS (2007) WIRELINE LOG

ESTIMATION. ................................................................................................................................... 189

FIGURE 5.8 COMPARISON BETWEEN CORE PROXIMATE ANALYSIS AGAINST CALVERT ET AL. (2011)

WIRELINE LOG ESTIMATION. ............................................................................................................ 190

FIGURE 5.9 COMPARISON BETWEEN CORE PROXIMATE ANALYSIS AGAINST W METHOD WIRELINE LOG

ESTIMATION. ................................................................................................................................... 191

FIGURE 5.10 COMPARISON BETWEEN CORE PROXIMATE ANALYSIS AGAINST V METHOD WIRELINE LOG

ESTIMATION. ................................................................................................................................... 191

FIGURE 5.11 CORE RECIPROCAL CORE DENSITY AGAINST AS ANALYSED GAS CONTENT IN THE LOWER

JUANDAH. ....................................................................................................................................... 192

FIGURE 5.12 CORE RECIPROCAL DENSITY AGAINST AS ANALYSED GAS CONTENT AND PRESSURE IN THE

LOWER JUANDAH USED IN METHOD W FOR GAS ESTIMATION. ........................................................ 193

FIGURE 5.13 ORGANIC MATTER IN V/V AGAINST AS ANALYSED GAS CONTENT USED IN THE LOWER

JUANDAH IN METHOD V FOR GAS ESTIMATION. ............................................................................... 193

FIGURE 5.14 RECIPROCAL DENSITY AND PRESSURE AGAINST AS ANALYSED GAS CONTENT IN THE LOWER

JUANDAH USED IN METHOD WP FOR GAS ESTIMATION. ................................................................... 194

FIGURE 5.15 ORGANIC MATTER AND PRESSURE AGAINST AS ANALYSED GAS CONTENT (V/V) IN THE LOWER

JUANDAH USED IN METHOD VP FOR GAS ESTIMATION..................................................................... 194

xiii
FIGURE 5.16 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED TOTAL GAS CONTENT

ESTIMATED USING THE MAVOR AND NELSON (1997). ..................................................................... 195

FIGURE 5.17 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED TOTAL GAS CONTENT

ESTIMATED USING THE ROGERS (2007). .......................................................................................... 195

FIGURE 5.18 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED TOTAL GAS CONTENT

ESTIMATED USING THE CALVERT ET AL. (2011). ............................................................................. 196

FIGURE 5.19 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED TOTAL GAS CONTENT

ESTIMATED USING W METHOD. ....................................................................................................... 197

FIGURE 5.20 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED TOTAL GAS CONTENT

ESTIMATED USING V METHOD. ........................................................................................................ 197

FIGURE 5.21 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED TOTAL GAS CONTENT

ESTIMATED USING WP METHOD. ..................................................................................................... 198

FIGURE 5.22 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED TOTAL GAS CONTENT

ESTIMATED USING VP METHOD. ...................................................................................................... 198

FIGURE 5.23 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED COAL GAS CONTENT

ESTIMATED USING THE ROGERS (2007). .......................................................................................... 199

FIGURE 5.24 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED COAL GAS CONTENT

ESTIMATED USING THE CALVERT ET AL. (2011). ............................................................................. 199

FIGURE 5.25 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED COAL GAS CONTENT

ESTIMATED USING W METHOD. ....................................................................................................... 200

FIGURE 5.26 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED COAL GAS CONTENT

ESTIMATED USING V METHOD. ........................................................................................................ 200

FIGURE 5.27 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED COAL GAS CONTENT

ESTIMATED USING WP METHOD. ..................................................................................................... 201

FIGURE 5.28 PLOTS OF GAS CONTENT FROM CORE ANALYSIS AGAINST ESTIMATED COAL GAS CONTENT

ESTIMATED USING VP METHOD. ...................................................................................................... 201

FIGURE 5.29 AN EXAMPLE OF WIRELINE DATA ROGERSS, MAVORS AND CALVERTS INTERPRETATION IN

A MULTIPLE COAL BED BEARING ZONE IN WELL 3. ........................................................................... 202

FIGURE 5.30 AN EXAMPLE OF WIRELINE DATA INTERPRETATION IN A MULTIPLE COAL BED BEARING ZONE

IN WELL 3 USING W AND WP METHODS. ......................................................................................... 203

xiv
FIGURE 5.31 AN EXAMPLE OF WIRELINE DATA INTERPRETATION IN A MULTIPLE COAL BED BEARING ZONE

IN WELL 3 USING V AND VP METHODS. ........................................................................................... 204

FIGURE 5.32 BASIC TWO STEP WORKFLOW METHODOLOGY TO ESTIMATE NET COAL (AS DEBATED IN

CHAPTER 4) AND IN SITU GAS CONTENT WITH INPUT OF WIRELINE LOGS AND CORE AND CORE

ANALYSIS DATA. ............................................................................................................................. 207

xv
Chapter 2

CHAPTER 1 Introduction

1.1. Preamble
Unconventional gas reservoirs are expected to play a vital role in providing

worldwide secure sources of energy. In the last two decades the global gas market has

focussed on unconventional gas resources to fulfil a worldwide increasing energy

demand (Ahlbrandt, 2002; International Energy Agency, 2013). With experience gained

in the last few years, investigating the possible development of unconventional

reservoirs, such as shale gas or coal seam gas, promising projects have emerged around

the world. At the present time natural gas is the third largest global energy source.

Global exploration of natural gas continues to grow, fuelled by increasing consumption

and because it is considered to be a cleaner energy source compared to other fossil fuels

(Unconventional Energy Resources", 2011). Local activity is expected to increase but

depends on availability of the resource and knowledge of how these reservoirs behave.

Coalbed methane (CBM) or Coal seam gas (CSG) is a natural gas source formed by

geological and biological processes and associated with coal seams. CSG resources

typically lie in non-mined or unlikely-to-be-mined coal seams. Natural gas,

predominantly methane, is stored by sorption as opposed to compression as in

conventional reservoirs. Sorption provides a much greater storage capacity at shallow

depths, and accessible production from coal at shallow depth looks attractive in a global

market. Possible usage includes electric power generation, industrial processes and

industrial and domestic heating. Furthermore, CSG is considered a clean fuel because

its combustion releases low emissions of oxides of sulphur and nitrogen. Compared to

the combustion of coal or oil, it has limited carbon dioxide per unit of energy as a by-

1
Chapter 2

product, benefitting the environment with the reduction of greenhouse gas emissions to

the atmosphere (Ahmed et al., 2009; Moore, 2012).

Humans have a long history of exploring coal as a cheap accessible source of energy.

More recently, over the last 3 decades, coal has been targeted for its ability to generate

and sustain significant reserves of methane gas. Exploration has benefitted from the

accumulated knowledge and has adopted evaluation techniques from coal mining,

where for example the units used for the gas content in coal are still m3/tonne,

expressing the gas volume per mass of rock. Now coal seams that are not targeted for

mining purposes are being drilled for coal seam gas with the best examples found in the

United States (San Juan and Black Warrior Basins) and Australia (Surat and Bowen

Basins).

In coal mining wireline logging tools have been used to estimate thickness and coal

quality by knowing the physical properties of coal are typically low density, low natural

gamma ray emissions and highly electrical resistivity. In CSG, gas-in-place estimation

also relies on special core analysis, where in situ gas content and sorption capacity are

amongst the properties measured. Permeability is determined by production history

analysis or by pressure transient tests. Hence, to evaluate coal as a methane reservoir

the geologist needs to analyse these data and establish correlations between within it so

CSG reservoir development can be assessed (Mavor and Nelson, 1997; Rogers, 2007;

Seidle, 2011).

CSG projects always have challenging factors related to understanding coal as a

reservoir. Difficulties appear in calculating fundamental parameters including reservoir

area, thickness, coal permeability, in situ gas content analysis and ultimately in gas-in-

place estimation. This research project exposes the challenges of CSG formation

evaluation by studying coal and coal bed properties through downhole logs, core data

2
Chapter 2

and production test results. The project demonstrates that, while using techniques

already in practice, novel interpretations can still improve and optimise the

petrophysical analysis of CSG gas-in-place estimation.

1.2. Structure and Aims of the thesis

The overall aim of this thesis is to determine the best strategy for the petrophysical

analysis of coal and coal seam gas. This objective is reached by petrophysical analyses

of a dataset, which includes downhole logs, core data and drill stem test results, from

numerous coal beds of the Walloon Sub-group coal seams, towards a better

understanding of wireline logs, gas content estimation and methane producibility. The

dataset were kindly provided by BG group, but inevitably the range of data available

(provided) limits the direction and extent of the research presented here.

This thesis first describes the origin and the physical and chemical nature of coal, in

Chapter 2. Also in the same Chapter an overview of the measurements analysed and

the study area where the dataset was acquired is provided. After this introductory

chapter the following three chapters focus on testing scientific questions.

In Chapter 3 while trying to answer the questions below a dataset gathered from the

Walloon Sub-group coal beds was analysed. These coals are well accepted to be a part

of the non-marine Jurassic super sequence K of the Surat Basin, being made of laterally

discontinuous high-volatile sub-bituminous coal seams (Crosdale, 2004; Hoffmann et

al., 2009; Papendick et al., 2011; Scott et al., 2007).

Martin et al., (2013) has divided the Walloon Sub-group coal seams in terms of

thickness by dividing them into two major groups (Class I and II) indicating that studies

are still needed to understand coal type distribution and cleat development for a more

refined petrophysical and gas production potential understanding. Hamilton et al.,

3
Chapter 2

(2012) and Scott et al., (2007) show that, like other CSG Basins, there is no single

explanatory model that can indicate the gas content of the coals at a regional scale

beyond what is observed locally.

Coal is usually analysed as part of separate reservoirs or coal seams and not as

distinct lithotypes (Calvert et al., 2011; Kabir et al., 2011). There is a general lack of

published information not only regarding the nature of coal lithotypes found in the

Walloon Sub-group but also how they influence the petrophysical properties of the coal

seams and gas producibility in CSG.

Question 1:

What is the nature of the Walloon Sub-group coal beds regarding their

composition, thickness and fracture system?

Question 2:

How are the different coal lithotypes reflected in the petrophysical

properties?

A critical parameter to be estimated when drilling for CSG is coal thickness or net

coal. Due to the low mineral content and high organic nature of coal, the standard

approach used to quantify coal thickness is based on the low density of organic matter

compared to other minerals. This approach can be applied to the density wireline log

interpretation. Nevertheless the thin nature of the coal seams that CSG exploration most

frequently targets worldwide leads to difficulties defining what should be considered

net coal in this way, and is one of the major sources of error in estimating gas-in-place

analysis with consequent significant impact in productivity forecast (Lagendijk and

Ryan, 2010; Mavor and Nelson, 1997; Nelson, 1999; Zuber and Olszewski, 1992).

4
Chapter 2

Economic viability of CSG exploration relies on drilling a large number of wells

compared to conventional oil & gas resources because of the variability in coal seams

laterally. Formation evaluation of these wells is done by interpreting standard wireline

log measurements. In addition to the primary nuclear and acoustic tools, different types

of resistivity tools are also run. Several interpretations have been given to the different

resistivity log responses provided by these tools. Samworth and Cherrie (1976)

identified that the focused electric log response had a different character to the short

spaced density log, giving a possible explanation to the higher sensitivity to moisture of

the resistivity measuring tool. Hoyer, (1991) and Yang et al. (2006) have assumed that

the response character of the laterolog and the spherically focused logs is associated

with permeability and porosity found in cleats (fractures) in coal respectively. Due to its

poor vertical resolution Induction logs are often considered to be inadequate for thin

beds and coal seam gas (Colson, 1991; Mavor et al., 1990a; Samworth and Cherrie,

1976; Yang et al., 2006)

Following this discussion of which coal features most influence wireline resistivity

readings in CSG wells, a novel approach is tested in Chapter 4 with the aim of creating

an improved economically-viable way to estimate net coal thickness using deep and

shallow penetrating resistivity measurements in coal.

Question 3:

What is the accuracy of estimating net coal thickness in thinly bedded

coal seams using density logs and other downhole standard wireline

measurement?

Question 4:

Can resistivity measurements be used to improve net coal estimation

both in cored and production wells?

5
Chapter 2

As part of the process for determining the optimal strategy for the petrophysical

analysis of coal seam gas, after the estimation of net coal thickness, the natural next

step is to determine in situ gas content. The interpretation of core analysis methods

establishes the relationship(s) with gas content that are later applied downhole through

the evaluation of downhole measurements. Based uniquely on core analysis or together

with wireline measurements, several authors have established (over the last 40 years)

procedures to estimate in situ gas content (Calvert et al., 2011; Colson, 1991; Hamilton

et al., 2012; Kempton and Peeters, 1977; Kim, 1977; Mavor and Nelson, 1997; Mavor

et al., 1990a; Mullen, 1989, 1988; Nelson, 1999; Rogers, 2007). These techniques

depend on relationships between gas content in coal with coal properties such as rank,

core density, core reciprocal of density and coal composition (namely ash content).

These relationships inevitably have a degree of uncertainty due to the heterogeneous

nature of coal and also due to the limitations associated with core and wireline log

measurements.

To constrain the analysis and improve the estimation of error in the in situ gas

content using wireline measurements, Chapter 5 investigates previous published

techniques and constructs an improved methodology. To do so core analysis and

wireline measurements were interpreted to answer the following questions:

Question 5:
What is the level of error in the estimation of in situ gas content using
published methodologies?

Question 6:
Can a methodology be created to optimize the estimation of in situ gas
content?

Chapter 6 summarises the most important aspects and conclusions of this work by

addressing the results to the questions posed here, which in themselves are presented in

more detail in chapters 3 to 5. Recommendations for further research are also proposed.

6
Chapter 2

CHAPTER 2 CSG Geological and

technical background

2.1. Coal Seam Gas


Coal Seam Gas (CSG) also known as Coalbed Methane (CBM), is the designation of

coal as a natural gas source, with a unique mechanism of gas storage (Harpalani and

Schraufnagel, 1990a). In contrast to coal mining, where coal is the target energy source,

CSG exploits only the gas the coal stores.

It is designated as unconventional gas resource due to the different way gas is

extracted compared to conventional clastic or carbonate rocks. Due to past mining and

oil & gas exploration in many Basins worldwide, the location of coal deposits, their

structure, stratigraphy and, to a certain degree, their thickness are relatively well known

(Figure 2.1). Most of the worlds CSG resources lie in seams that have not yet been

mined, or are unlikely to be developed. A CSG reservoir is an intensively stratified

continuous gas accumulation stretched along a relatively large geographic area, usually

termed a fairway (Creedy et al., 2001; Moore, 2012). Wells are drilled onshore typically

reaching shallower depths than in conventional gas reservoirs (Hewitt, 1984; Nelson,

2000a). Today, commercially viable exploitation of these resources are largely confined

to the USA (San Juan, Black Warrior, Powder River and Central Appalachian Basins)

and Australia (Surat and Bowens Basins). While in the USA the gas is mostly used as a

local domestic source of natural gas, although Australia are pioneering the export of

methane from CSG reserves (Bradshaw et al., 2012; Weijermars, 2012).

7
Chapter 2

Figure 2.1 Worldwide coal bed methane reserves.


Darker shading indicates coal in outcrop or subcrop (from Ahmed et al., 2009).

Besides being unconventional in an exploration point of view coal stores gas in a

different way to a conventional gas reservoir, combining three elements typical of a

hydrocarbon reservoir system. A coal seam in CSG is the source rock, the reservoir and

the seal, in a single rock formation (Faiz et al., 1992; Nelson, 2000b). As discussed in

the following chapter, and as Jones et al. (1988) and Kim (1977) have shown, coal is a

solid, usually highly microporous, highly complex, heterogeneous rock (Koenig and

Stubbs, 1986; Wold et al., 1995). The coal matrix creates conditions that allow higher

volumes of methane gas to be stored in a rock type with limited total porosity (Table

2.1). Molecules of the gas are physically bonded (adsorbed) to the walls of pores found

within the coal matrix (Curl, 1978; Gray, 1987a). These molecules can also be found,

although in much less quantities, within the fracture network as free gas or in water

solute (Levine, 1992a).

The following chapters debate the main challenges faced today in CSG formation

evaluation. It is not a challenge to find gas, but rather to find significant accumulations

with sufficient natural permeability, for it to be commercially viable. The estimation of

8
Chapter 2

reservoir volume, and thus the gas-in-place, is challenging due to the limited thickness

and uncertain lateral continuity of the coal beds normally associated with this type of

reservoir. Also permeability provided by natural fractures is an additional significant

uncertainty. In CSG Basins, these uncertainties can only be quantified by a drilling,

coring, wireline logging and production test programs (Mavor et al., 1990b) forms the

data source of this study.

Reservoir Type Permeability (md) Porosity (%)


Tight Gas Sand <0.1 6 - 12
Unconventional Shale Gas 0.001 0.01 2 -8
Coal Seam Gas 0.1 5 1 - 10
Sandstone 0.5 2900 6 - 35
Conventional
Carbonate 0.1 2200 3 - 30
Table 2.1 Modified table from Gaurav et al. ( 2012) and Ehrenberg and Nadeau (2005).

2.2. From Peat to Coal

Compositionally, coal is comprised of organic and inorganic components that have

undergone a process of diagenesis of the original organic material deposited in the peat.

Plants provide the material that later influence humic coal characteristics such as

porosity, gas content and the degree of cleating (coal fractures). The continuous

transformation of plant material from peat, to what we consider today as coal, is known

as coalification, a process that can be divided into biochemical and geochemical stages.

The main agents during the early stages of coalification (the biochemical stage:

biogenesis) are biological (Kim, 1978). The low drainage rate and lack of fresh water

help to create an acidic, low oxygen environment suited to anaerobic microbial

communities that minimise decomposition and aid the accumulation of carbonaceous

material. Originally deposited vegetation type compounds, mainly lignin and cellulose,

give origin to the microscopically recognizable organic components in coal called

macerals. With continued deposition of overlying inorganic sediments, the biochemical

9
Chapter 2

phase ends at depths of the order of several hundred metres, where percolating water

supplying microorganisms with nutrients is no longer available. The degree of

maturation is referred to as coal rank, and at this point coal can be already classified as

lignite up to sub-bituminous (Langenberg et al., 1990; Figure 2.2).

Figure 2.2 Coalification stages during coal generation.

As coalification enters the geochemical stage, the sediment undergoes diagenesis

driven primarily by temperature (thermogenesis) and to a lesser extent pressure

(Rightmire, 1984). With increasing depth, the plant material is transformed by releasing

volatile matter (water, carbon dioxide and light hydrocarbons, including methane) and

becoming progressively enriched in carbon and structurally more homogenous (Kim,

1978; Langenberg et al., 1990; Teichmller and Teichmller, 1968). If the whole

10
Chapter 2

coalification range is complete the coal reaches the highest rank and is classified as a

meta-anthracite.

As coal formation is a dynamic process, with a great variety of factors involved,

there is no single molecular structure that can represent a coal molecule; the

heterogeneity in structure and composition is simply too great. The only credible

description of a coal molecule is a simplified structure formed by clusters of aromatic

rings with weaker links between them that break thermally during coalification (Kim,

1977; Figure 2.3). As coalification progresses, the coal molecules realign liberating

volatiles (CO2, CH4 and H2O) and hydrocarbon in continuous process.

Figure 2.3 Changes in the chemical and physical properties of coal with increasing rank.
Modified from Teichmller and Teichmller (1968).

11
Chapter 2

2.3. Coal Petrographical Characteristics

2.3.1. Lithotypes

The macroscopically recognizable beds in humic coals beds are called lithotypes.

They are indicative of variations in composition, conditions during deposition and

biochemical maturation of coal (Hower et al., 1990; Stach et al., 1982; Stracher et al.,

2010; Taylor et al., 1998). The macroscopic appearance of humic coal is usually brown

to grey to black, with a banded appearance due to accumulation of duller and brighter

coal laminae (Gamson and Beamish, 1991; Gamson et al., 1993; Thomas, 2002). Each

bed represents different physical and chemical characteristics. Brown coals generally

have a low rank (lignite and subbituminous coals), whereas black coals generally have a

higher rank (from bituminous to anthracite coals). Lithotypes to be recognised as such

must be at least 1cm thick; if smaller these are considered as bright or dull coal laminae

(Langenberg et al., 1990). The more bands a coal has the more variable is its

composition (Hower and Wagner, 2012), indicating changes in type and percentages of

maceral (organic) and mineral (inorganic) components.

For the purpose of this study, the Australian standard system for classification was

followed (AS 2916-2007). This classification defines coals as bright and dull, whereas

under the ASTM (American Society for Testing and Materials) classification scheme,

the same coals would be classed as vitrain and fusain respectively; thus the reader

should beware comparisons using different terminology. The classification schemes and

terminology are compared in Table 2.2. The categories with which the Australian

standard is defined are composed of major and minor constituents of bright to dull

layers in each coal bed. In addition to these categories, there is also the designation of

stony coal for highly impure coal types rich in inorganic components. An example

core and its description are shown in Figure 2.4. It should be noted, that it is not

12
Chapter 2

expected for each lithotype to have a completely uniform composition, even when

showing identical macroscopic appearance (Cao et al., 2011). Although the

identification of lithotypes can considerably aid a better understanding of coal facies

and its characteristics at the regional scale (Lamberson et al., 1991).

(Stopes, 1919) (Schopf, 1960) ASTM standard (1978) Australian standard


Vitrain Vitrain Bright
Clarain Bright
Moderately bright Bright minor dull
Durain Attrital coal Mid. Lustre Dull and bright
Moderately dull Dull banded
Dull Banded Dull minor bright
Fusain Fusain Dull
Impure coal Bone coal Stony or Shaley
Table 2.2 Macroscopic description of coals.

13
Chapter 2

Figure 2.4 Examples of special core analysis tested samples description followed in this research. Pictures provided by the operator.

14
Chapter 2

2.3.2. Macerals
Macerals are organic components and correspond to minerals in inorganic rocks.

They originate from plant material remains and are the primary constituent of coal. The

formation of macerals in the early stages of coalification depends on the local plant

community, climatic and ecological controls, and the depositional environment. There

are three major maceral groups: vitrinite, inertinite and liptinite group (Table 2.3). Each

of these groups contains subgroups with similarities of origin, composition and optical

properties (Diessel, 1992). Dyrkacz and Horwitz (1982) and Dyrkacz and Bloomquist,

(1992) concluded that members of the liptinite group give coal lower densities than

other macerals. Several publications have defended the following sequence of maceral

group density Liptinite<Vitrinite<Inertinite (Choi et al., 1987; Parkash et al., 1983;

Stankiewicz et al., 1994; Taulbee et al., 1989). The organic matter of a coal bed is made

of different macerals due to the fact that coals originating from different vegetation

tissues mixed together in the plants body. These later on give form to a mix of different

macerals that can only be distinguishing using a microscope. These macerals groups

have each their particular density since each maceral originated from a different tissue

(Table 2.3). Another factor contributing to the heterogeneity of coal is the fact that each

maceral matures differently. Dyrkacz and Horwitz (1982) for example concluded that a

motive for the broad variability of liptinite density compared to vitrinite group is the

significant chemical structure changes occurring with maturation. The properties of the

major maceral groups are described in turn below:

15
Chapter 2

Vitrinite

Vitrinite is typically the most abundant maceral in coal, the most homogeneous and

the more conducive to form a natural fracture system (Ahmed et al., 1991; Nelson,

2000a; Paterson et al., 1992). It is in part formed by the remains of lignin, an

amorphous polymeric substance that gives structure to the plant cell wall in conjunction

with cellulose. Vitrinite is chemically richer in oxygen than liptinite macerals and

therefore indicative of high accommodation rates and rapid deposition of the original

plant material (Cohen and Spackman, 1972; Diessel, 1992). High vitrinite content gives

the coal a bright appearance due to extensive changes caused by coalification

(gelification phase) that also leads to its homogeneous structure. Harris and Yust (1976)

analysed high-volatile bituminous coal samples using electron microscopy, and

concluded that vitrinite has the smallest range in diameter (between 2nm and 20nm),

comprising the majority of smaller particles in coal. Due to its highly microporous

structure, vitrinite rich coals are also shown to have the slowest desorption rates if it

does not have an extensively developed fracture system (Clarkson and Bustin, 1997;

Crosdale et al., 1998).

16
Chapter 2

Liptinite

Liptinite, also called exinite, originates as spores, pollen, resins, oils, algae and

waxes. Liptinite macerals have chemical structures high in hydrogen compared to other

macerals (Cohen and Spackman, 1972; Diessel, 1992). In contrast to vitrinite, liptinite

has more resistance to mechanical degradation, since in this case the original plant

material was in part meant for reproductive dispersal. They are also more resilient to

oxidative degradation, so a high liptinite coal is associated with slow deposition rate

with original biological material exposed to oxidising mire acrotelm before

accommodation (Petersen et al., 1998; Taylor and Liu, 1987).

Inertinite

Though originating from the same material as other maceral groups, inertinite

macerals are the oxidized or charcoaled remains of cell walls, having been exposed to

forest fires, bacterial action and oxidation prior to coalification. It is therefore high in

carbon but contains less hydrogen than other maceral groups. As coal matures, it loses

porous volume, though this depends on the maceral constitution (Unsworth et al.,

1989). Inertinite undergoes the least volume change as coal matures, and adds the most

macroporosity to coal of the three principal maceral groups. Using electron microscopy

techniques in high-volatile bituminous coals, Harris and Yust (1981, 1976) observed

that inertinite create pores of 5 to 60nm. It is also the hardest of the macerals and the

least likely to form a good fracture system (Ahmed et al., 1991). Although having less

cleats this maceral type is known to have higher gas diffusivity due to the larger

macropore volume (Clarkson and Bustin, 1997; Crosdale et al., 1998; Karacan and

Okandan, 2001; Unsworth et al., 1989).

17
Chapter 2

Maceral Maceral
Definition/Origin Maceral
group Subgroup
Textinite
Intact cell walls that may or not be visible. From
Texto-ulminite
Telovitrinite the parenchymatous and woody tissues of roots,
stems, barks and leaves. Eu-ulminite
Telocollinite
Isolated of as cemented fine grained macerals, Attrinite
Vitrinite
Detrovitrinite originated from strong decay of parenchymatous Densinite
and woody tissues of roots, stems and leaves. Desmocollinite
Void filling of vitrinitic material. Precipitated Corpogelinite
Gelovitrinite content of plant cells or humic fluids during decay Porigelinite
and diagenesis. Eugelinite
Spores and pollen grains. Sporinite
Cuticles of leaves, roots, stems, needle and shoots. Cutinite
Resins, fats, waxes and oils. Resinite
Degradation residues. Liptodetrinite
Algae. Alginite
Liptinite Cork cell walls. Suberinite
From Vegetable oils. Fluorinite
From hydrogenated substances and Liptinite
macerals.
Exsudatinite
Degradation product from algae, bacterial,
zooplankton.
Bituminite
Partly combustion or biological oxidation of plant
material.
Fusinite
Telo-inertinite Partial oxidized plant material. Semifusinite
Inertinite Mostly fungal remains. Sclerotinite
Redeposited inertines. Inertodetrinite
Detro-inertinite
Degradation of macerals during coalification. Micritine
Gelo-inertinite Oxidized gel material. Macrinite
Table 2.3 Maceral analysis classification following the Australian Standards (AS 2856.1-
2000, AS 2856.2-1998, AS 2856.3-2000, AS 2856-1986).

2.3.3. Cleat system & Permeability


Besides its remarkable capacity to store gas, coal has another attribute that

differentiates it from other sedimentary rocks, which is the capacity to develop closely

spaced cleats at nearly all ranks of maturity. Natural fractures, or cleats, are a result of

the original material found in the peat and the physical and chemical changes which the

coal undergoes during its maturation (Nelson, 2000; Ting, 1977). Coal volume and

mass decrease due to water drainage and devolitisation during maturation, enhancing

the coal bed fracture system (Anderson et al., 2003; Rodrigues and Lemos de Sousa,

18
Chapter 2

2002; Stach et al., 1982). Coal with low mineral content has been observed to have a

larger cleat spacing, indicating how the geochemical transformations during coal

maturation such as shrinkage are important to cleat development (Kendall and Briggs,

1933; Stach et al., 1982). While vertical shrinkage of coal is accommodated downward

by boundary formations, horizontally there are no lateral movements of accommodation

creating vertical spaces (cleats) within the coal beds. Moreover, as coal undergoes

deformation, the regional minimum stress orientation promotes cleat formation by

folding, cooling and uplifting the coal seam (Barton et al., 1997; Ward, 1984). For

example, Enever and Hennig (1997) concluded that the in situ stress regime has a

strong influence on hydraulic conductivity of fracture networks and consequently coal

permeability in Australia. Considering that the regional stress field is normally

anisotropic, these vertical fractures will form normal to the direction of minimum

horizontal compressive stress (Gash et al., 1992).

Cleats form two sets: face cleats sets; the oldest and more continuous, and butt

cleats; less continuous and ending in contact with face cleats (Dron, 1925; Kulander and

Dean, 1993; Laubach and Tremain, 1991). They are orthogonal between each other and

both perpendicular to the bedding (Figure 2.5a), with no significant offset parallel to

the cleat wall and have apertures of less than 0.1mm at surface conditions (Kulander et

al., 1990; Laubach et al., 1998). Macro and micro fractures may be also present. Macro

(master) fractures, are joints and faults that cross the coal seam regardless of the

different lithotypes and non-coal interbeds (Figure 2.5b). Microfractures, not visible to

the naked eye, are considered to be the most recent to form (Figure 2.5c). Master

fractures are relatively rare in coal seams, especially if their numbers are compared

against cleats and smaller fractures, appearing randomly throughout a well regardless of

the type or rank of the coal bed (Dawson and Esterle, 2010; Wold and Jeffrey, 1999).

19
Chapter 2

Cleat horizontal connectivity, depends on the frequency of all cleats (macro to micro)

while vertical connectivity is usually limited by the boundaries between coal lithotypes.

a) b)

c)

Figure 2.5 Coal matrix and natural fracture network.


a) The primary cleat patters in plain view from Laubach et al. (1998);
b) Natural fractures hierarchies in cross-section view from Laubach and Tremain (1991);
c) Top side view of coal matrix and fracture network from King et al. (1986).

Cleat attributes can be relatively uniform at the regional scale but also can change

abruptly vertically and laterally between short spaced wells (Dron, 1925; Kendall and

20
Chapter 2

Briggs, 1933). The cleat spacing is a fundamental physical reservoir characteristic, as it

directly determines the permeability and the friability of the coal bed. Cleat spacing is

considered to be a function of various factors that may or may not change from coal bed

to coal bed, such as coal lithotype, coal rank and coal composition (Clarkson and

Bustin, 1997; Close and Mavor, 1991; Close, 1993; Nelson, 2000a). Similar to gas

content, rank is the primary factor influencing cleat development. It is well recognised

that cleat spacing decreases from lignite to medium volatile bituminous coal and

increases throughout anthracite rank (Ammosov and Eremin, 1963; Ting, 1977).

Nonetheless, coals with identical rank often have very different cleat spacing. Whilst

describing Bowen Basin coal beds, Gamson et al. (1993) registered that the spacing of

the cleats is associated with one individual bed, therefore both sets terminate at the coal

bed boundaries and are non-pervasive. Moreover, since the physical and chemical

nature of the coal changes from coal bed to coal bed vertically, the fracture

development may also change dramatically even for closely spaced coal beds (Clarkson

and Bustin, 1997; Dawson and Esterle, 2010; Tremain et al., 1991). This is in line with

publications which suggest that cleat spacing is dependent on coal composition and type

(Daniels et al., 1996). For instance, bright coal has a smaller cleat spacing than dull

lithotypes (Kendall and Briggs, 1933; Stach et al., 1982). Publications such as Close

and Mavor, (1991), Spears and Caswell (1986) and Tremain et al. (1991) contradict this

however, and suggest that cleat spacing in identical rank coals is linearly proportional to

coal bed thickness regardless of composition or coal type. Cleats may be filled,

partially filled or have no clay, calcite, quartz or pyrite filling, varying within close

spaced beds. These authigenic minerals found in cleats or fractures can block gas flow

pathways or in contrast preserve at least part of the cleat space (Daniels and Altaner,

1990; Daniels et al., 1996; Hatch et al., 1976; Spears and Caswell, 1986).

21
Chapter 2

2.4. Coal as a Gas reservoir

2.4.1. Coal gas origin


The economical interest in CSG comes from its gas producing potential. This gas

can have two primary origins: biogenic and thermogenic (Figure 2.6).

The first methane generated from coal is of methane biogenic origin. As coalification

starts, carbon dioxide and water are the first volatiles generated from coal. They are

generated while subsidence has not yet reached temperatures between 100-150C

(Clayton, 1998; Rightmire, 1984). Very little thermogenic methane is generated below

these temperatures. Methane generation begins in the lignite stage, although, unless

impermeable material is quickly deposited to form a trap, most of the initial biogenic

gas produced is lost due to high pressures (compaction) and/or temperatures (resulting

in devolatilisation) (Scott et al., 1994). Also, and more importantly, in post-coalification

if the coal is uplifted, secondary biogenic gas can be generated, and, under low

temperatures meteoric waters are introduced to the permeable coal beds. This provides

the conditions required for bacteria to metabolize organic components found in coal and

to generate methane and carbon dioxide. CSG production from biologic sources is

normally associated to thick immature (subbituminous) coal at shallow depths (Jones et

al., 2008; Scott et al., 1994).

Before coal reaches a bituminous stage, thermogenic gas starts being produced by

the breakup of carbon-carbon bond chains mainly from liptinite macerals (Das et al.,

1991). The genesis of gases such as methane, carbon dioxide and ethane are primary a

result of temperature increases with time (Whiticar, 1994). Large quantities of

thermogenic methane are generated in high volatile bituminous coals.

During the early stages of the coalification process, before bituminous rank, methane

is slowly generated and occupies the microporous system within the coal. After this

22
Chapter 2

phase, generated methane acts to expel gas into the macropore system (Thimons and

Kissell, 1973). The biggest reserves in CSG come from thermogenic gas, although

biogenic gas can have also economic value (Scott et al., 1994). Nevertheless there is no

clear transition between biogenic and thermogenic genesis which may result from both

being produced from the same coal seams (Ahmed et al., 2009).

Figure 2.6 Coal rank and methane generation from Ahmed et al. (2009).

2.4.2. Gas Storage


In a conventional reservoir gas is stored mainly by compression. In the case of CSG,

minimal gas content is stored by the compression of free gas in the macroporosity and

cleat network (Faiz et al., 2007). Porosity in coal depends on fracture development

which usually does not reach 5% and is water saturated (Ahmed et al., 1991; Gash et

al., 1992; Purl et al., 1991; Young et al., 1991). In CSG, methane and other gases are

stored within the coal, mainly in an adsorbed state (Curl, 1978; Gray, 1987a). The

attraction between gas molecules to a solid surface leads them to pack closer together, a

phenomenon known as sorption (Krevelen, 1981). In coal, sorption occurs when gas

molecules are bound by weak intermolecular attractions, or Van der Walls forces, to the

23
Chapter 2

organic materials (the solid surfaces) that makes up the coal (Brunauer et al., 1940;

Gregg and Sing, 1967).

The gases found in coal beds include: methane, nitrogen, carbon dioxide and ethane.

The principal gas produced from coal beds is methane, reaching a purity of 99%

methane in many cases (Gray, 1987a). Also, ethane and heavier saturated hydrocarbons

could be produced in very low percentages. Hydrocarbons heavier than methane are

more strongly adsorbed and stay this way unless thermal or strong solvent extraction

methods are applied (Levine, 1992a, p. 199). In contrast, nitrogen is less bound to coal

(Surez-Ruiz and Crelling, 2008), and associated with early production or widening of

fractures during production.

2.4.3. Controls on gas content


In CSG there is no currently known trend that relates the nature of coal to its

methane adsorption capacity and gas content due to the following controls:

Maturity

Rank is considered to be a critical factor and determines the methane storage

capacity (Yee et al., 1993). There are two opposing relationships published today

that relate rank with gas storage capacity. Firstly, that gas storage capacity increases

with rank due to an increase in microporosity (Clarkson and Marc Bustin, 1996; Faiz

et al., 1992; Gan et al., 1972; Prinz and Littke, 2005; Prinz et al., 2004), explaining

the direct correlation between gas storage capacity and rank identified in the Black

Warrior Basin (Carroll and Pashin, 2003; Kim, 1977) (Figure 2.7). Secondly, that

this relationship has a U shape behaviour, with a broad minimum values of gas

storage capacity found in the medium volatile bituminous stage (Ettinger et al.,

1966; Schwarzer and Byrer, 1983; Figure 2.8).

24
Chapter 2

Figure 2.7 Adsorption curves against pressure at a given rank from Kim (1977).

Figure 2.8 Trend in gas storage capacity against rank (%, Ro max) in different coal
lithotypes from Laxminarayana and Crosdale (1999).

25
Chapter 2

Porosity & surface area

Porosity volume is distinct from surface area. Pore surface area is defined by pore

size distribution and pore morphology, and these factors influence how gas is stored.

Microporosity is associated with a higher surface area (Mares et al., 2009b). Since

the sorption process is a bond with the pore surface inside the coal matrix, it is

directly related with methane adsorption capacity and may be its main controlling

factor (Clarkson and Bustin, 1999; Faiz et al., 1992; Levy et al., 1997; Mahajan and

Walker, 1978). The relative quantity of microporosity available can therefore be a

fundamental reservoir characteristic, while larger pores can be mainly seen as

important pathways or connectors to fractures and thus gas flow. CSG is usually

produced from shallow depths (less than 1000m) under lower pressures than

conventional reservoirs. In the micropores gas molecules are stored under low

pressures in high concentration with direct bonds to the surface of the pore (Mahajan

and Walker, 1978). If the coal is under high pressure, gas storage in the macropores,

besides the gas being adsorbed on the surface of the pore, is also by forming an

additional layer on top of this one. Any free gas that is present within the cleat

system will not be significant compared to the gas adsorption within the matrix.

Maceral composition

The influence of maceral composition on gas content and gas storage capacity is a

controversial subject. Ettinger et al. (1966) suggests that the greater storage capacity

of gas is due to higher inertinite composition. Creedy (1979), Lamberson and Bustin

(1993), Laxminarayana and Crosdale (1999) and Levine et al. (1995) suggest that

vitrinite has a positive impact on the adsorption capacity of coal, although pointing

out that the rank can be an inhibitor to these effects. Direct comparison between

26
Chapter 2

telocollinite (bright coal) and semifusinite (dull coal) rich coals from the Australian

Bowen Basin gave more pore volume to the inertinite but more surface area to the

vitrinite rich coal (Beamish and ODonnell, 1992; Crosdale and Beamish, 1993)

indicating the higher importance of vitrinite. Liptinite is generally considered to be

the middle maceral in terms of its affinity to store gas, although Karacan and

Mitchell (2003), Grdal and Yaln (2001) and Scott et al. (2007) stated that coal

with liptinite is associated with a higher gas content but present no arguments to

explain this observation. Opposing this other authors did not find in any systematic

way a relation between gas sorption and maceral composition (Carroll and Pashin,

2003; Faiz et al., 1992; Krooss et al., 2002; Schwarzer and Byrer, 1983) giving more

weight to the influence of rank.

Coal lithotype

In terms of gas storage capacity the influence of different lithotypes is as debatable

as maceral composition. Ettinger et al. (1966) suggest that gas yield increases with

dull bands, although brighter coal is usually associated with vitrinite-rich samples

and thus more microporous and a higher methane adsorption capability (Bustin and

Clarkson, 1998; Clarkson and Bustin, 1999; Crosdale and Beamish, 1993; Crosdale

et al., 1998; Hildenbrand et al., 2006; Lamberson and Bustin, 1993; Laxminarayana

and Crosdale, 1999; Mastalerz et al., 2004). Levine et al. (1993) also found that

brighter coal has a higher adsorption capacity than the same rank dull counterparts in

the coal of the Sydney Basin. The same author, testing coal samples from the Bowen

Basin concluded that the different gas adsorption capacity between the bright and

dull coal is minimal, vanishing in higher rank coals and a similar trend was also

found by Laxminarayana and Crosdale (1999).

27
Chapter 2

Coal composition

Moisture content is an influential factor that negatively affects gas yield and gas

storage capacity (Bustin and Clarkson, 1998; Crosdale et al., 2008; Joubert et al.,

1974; Levine et al., 1993; Levy et al., 1997; Yalin and Durucan, 1991; Yee et al.,

1993). Moisture is considered to affect gas holding capacity in coals by filling the

micropores which would otherwise be gas saturated, competing with methane for

adsorption sites, or by blocking the access of the gas into the micropores (Levy et al.,

1997; Mares et al., 2009b). There are several relationships in the literature that relate

the methane adsorption capacity of coal to moisture content following the equation

2-1:

2-1
Vd=Vm(1 + x*m)

Where:
Vd Methane storage capacity in dry coal, g/cm3
Vm Methane storage capacity in moist coal, g/cm3
x Multiplier
m Moist yield in coal, wt.%

Depending on the coals analysed, the methane storage capacity is differently affected

by moisture, changing the value imputed in the multiplier x in equation 2-1. Ettinger

et al.(1966) and Joubert et al. (1974) reported x should be 0.31 and 0.23 respectively

in different American Coals. Studying coals from the Bowen Basin Levy et al.

(1997) concluded that x should be 0.39. For example, in bituminous coal the

maximum sorption capacity of coal reaches its peak at zero moisture content and

decreases until the maximum sorption holding capacity of moisture is achieved,

known as equilibrium of moisture, at which point methane adsorption is no longer

affected (Joubert et al., 1974). In another example, analysing Bowen Basin coal

28
Chapter 2

samples, Levy et al. (1997) observed that at a pressure of 5MPa, a decrease in gas

adsorption capacity for each 1% increase of moisture content by a rate of 4.2g/cm3.

In rich vitrinite coals this could be exaggerated, as indicated by Levine (1993), and

that these macerals have a more open structure compared to others, with hydrophilic

groups able to adsorb more water molecules.

The amount of fixed carbon content is associated positively with coal maturity and is

strongly related to gas storage capacity (Faiz et al., 1992; Levy et al., 1997). There is

a generalised acceptance that ash content constrains gas content, gas generation and

even natural fracture (cleat) development (Mavor et al., 1990b; Schwarzer and Byrer,

1983; Spears and Caswell, 1986; Tremain et al., 1991). Laxminarayana and Crosdale

(1999) concluded that ash content works as a dilutent in terms of the gas adsorption

capacity of coal. Moisture and ash are undesirable in conventional coal mining and

CSG, as fixed carbon and volatile matter are the only components that could lead to

the production of energy. The sum of these two are what is called organic matter

content, and is directly related to sorption capacity and gas content (Faiz et al., 2007;

Levine, 1992a). Therefore coal quality is usually corrected from as analysed (or raw)

to a dry-ash-free basis, where the received sample mass is normalized to 100%

organic content.

Reservoir Pressure & Temperature

Reservoir pressure & temperature are both conflicting parameters when it comes to

gas storage capacity and gas content. Gas storage capacity is directly related to

pressure. As pressure increases, so should the gas storage potential. At in situ

conditions, if the gas generation occurs in coal simultaneously as reservoir pressure

is increased, i.e., if the gas is purely of thermogenic origin, and is not lost with time,

29
Chapter 2

gas content increases with reservoir pressure (Scott, 2002). In contrast, as

temperature increases, the sorption capacity of coal reduces (Khavari-Khorasani and

Michelsen, 1999; Kim, 1977; Scott, 2002; Yang and Saunders, 1985). Figure 2.7

and Figure 2.9 demonstrate these two opposing effects.

Figure 2.9 Adsorption curves against temperature at a given rank from Kim (1977).

2.4.4. Coal Gas Saturation


CSG exploration tends to focus on gas content estimation and generally disregards a

fundamental associated parameter, gas saturation. The term saturated is used when the

coal at a given pressure has the maximum gas content it can sustain. As in conventional

reservoirs, besides the fact that coal is also the source rock, gas is created in coal by

thermogenic or biogenic processes and not related with the physical structure of the

coal (Mares and Moore, 2008). Therefore gas content and gas saturation can have

distinctly different profiles downhole (Figure 2.10).

30
Chapter 2

Figure 2.10 Examples of vertical profiles of desorbed gas content.


(a) and gas adsorption capacity;
(b) from subbituminous Waikato coal seams in New Zealand from Mares et al.( 2009a).

Depending on gas saturation and reservoir pressure, coal as a reservoir will show

different production behaviours with critical economic significance (Mares et al.,

2009a; Nelson et al., 2000). Permeable coal beds that are saturated will produce gas

immediately, while on the contrary, under-saturated coals may or may not produce. If a

coal is under-saturated it can only produce when the reservoir pressure is reduced to the

point of the saturation pressure of the coal (Jenkins and II, 2008). This process of

reducing reservoir pressure is triggered by dewatering the coal, which can take days,

months or even years depending on the saturation and permeability of the coal. Gas

saturation can vary significantly with depth and laterally (Malone et al., 1987; Pashin,

2010).

Since there is no measurement today that can estimate gas saturation downhole, it

adds a significant margin of error to gas-in-place and production estimation (Mares et

al., 2009a). As will be demonstrated in CHAPTER 3, the general assumption that gas

saturation increases with depth, because of the increasing reservoir pressure, is not

31
Chapter 2

straightforward. Gas saturation can change within a coal seam group through time and

with depth (Figure 2.11). There are several possible explanations for these variations.

Sorption capacity can increase or decrease as a result of uplift, where a coals initial

saturation level can change as a result of the new conditions of pressure and

temperature (Faiz et al., 2007; Scott et al., 1994; Yang and Saunders, 1985), especially,

as both of these parameters decrease, they have opposite effects on coal sorption

capacity. If the case is of a coal that was saturated, then the gas can migrate over time to

overlying porous beds (Bachu and Michael, 2003; Bustin and Bustin, 2008; Zhang et

al., 2008) or even to other coals beds (Faiz et al., 2007). The gas content could also be

increased due to introduction of meteoric water in permeable coals creating conditions

for bacteria to thrive and form secondary biogenic gas (Faiz and Hendry, 2006; Faiz et

al., 2007).

Figure 2.11 Schematic gas generation against gas storage capacity related with depth and
time from Moore (2012).

32
Chapter 2

2.4.5. Gas flow


Desorption is observable to the naked eye when core samples are brought to the

surface and bubbles of gas form on the surface of the coal sample in a mixture of gas

and formation and borehole fluids. At this stage gas desorption test samples are selected

from the recovered core and placed in leak proof canisters. Diffusion (K) is a

fundamental process in CSG producibility, as it is the process of transportation of the

gas within the matrix, where 95% of the gas resides. In coal not yet drilled, the gas

pressure in the reservoir (matrix and fracture systems) is in equilibrium. Gas flow is

triggered by changes in concentration provoked by the drilling of nearby well until a

new equilibrium is achieved (Cervik, 1967). Desorption is the process by which gas

molecules after disconnecting from the porous surface system within the coal matrix

flow as free gas (Figure 2.12). While adsorption relates to the resource potential, the

desorption relates to the actual resource production (Ahmed et al., 1991). This

mechanism has long been considered as hazardous when mining coal, as mining works

release accumulations of methane within a coal seam, and if not prevented may lead to

fires, explosions and collapse of the mining structure (Cervik, 1967; Gray, 1980; Patton

S.B., 1996; Rightmire, 1984). In CSG exploration, desorption data is fundamental, as it

is the only analytical source to date that provides knowledge of the coal methane

charge, and allows the gas reserves to be extrapolated.

33
Chapter 2

Figure 2.12 Methane flow model through coal from (Gamson et al., 1996).

The rate at which desorbed gas leaves the matrix is a major factor in the initial stage

of gas production. The faster the gas desorption is the shorter is the time required for

gas to be produced (Swayer et al., 1987). At depth, under reservoir pressures, methane

molecules attached to coal surface pores are released from their Van der Vaals forces

and flow in complex systems of micropores (less than 2 nm diameter), mesopores

(between 2 and 50 nm) and macropores (>50nm) that the organic content and mineral

content form, before reaching the cleat system. The desorption rate depends on the

following factors:

Rank & depth

As coal increases in rank, the speed at which gas is released diminishes. As maturity

increases the microporosity becomes more dominant with a decrease in

mesoporosity and macroporosity (Gamson et al., 1996; Gan et al., 1972;

Laxminarayana and Crosdale, 1999). Desorption decreases with maturity, especially

in bright coals. The rearrangement of coal to a highly microporous structure while

losing volume is an inhibiting factor to diffusion and permeability (Krevelen, 1981;

Nandi and Walker Jr, 1970) (Figure 2.13). In addition, besides the matrix, the

natural fractures change with rank and depth. Ryan (2003) concluded that

34
Chapter 2

permeability increases with cleat development. This increase occurs up to the

pressure (depth) at which the ability of the coal matrix to withstand lithostatic

ceases, promoting the closing of the cleats and exponentially decreasing the

permeability. As a consequence, economically viable CSG prospects are usually

immature and at shallow depth (pressure).

Coal type

The coal type, composition and pore structure all have a strong influence on methane

desorption. Purely bright coal has been reported to have slower desorption rates than

their dull counterparts (Jones et al., 1988). Laxminarayana and Crosdale (1999) and

Crosdale et al. (1998) observed that in many cases dull (inertinite rich) coal beds

have desorption rates two to three times faster than bright (vitrinite rich)

equivalents. Brighter coal beds tend to be more homogeneous, not only in terms of

being more microporous because of their maceral composition, but also due to their

generally lower mineral content, which is usually dominated by macropores and

mesopores (Clarkson and Bustin, 1997; Karacan and Mitchell, 2003). Furthermore,

coal beds are likely to occur with a variable number of centimetric bedding planes

(or bands) creating an additional gas flow pathway (Gamson et al., 1996;

Laxminarayana and Crosdale, 2002; Mares and Moore, 2008).

Natural fracture network

The degree of fracture development (spacing, connectivity and width) and the

regional stress field also determine the gas flow within the coal bed (Harpalani and

Schraufnagel, 1990a, 1990b; King, 1985). Besides the complex multi scale porous

system, coal also has a multi scale cleat system with macroscopic and microscopic

35
Chapter 2

fractures, for which the interconnectivity can be observed using electron microscopy

(Gamson et al., 1996, 1993; Harris and Yust, 1976). The closer the spacing of the

cleats, the greater the number of beds fractured within a coal seam; bigger apertures

and less fracture filling will result in faster methane diffusion out of the coal. With

fewer cleats, methane will need to travel greater distances and will be less sensitive

to pressure reductions as the coal beds produce.

Figure 2.13 Estimates of percent volume decrease versus rank and temperature from Ryan
(2003).

2.5. CSG reserve & production

2.5.1. Gas-in-place estimation


To economically produce methane from coal, a knowledge of coal distribution, coal

maturity, gas content, coal natural permeability, deposition and structural setting, and,

in the case of biogenic gas production, ground water flow is needed (Scott et al., 1994).

Gas-in-place is the volume of methane stored within a defined rock volume. Four

parameters are needed to calculate the gas-in-place in CSG (Jenkins and II, 2008;

Mavor and Nelson, 1997; Nelson, 1999): coal reservoir area, gross coal reservoir

36
Chapter 2

thickness, average reservoir coal density, and average in situ methane content. The

equation (2-2) mostly used today to estimate gas-in-place is:

2-2
  1359.7
c
c

Where:
G Gas-in-Place, scf
A Reservoir Area, acres
h Thickness, feet
 Average in situ coal density, g/cm3
c

c Average in situ gas content, scf/ton

Its a challenge to accurately determine these parameters and they are commonly

under or overestimated. These parameters are estimated based on: structural maps,

seismic data, wireline logs, special core analysis and production well tests (Koenig and

Stubbs, 1986; Rieke III et al., 1981). Other methods can be also used, though with a

significant increase in uncertainty, such as using conventional drill cutting for core

thickness and gas content estimation (Mavor et al., 1994). Nelson, (1999), Mavor et al.

(1996) and Lagendijk and Ryan (2010) listed the main sources of error in gas-in-place

estimation that still have a impact today:

Lateral reservoir variability

Geologic heterogeneities like coal bed pinch out, compositional and permeability

changes create unexpected reservoir behaviour identified in gas and/or water

production (Nelson, 1999; Figure 2.14). Differences in mineral and moisture

content, sorption capacity of the organic content, variability of the cleat

development, thinning and thickening of coal beds all affect gas content distribution

locally and regionally (Scott, 2002).

37
Chapter 2

Figure 2.14 CSG reservoir heterogeneity.


Structural, stratigraphic and petrologic variations disrupt the lateral continuity of the coal
seams and these complicate the determination of reservoir geometry, coal permeability and the
estimation of gas in place from Mavor and Nelson (1997).

Net coal thickness estimation errors

The main source of data for determining potentially productive CSG coal beds is by

the interpretation of wireline logs. The selection of the maximum density cut-off

used to estimate the thickness of the coal bed can lead to underestimation or

overestimation depending on the approach taken to define what should be considered

as a potential methane producing reservoir (Nelson, 1999). Calculations by

Lagendijk and Ryan (2010) indicate that the uncertainty related with coal thickness

estimation has a great impact on regional development models.

Gas content errors

The in situ gas content can only be determined by measurements of gas desorbed

from fresh cut core samples. Such estimation has usually several sources of error. The

loss of gas during sample coring needs to be estimated indirectly and in situ residual gas

38
Chapter 2

can vary significantly (Mavor and Nelson, 1997; Nelson, 1999). The estimated sorption

capacity of a given coal is another source of error, especially considering that this

analysis is typically under-sampled (Mares et al., 2009a).

2.5.2. Production
Fractures are considered to be a critical factor, which if not present can jeopardise a

CSG prospect (Close, 1993; Gash et al., 1992; Gray, 1987a, 1987b; Kolesar et al.,

1990a, 1990b; Mavor et al., 1990a; Ting, 1977). For methane production to start,

reservoir pressure must be reduced. Reduction of the reservoir pressure is achieved by a

de-watering process, where producing water through the cleat system triggers gas

desorption (Sung and Ertekin, 1987). Therefore productivity in coalbed reservoirs is

associated with the distribution of natural fractures. The dominant type of movement

depends on cleat development. If the coal is highly fractured the flow is mainly caused

by pressure gradient Darcy flow, if that is not the case it is mainly a result of

concentration gradient by diffusion within the matrix (Cervik, 1967; Hemela et al.,

1982; Law et al., 2002; Wyman, 1984). The distinction of these two regimes also

depends of the pore diameter. If the reservoir structure is dominated by macroporosity,

for example, coal theoretically can be associated with the Darcian flow of a

conventional reservoir (Gray, 1987a).

There is debate as to whether the role of conducting fluids is provided exclusively by

the vertical surfaces (the face and butt cleats) (Figure 2.15a). The matchsticks model

was tested over laboratory data following the idea that the gas flow is mainly supported

by the cleat system, presuming that the bedding plane surfaces are closed by lithostatic

pressure and contribute little or nothing at all to the gas flow (Seidle et al., 1992). This

contradicts the previous cubic model associated to coal, that also included the

39
Chapter 2

bedding planes as paths of water and gas flow (Reiss, 1980) (Figure 2.15b). Gamson et

al. (1996) maintain that both models are too simplistic, stating that, besides the cleat

network and bedding planes, gas flow is also dependent on the degree of mineralization.

This is considered further in CHAPTER 3.

a)
b)

Figure 2.15 Conceptual free gas and water flow models.


a) matchstick model from Seidle et al. (1992).;
b) cubic model from Reiss (1980).

There is evidence of high variability in terms of productivity even when production

comes from vertically close spaced seams. One of the primary explanations for this is

significant changes in fracture development (Weida et al., 2005). In CSG, hydraulic

fracturing of the coal seam is usually an exceptional procedure as coal seams in CSG

wells are usually thin, vertically separated from each other and could be horizontally

discontinuous, making it sub-economical to fracture each well one by one (Jones et al.,

1988). Exceptional cases where the coal seam is targeted to run a small hydraulic

fracturing process are the ones where it has been proven to be gas saturated, relatively

thick and horizontally continuous (Koenig and Stubbs, 1986). Therefore, CSG reservoir

producibility at a regional scale needs to rely on natural factures (Arastoopour and

Chen, 1991; Cervik, 1967; Jenkins and II, 2008; Wold and Jeffrey, 1999). It is then

common practice in CSG exploration to describe coal core samples in terms of fracture

characteristics such as cleat spacing, length and mineral filling degree and type.

40
Chapter 2

Permeability over time, with production, changes in three primary ways (Guo et al.,

2007; Sparks et al., 1995). The initial permeability may depend on the saturating phase

in the cleats and the presence of methane in the microporous system. This effect is well-

known in the conventional oil and gas industry. Second, the effective stress change

during production, where decreases in fluid (water) pressure increase the effective stress

and constrain cleat width. The third mechanism that changes the natural permeability of

the coal is due to an increase in permeability with time as a result of coal matrix

shrinkage with the slippage of gas. Lama and Bartosiewicz (1982) mentioned that

although the shrinkage, even being less significant compared to the increasing effective

horizontal stress, could alleviate the horizontal stress of the vertical and sub-vertical

cleats, and therefore maintaining or even increasing the permeability. Also St. George

and Barakat (2001) stated that the desorption of methane is equivalent to a pressure

increment of more than 2.5 times the stress change over the coal bed. Equally, swelling

with gas adsorption has been well documented (Cody et al., 1988; Green and West,

1985; Gregg, 1951; Hargraves, 1963). It is presumed that both increasing/decreasing

permeability processes are happening simultaneously during production in different

parts of the coal seam (Palmer and Mansoori, 1996). Seidle et al. (1992) reported that a

decade of production did not change the coal permeability of the Oak Grove Field in the

Warrior Basin or the Cedar Hill field in the San Juan Basin, as no decreasing

permeability data were found in any literature from actual field production data. In

contrast, Harpalani and Chen (1997) and Gray (1987b) concluded that coal permeability

can increase significantly with decreasing reservoir pressure. Besides volume and stress

changes during production, other parameters such as capillary pressure and phase

relative permeability, both of which are also highly variable, make estimating CSG

41
Chapter 2

reservoir performance and permeability characteristics through time a difficult

nonlinear problem (Gray, 1987a) .

As previously mentioned, the de-watering process can take a very significant time.

The natural storage capacity of coal gives this unconventional reservoir a typical three-

stage production character (Figure 2.16). Usually, during the initial stages of

production most coal cleat systems are water saturated. As pressure decreases by water

production, gas desorbs and flows along with the water through the cleat system. CSG

therefore is actually a two-phase (gas and water) flow.

Figure 2.16 Typical coal seam gas and water production characteristics from Anderson et
al.(2003).
Normally in the first stage, initial production is almost completely dominated by water (Ried
et al., 1992; Sung and Ertekin, 1987). With a reduction of pressure and water saturation in the
cleats, the production will enter stage 2 where significant gas production starts. Stage II starts
when the critical desorption is reached, and this can lead to a significant jump in gas
production. Gas production will keep rising, reaching a stable production stage (sometimes
considered to be a separate stage). After gas production reaches its peak, in the beginning of
stage III, a slow decline phase starts.

42
Chapter 2

2.6. CSG formation evaluation


Drilling for CSG has distinctive particularities from conventional reservoir drilling

operations (Gray, 1987a). The first difference is the number of wells needed to

investigate and produce methane from coal beds. The lateral variability of coal

characteristics such as coal thickness, composition, gas content and permeability

underline the necessity for intensive drilling in what can be considered to be a relatively

small geographical area (Boyer, 2004; Johnston et al., 1991; Pashin, 1998). Wells are

commonly drilled with a spacing of less than 1km. A minimum of 5 wells are needed to

evaluate the performance of a reservoir. For a given area of the prospect, one of the five

wells will core the entire target formation. This becomes a key well, and the recovered

coal is submitted for special core analysis. The remaining four exploration wells are

used for permeability and production testing (Jenkins and II, 2008; Weida et al., 2005)

and no core is recovered. Wireline tools are generally run on all five wells and aid in the

correlation between the key well and production wells. Wells that are planned to be

cored have drilling practices that differ slightly from the drilling of a production well.

These two types of well differ in diameter, where the key wells are 3 3/4inches (9.6cm)

and the production wells are 8 1/2inches (21.6cm). Another common characteristic of

the wells is their shallow depth. Coal, although it increases its gas storage capability

with depth, loses permeability, so that CSG exploration usually only targets shallow

coal beds (Bodden and Ehrlich, 1998). Wells usually do not reach a total depth of 1km

and are drilled using underbalanced rotary-percussion methods, where despite the rapid

drilling rate, the objective is to minimise formation damage (Boyer et al., 1986; Mavor,

1994; Montgomery, 1999).

43
Chapter 2

In CSG exploration the heterogeneous nature of coal requires a series of distinct

measurements. Core analysis, together with wireline measurement and well tests

comprise the bulk of the data set for CSG formation evaluation. To predict the reservoir

behaviour and economic productivity it is necessary to relate all of these measurements

under very different measurement scales. The interpretation of all these data creates

obvious concern over production accuracy (Zuber and Olszewski, 1992). In the next

three sections the measurements relevant to CSG interpretation are scrutinised.

2.6.1. Core analysis


Core analysis data acquired for this research consist of a series of core tests which are

common practice in CSG exploration as part of the gas-in-place estimation procedure

(Mavor and Pratt, 1996). The most relevant reason for core recovery is to estimate the

in situ gas content, this can only be accurately achieved by collecting freshly cut coal

samples and placing them in desorption canisters immediately after they reach the

surface. Additionally this special core analysis determines the basic composition of the

sample, i.e. proximate analysis, maceral composition, rank and gas storage capacity

(Mavor et al., 1990a). To a certain extent, in conventional reservoirs formation

evaluation, lack of core data is often compensated by the analysis of drill cuttings. In

CSM, however, the multiple bed distribution of the coal associated with the major CSG

exploration Basins leads to this source of data (cuttings) to lose its relevance due to

difficulties in determining the sample depth origin. Also, gas content estimation is

seriously compromised due to the immediate escape of gas after the coal has been

crushed in situ by the drill bit (Mavor et al., 1990a).

44
Chapter 2

2.6.1.1. Gas Desorption Testing

Samples recovered using a slimline coring system are sealed immediately upon

retrieval at the wellsite. The gas volume released from the sample is measured with

time, over variable time periods, following the decrease in rate of desorption, starting

with two minute intervals up to a weekly basis with a maximum period set by operator

which in this case was no more than six months.

The reported total gas yield of a sample is done by adding what is normally

designated as lost gas (referred as Q1), desorbed gas (Q2) and residual gas (Q3). Q1 is

the portion of the gas lost during recovery of the core sample prior to its confinement

inside the desorption canister. Q1 is estimated through numerical analysis of measured

gas desorption rate data (MacLennan et al., 1995) and is therefore generally considered

the least reliable component of the total gas content (Diamond and Schatzel, 1998). It is

usual to find great variation in gas content in the coal core samples recovered (Jenkins

and II, 2008). The reliability of such an estimation depends on several factors including

technical issues such as: recovery time and type of drilling fluid and geological issues

such as: coal type, physical condition of the sample, water saturation, permeability, gas

yield and amount of free gas (Kissell et al., 1973; McCulloch et al., 1975). Once core is

retrieved to the surface, cut and sealed in a desorption canister, the gas desorbed is

measured directly (Q2). During this phase gas samples are taken to measure gas

composition. The desorption rate of the Q1 and Q2 is influenced, besides coal nature, by

the physical character of the retrieved sample (Kissell et al., 1973; McCulloch et al.,

1975). After the desorption test is complete the sample still contains some gas due to

low permeability/diffusivity and because at a pressure of 1 atm. coal may still hold

methane in equilibrium (Levine, 1992b). Q3 is then calculated by crushing the sample,

in a mill at the same temperature as the samples had previously desorbed, and

45
Chapter 2

measuring the gas release. Q1 and Q2 are considered to be the parts which can actually

be produced from the reservoir in contrast with Q3 which cannot (Eddy et al., 1982).

The argument of the unproducibility of residual gas is that this gas is only liberated in

the laboratory under atmospheric, thus below in situ conditions. Furthermore, as

indicated before (section 2.5.1), gas content in CSG exploration is a source of error

because of the lack of data regarding how much of the in situ gas is residual. The US

Black Warrior Basin coal samples indicated that residual gas content is usually 5 to 8%,

but it could be as high as 45% (Eddy et al., 1982). Desorption procedures cannot be

rerun, or at least if they are rerun, they are not be representative of in situ conditions,

due to changes coal properties due to oxidation, water, gas and volume loss (Clarkson et

al., 2011; Guo et al., 2007; Mavor and Gunter, 2006; Purl et al., 1991).

The gas yield is reported in raw (as received) and as dry-ash-free (d.a.f.) basis. The

raw basis is relevant to understanding the gas content considering the entire sample,

regardless of non-coal layers or moisture content. Using d.a.f. the weight of the sample

is corrected for non-coal components such as ash and moisture by considering that the

organic content is the only component that contains gas (Faiz et al., 2007). In terms of

analysis and correlations, results in d.a.f. are desirable (Scott et al., 1995) as long as the

methodology is consistent and stays faithful to the same basis. The final values of both

bases are then presented in a volume to weight ratio (m3/tonne).

46
Chapter 2

Figure 2.17 Apparatus required for desorbed gas (Q2) estimation.


Photo from http://www.spe-qld.org/useruploads/files/short_course/fes10-
01_short_course_oct_15_2010_crosdale.pdf.

2.6.1.2. Maceral Analysis

The maceral analysis is a procedure that identifies the various organic constituents of

coal (macerals), which have distinct properties when observed under transmitted or

reflected light subjected to UV-induced fluorescence (Stach et al., 1982). There are

three major maceral groups: Vitrinite, Inertinite and Liptinite group. Each of these

groups contains subgroups with similarities in origin, composition and optical

properties (Table 2.3). The vitrinite group has high reflectance compared to liptinite

and low reflectance compared to inertinite. In contrast Liptinite is associated with dull

coals, as the members of this group show low reflectance under the microscope.

Inertinite is distinguishable from both Vitrinite and Liptinite by its highest reflectance.

47
Chapter 2

The testing procedure and maceral classification in this study was done following

Australian Standards (AS 2856.1-2000, AS 2856.2-1998, AS 2856.3-2000, AS 2856-

1986).

2.6.1.3. Proximate Analysis

In the data set used in this study all core selected for desorption testing underwent

proximate analysis acquired from the operator to provide percentages of the four basic

constituents, fixed carbon, volatile matter, moisture, and ash yield of coal following the

Australian standard (AS1038.3-2000). This procedure determines coal quality which

has a determinative impact on coal as a reservoir (Bustin and Clarkson, 1998; Carroll

and Pashin, 2003; Kim, 1977; Lamberson and Bustin, 1993; Levine, 1993). This

analysis is based on relatively simple testing conditions where the distribution of the

different components of the rock is determined in three separate steps as the sample is

subjected to increasing heat under specified conditions.

The first component of coal to be estimated is moisture. This moisture can have

several sources, from the original vegetation which coal was formed, where water was

both physically and chemically bound, meteoric water later introduced, or borehole

fluid introduced while drilling (Riley, 2007). In this context moisture content

terminology refers to the results of a procedure to determine what is designated as air-

dried moisture. This is estimated by calculating the loss of weight of the sample

received in the laboratory after it has been exposed to low, slightly above ambient

temperatures. Moisture content has a negative relationship with rank, for example

subbituminous coals have very high moisture content (>25%) while bituminous coal are

known to have much lower values (<10%) (Nelson, 1999; Pratt et al., 1999; Rightmire,

48
Chapter 2

1984). Scott et al. (2007) note that as coal matures and compaction progresses moisture

is lost and therefore moisture content is expected to decrease with increasing depth.

Ash content is the fraction of non-coal material in the sample (Mavor et al., 1990b).

It is determined by weighing the remaining sample after it has gone through

combustion. It is composed of mineral matter disseminated in the coal, interbeds of

other rock types in the sample or mineral deposits inside the natural fractures. Ash

content can be more, equal to, or less than the mineral matter, depending of the nature

of the mineral matter and any chemical changes during heating (Riley, 2007). The ash

content varies widely depending on the core sample selection where the presence of

other rock type interbeds will provoke a significant ash content increment.

Volatile matter represents the components of coal liberated when it is heated at

standard conditions and a temperature of 900oC, such as hydrogen, carbon monoxide,

methane, ethane, carbon dioxide and also water vapour. Volatile materials derive from

thermal decomposition of the various components of coal during maturation, that are

only liberated if subjected to laboratory high temperatures (Langenberg et al., 1990;

Riley, 2007). Volatile matter has been associated with maceral composition of coal.

Inertinite has the lowest and liptinite the highest concentration of volatiles (Mathews et

al., 1997; Stach et al., 1982). The volatile content is calculated by loss of weight minus

the air-dried moisture.

Fixed carbon is mainly carbon and minor quantities of sulphur, hydrogen, nitrogen

and oxygen that remains after the removal of moisture, ash and volatiles. This

component is measured indirectly, by the difference of the sum of the percentages of

other components. As can be seen in Table 2.4, ASTM defines rank classes using

volatile matter and moisture in percentage of total weight and fixed carbon on a d.a.f.

basis.

49
Chapter 2

Together with proximate analysis, moisture, ash, volatile matter and fixed carbon,

two other physical properties are measured, relative density and sample volume. The

sample volume is calculated by the difference between the empty desorption test

canister volume and the headspace volume. Also designated as core density or apparent

density, relative density is determined by weight loss of the original core sample when

immersed in water (AS1038.3-2000; Calvert et al., 2011; Mavor and Nelson, 1997;

Thomas, 2002). Although there is variable air remaining inside the core analysed, this is

the closest measurement to what could be considered true density. Helium expansion

density is also a standard measurement in special core analysis in CSG exploration.

This measurement is associated with adsorption isotherm determination and therefore

the sample tested in this case is a pre-crushed composite of several samples from the

same seam.

Up to 5 gram sample size

Figure 2.18 Thermogravimetric Analyzer used for the proximate analysis estimation.

50
Chapter 2

Vitrinite reflectance Volatile matter Moisture


Coal rank
(Ro max) (wt.% dmmf) (wt.%)
Class Group
Peat
75
Lignite B
Lignitic
Lignite A 35
0.42
Sub-bituminous C
Sub-bituminous Sub-bituminous B 25
Sub-bituminous A
0.50 8 - 10
High-volatile C bituminous
High-volatile B bituminous
0.75
Bituminous High-volatile A bituminous
1.12 31
Medium volatile bituminous
1.51 22
Low volatile bituminous 14
1.92
Semi-anthracite 8
2.50
Anthracitic Anthracite
2
Meta-anthracite

Table 2.4 ASTM rank classification of coal approximate boundary values from Stach et al.
(1982) and Ward (1981).

2.6.1.4. Adsorption Langmuir Isotherm Analysis

The most common model used to relate storage capacity of coal, pressure and

moisture is the Langmuir Isotherm and isotherm analysis has become essential practice

in CSG exploration worldwide (Mavor and Nelson, 1997). Gas sorption capacity of coal

contributes to knowledge of gas storage capacity and the gas saturation of the reservoir,

the pressure at which the gas is released, the amount of methane produced as pressure is

reduced and the remaining gas when the reservoir is abandoned. Samples are tested at

increasing pressures including at in situ conditions, up to 10MPa, at equilibrium of

moisture and at constant (reservoir) temperature. The shape of the curve is used to

predict production performance from a coal as the reservoir pressure is reduced (Figure

2.19). Analysis of the data is performed by presuming that the sample is representative

of a typical Langmuir isotherm curve (equation 2-3). Based on work by Langmuir

51
Chapter 2

(1918) the experiment focuses on the nature of gas adsorption onto solid surfaces;

Langmuir pressure is the pressure value at which gas capacity equals 50% of the

maximum storage capacity (Langmuir volume), and is an important indicator of the

shape of an isotherm.

2-3
P
s  sL1  a  m
PL + P
Where:
GS Gas storage capacity, m3/ton
GsL Dry, ash-free Langmuir storage capacity, m3/ton
P Pressure, psia
PL Langmuir pressure constant, psia
a Ash content, fraction
m Moisture content, fraction

Equation 2-3 represents the adsorption of a given gas, methane alone or together

with other gases (called an extended Langmuir isotherm), in coal under increasing

pressures based on monolayer coverage of the surface area found in the pore walls of

the coal. The plateau of the isotherm indicates the filling of the monolayer at higher

pressures and establishes the point at which the coal is saturated (Gregg and Sing,

1967). Wyman (1984) stated that gases in the absorbed state form a 4- thick layer on

coal surface.

52
Chapter 2

Figure 2.19 Retention of gas in coal seams redrawn from Surez-Ruiz and Crelling (2008).
Saturated coal bed will only produce gas when the pressure is lowered under the critical
desorption pressure. Many of the coal bed found are not completely gas saturated and is
necessary to lower the reservoir pressure even lower to achieve production critical desorption
pressure of undersaturated coals. The amount of gas produced is the distance between the

points of initial production () and the abandonment pressure ().

The analysis is based on testing a sample for maximum CH4 or a CH4 + CO2 storage

capacity at increasing pressures and constant (reservoir) temperature. Carbon dioxide

sorption is greater than methane (Figure 2.20) (Mares et al., 2009b; Mastalerz et al.,

2004) with a ratio of 2:1 being reported (Clarkson and Bustin, 1999; Levy et al., 1997;

Pashin et al., 2003; Rodrigues and Lemos de Sousa, 2002) or greater depending upon

the pore dimension (Faiz et al., 1992) and vitrinite content (St. George and Barakat,

2001). This is due to the ability of carbon dioxides smaller kinetic molecular size to

reach smaller pores, and to block the entrance to larger pores (Cui et al., 2004) and also

the stronger polarity of CO2 and electrostatic forces (Gregg and Sing, 1967).

53
Chapter 2

Figure 2.20 Example of adsorption performances of different gases produced from low
bituminous coal sample from western Canada from Surez-Ruiz
Ruiz and Crelling (2008).

2.6.1.5. Reflectance Analysis


As Ahmed et al., (1991
1991) indicated, all the commercially viable CSG in Australia and

United States are in the bituminous


ituminous range. The stage of maturation can be quantified by

measuring under reflected light the vitrinite maceral constitution of a given coal. Due to

a fairly regular alteration through maturation and the common abundance of vitrinite in

coal, the reflectance measurement of this maceral group is the most common method of

estimating rank. As shown in Table 2.4 volatile matter, fixed carbon and moisture

content estimated through proximate analysis are also parameters correlatable with

rank. Macerals can be distinguished


distinguishe under the microscope, and submerged under oil the

incident light can identify vitrinite due to its dark-light


dark light gray medium-reflectance
medium

appearance, while liptinite is dark and inertinite is highly reflective (Stach et al., 1982).
1982)

The reflectance analysis was conducted by Energy Resources Consulting Pty Ltd

following Australian Standards (AS 2856.1-2000; AS 2856.2-1998;


1998; AS 2856.3-2000).
2856.

54
Chapter 2

2.6.2. Downhole wireline measurements


Wireline logging has a long history in coal exploration, and was considered a quality

method for acquiring fundamental data like coal thickness and coal quality as well as

for identifying the presence of aquifers (Rogers, 2007). With decades of experience

measuring coal, wireline formation evaluation in coal is still nevertheless limited

compared to conventional oil and gas reservoir interpretation due to the complexity and

heterogeneity of coal itself. Today, wireline log measuring tools in CSG are used

ubiquitously to estimate net coal (or coal thickness) and coal density and also to

determine the orientation of master cleats and major multi-bed fractures as well as the

stress orientation (Mavor and Nelson, 1997, p. 199). In conventional reservoirs,

interpreting down-hole measurements can determine rock storage capacity (porosity)

and hydrocarbon saturation. In CSG the former can be estimated indirectly and only to

a limited degree, while the latter cannot be estimated, as methane content has no

significant effect on any downhole measurement tool. Nevertheless, as wireline

measurements have proven to be very reliable, they can be used to establish correlations

with core analytical data. In the next section coal measurement responses from

commonly used wireline logging tools are described and explained.

2.6.2.1. Nuclear measurements

Natural gamma ray

The natural gamma ray measurement tool run in all wells used in this study was a

Weatherford MSG tool, the Compact gamma ray (Weatherford - MCG, 2012)

(APPENDIX I tools and logs acronyms). The MSG measures bulk gamma rays

emitted from radioactive elements (thorium, potassium and uranium), close to the

borehore wall (Rogers, 2007) and in the borehole fluid. Normally coal has a low

55
Chapter 2

radioactivity (Figure 2.21), and therefore a low reading or gamma ray response

contrasts with neighbouring shaly zones that generally contain larger amounts of

naturally occurring radioactive mineral. As a result the gamma response is principally

used as a lithology indicator, estimating the thickness of coal seams, and helps in

situations where wash-outs occur and no reliable density readings exist. There are

nevertheless two points that make defining the top and basal boundaries of coal seams

challenging. Firstly, the gamma ray reading is an arithmetic mean of the formation

radioactivity, leading to a smoothing of the curve. Secondly, the density of the coal is

normally much lower than that of the surrounding rock types, creating a difference in

the depth and volume of investigation in coal and other sequences. In other words, a

larger volume of material contributes to the gamma ray reading in coal as opposed to

adjacent shales/sands, and this can shift the midpoint in transition from one formation to

another. This measurement is also considered inadequate for proximate analysis and

quantitative purposes due to its inability to distinguish low ash from high ash layers

within the coal seam and therefore cannot resolve thin beds (Mavor et al., 1990a).

Neutron-porosity

The neutron-porosity (Weatherford - MDN, 2012) tool responds to the hydrogen

content in the formation matrix (Ahmed et al., 1991). The readings of the neutron log

are therefore associated with the volume of organic content, clays and other minerals

containing hydrogen and fluids. Since coal has very high hydrogen content, the log

responds as an apparently very porous reservoir (very high values) (Figure 2.21).

Besides coal identification, neutron logs offer limited precision in coal, with low count

rates and significant noise due to the high hydrogen content of coal (Rogers, 2007).

56
Chapter 2

Density & Pe

Coal, as an organic rich rock, has properties very distinct from other sedimentary

formations. The organic composition gives coal a considerably lower density than other

sedimentary rocks such as mudstones, siltstones or sandstones (Figure 2.21).

Measuring density downhole (Weatherford - MPD, 2010) or in fixed laboratory

conditions is of special importance. The density log is a critical measurement that

directly affects the estimation of gas-in-place (2-2), as through its interpretation,

reservoir quality and thickness is estimated (Mavor and Nelson, 1997; Nelson, 1999).

Coal density is dependent of geological variables such as depositional environment and

the characteristics of the coal lithotypes including: rank, mineral and moisture content

and maceral type (Neavel et al., 1986; Unsworth et al., 1989; Walker Jr. et al., 1988).

The density tool is also associated with the photoelectric (PE) reading, which is

considered to be a good coal identification method (Rogers, 2007). The Pe values in

coal zones are usually very low, below 1.0 barns/electron.

2.6.2.2. Resistivity measurements

Coal, as for conventional oil and gas reservoir rocks, is an electrical insulator. If

these rocks are shown to have no porosity or are measured on a dry state, the expected

resistivity is over a million ohm.m (Hilchie, 1982). Coal has limited macroporosity,

making this rock highly resistive. The electrical resistance recorded depends not only

the type of material but also its geometric shape. To define the ability of a material to

obstruct electric flow in well logging the term used is R (resistivity). The metric unit

used in resistivity measurements is the ohm.m (ohm-metres2/metre). In downhole

measurements resistivity has a long history of being plotted using a logarithmic scale

because this property typically varies over several magnitudes. In CSG a common

57
Chapter 2

interpretation of downhole resistivity measurements relates to permeability variations

within the coal bed zone (Li et al., 2011; Mavor et al., 1990a). This is considered in

more detail in CHAPTER 4.

To measure resistivity downhole there are different devices based on different

designs, electrode sizes and spacings. Although they can vary on a well to well basis,

one or more resistivity tools are always run, providing resistivity measurements with

different depths of investigation into the formation.

The tool of choice is usually associated with the type and salinity of the borehole

fluid. It is general practice to run induction tools if the mud system is resistive or non-

conductive and laterolog tool if the mud is more conductive. Development and

production wells make up the bulk of a drilling campaign. These, with a larger borehole

diameter, are drilled with light water-based muds containing bentonite to improve

viscosity; due to this practice the induction assembly is the most common resistivity

tool run in CSG. The running of the induction assembly tool (Weatherford - MAI,

2012) is associated with another resistivity device, the high-resolution shallow focused

electric (MFE) (Figure 2.21).

Cored wells in a commercially viable CSG project are rare as they are significantly

more expensive. Drilling fluids with potassium chloride (KCl) are often used to

maintain borehole stability and, in cored wells where the main goal is to recover core

samples in the best conditions, KCl fluid systems tend to be used.. The saltier the

drilling fluid is, the more conductive it becomes, affecting the received signal in

logging devices more sensitive to conductivity. In these cases the resistivity tool run is a

focused resistivity logging system, like the Compact dual laterolog tool (Weatherford -

MDL, 2012), designed to be used in salt-based drilling muds.

58
Chapter 2

2.6.2.3. The spontaneous potential


The spontaneous potential (SP) log mainly resolves the salinity difference between

the borehole fluid and the natural reservoir fluid, and is also a non-quantitative indicator

of high permeability variations and thick coal beds (Figure 2.21). It is also influenced

by the streaming potential and the electrochemical invasion. The measurement is based

on a voltage potential difference, where the zero baseline reference is a mudstone value.

If the plotted log moves left of this baseline it indicates the salinity of the borehole fluid

is lower than the natural salinity of the formation, if it moves right of the baseline the

phenomenon is the opposite. Difficulties found in this interpretation are found in thinly

bedded and tight formations (Rogers, 2007).

2.6.2.4. Acoustic travel time


The sonic tool (Weatherford - MSS, 2012) measures formation compressional

slowness (the inverse of velocity). Sonic measurements in coal zones read long transit

times, usually longer than any other formations in the well. Readings will be affected by

coal strength and factors like temperature. Compensated acoustic logs have the

advantage of being developed to eliminate wash-out effects. In terms of specific well

logging readings, coal will give high travel time values. For example DT (conventional

compressional travel time log) value for lignite ranges from 130 to 150 s/ft, but less

than 120 for anthracite (Yao & Han. 2007) (Figure 2.21).

2.6.2.5. Electrical and acoustic imaging


Over the past two decades high resolution micro-imager type tools has been

increasingly used in CBM exploration. Tools like the Electrical microimager

(Weatherford - CMI, 2013) have an array of micro-resistivity buttons mounted on 4, 6

or 8 pads to provide a representative view (image) of the borehole walls. Acoustic

59
Chapter 2

scanners tools also generate images of the borehole by emitting ultrasound pulses and

recording the amplitude and travel time of the reflected signals. They give particularly

good information about coal thickness, identification of master-cleats and major

fractures, natural fracturing direction, present day stress determination and borehole

breakouts (Figure 2.21).

2.6.2.6. Caliper measurements


Several different sources of borehole diameter (caliper) measurements were

available in the data set. The caliper readings throughout this thesis refer to the density

caliper (Figure 2.21). The density tool measures the distance between the skid face,

where the source and the detectors are, and the backup shoe pushes the skid face against

the borehole wall. This distance is designated as density caliper and is a common

reference of borehole conditions and a washout indicator.

60
Chapter 2

Figure 2.21 Typical tool readings from wireline log measurements against core description in coal seam gas exploration. Example from a coal seam
in well 3.

61
Chapter 2

2.6.3. Drill stem tests (DST)

Experience has taken the CSG exploration industry to add significant importance to

permeability testing as a means to validate the economic potential of a prospect (Ried et

al., 1992). Today, permeability analysis in CSG is through production history data.

Nevertheless the dual porosity found in CSG makes the interpretation of pressure

transient tests difficult. The coal matrix has a very low permeability. It is believed that a

determining factor influencing the production rate is the natural fracture system and its

development (Weida et al., 2005; Yang et al., 2006).

From the first well drilled in a prospect area, which is usually cored, it is common

practice to evaluate permeability by performing a series of DST (drill stem test) tests on

selected coals seams (Kabir et al., 2011; Mavor et al., 1990a; Weida et al., 2005).

Transient pressure tests have a long history in conventional reservoirs and they are the

main source of permeability data, at least in the first stages of reservoir evaluation.

They are conducted immediately after the well has been drilled and logged with

wireline tools, before the well is cased. It is necessity to do this procedure as soon as

possible after the well has been drilled, when the coal beds are less damaged. Therefore

the DST test should be conducted in intervals before the total depth is reached. For

economic reasons this procedure may only be done when the well is totally drilled.

Before the actual DST test is run the well is required to be logged first and the

measurements interpreted, so the depth intervals for the test can be selected.

Technically, the isolation of coal seams is conducted by installing packers on the

bottom and top of the interval. From DST tests, it is possible to determine permeability,

produced fluid properties, reservoir pressure and skin damage (Earlougher, 1977;

equation 2-4). As for conventional DST testing, after the first clean up well pre-flow

62
Chapter 2

and shut-in stages, a longer period of flow is conducted. This main flow is an essential

component of the test, from where the permeability is estimated.

2-4
162.6   



Where:
 permeability, md
 water injection rate, bpd
 water formation volume factor, reservoir bbl/stb
 water viscosity, cp
 slope of the semilog straight line, spi/cycle

net coal thickness of the tested interval, ft

Permeability and porosity change significantly in coal productive zones near the

wellbore during production (Gu and Chalaturnyk, 2010). Another parameter that DST

testing produces is skin values. If by any means possible the formation is damaged or

highly fractured, and induces an increment or reduction of the natural permeability, the

skin factor can be negative or positive respectively (Laubach and Tremain, 1991). This

estimation of formation damage is determined from pressure build-up analysis (Dake,

1983). Skin with high positive or negative values can be caused by a variety of issues,

from non-Darcy flow of formation water, drilling mud, fines and gas flow, drill bit

action, tortuous flow paths and anisotropic cleat spacings, effect of horizontal stress

anisotropy with directional permeability and other geomechanical processes (Kabir et

al., 2011).

63
Chapter 2

2.7. Study area


The data used in this study were provided by Queensland Gas Company (QGC is

part of BG group). Well location can be seen in Figure 2.22. This data set provided a

reasonably good representative sampling of the Walloon Sub-group CSG exploration

NW-SE fairway, a northeast-southwest trend where this Sub-group is the thickest

(Figure 2.23). The coals from the Walloon Sub-group were targeted for this study due

to the complete dataset including todays standard measurement in CSG.

Part of the Great Australian Basin the Surat intracratonic Basin covers an area of

300,000km2 in the south-eastern Queensland and northern New South Wales. From

centre to its margins it overlies the Permo-Triassic Bowen Basin and the Early

Devonian to Late Triassic crystalline basement (Martin et al., 2013; Figure 2.24;

Figure 2.25). After extensive erosion in the region during the Triassic, the Surat Basin

developed during a massive depositional phase due to subsidence driven by subduction

along the eastern margin of the Australian plate. It consists of 2,500m sub-horizontal

sedimentary formations of fluvial-lacustrine deposits laid down in the late Triassic to

the early Cretaceous. The Surat Basin has its maximum thickness in the Mimosa

syncline. The sequence K (Hoffmann et al., 2009) contains the coal of the Walloon

Sub-group as part of a series cycles of grain sizes fining upwards with faults being

generally small, localized with throws of 5 to 20 metres (Exon, 1976; Jones and Patrick,

1981; Scott et al., 2007).

Found throughout the Surat Basin, overlying transitionally the Durabilla Formation,

the Middle Jurassic Walloon Subgroup with a variable thickness between 50 and 700m

(Hoffmann et al., 2009) is made of tight, fine grained argillaceous sandstone, siltstone,

mudstone, coal and rare limestone and ironstone. The coal beds are believed to have

been deposited in poorly drained floodplains with the exception of the base of the

64
Chapter 2

formation, where rare coal beds indicate stream deposition (Exon, 1976) during a

greenhouse period when the Australian continent was at higher latitudes (Bradshaw

and Yeung, 1990).

The Walloon Sub-group today is considered to be a Sub-group divided by

lithological similar Juandah (upper and lower) Coal Measures, Tangalooma Sandstone

and Taroom Measures (Smyth and Cook, 1976). The correlation of these measures has

proved difficult to both academics and industry due to variable thickness and complex

cross-sectional geometry (Calvert et al., 2011; Fielding, 1996; Jones and Patrick, 1981;

Lagendijk and Ryan, 2010). Increasing thickness is associated with the less erosion due

thinning of overlying Springbok Sandstone. To aid the correlation, the measures are

subdivided into regionally correlatable coal seam or coal groups (Kabir et al., 2011;

Figure 2.26). Upper Juandah is then divided into the Kogan, Macalister - upper and

lower - and Nangram. The lower Juandah is subdivided into Wambo, Iona and Argyle

coal seams. In this study the QGC/BG Group operator non-formal subdivision of the

Taroom Measure was used, including Auburn, Bulwer and Condamine coal seams.

Volcanic debris and montmorillonite (bentonites) are also found forming thin tuff beds.

These sediments lead to some debate about their origin due to volcanic activity in the

subducting plate margin to the east of the Basin, as suggested by Exon (1976), or

because of intra-basinal volcanic activity related to the early stages of rifting of the

south-eastern Australian continental margin stated by (Fielding, 1996).

65
Chapter 2

a)

b)

c)

Figure 2.22 Location of the well data in the Surat Basin in Queensland Australia.
c) Detail of the Well 3 field location where besides the cored well was also acquired data
from DST production wells Well 3a, 3b and 3c.

66
Chapter 2

Wells
Field

10

12

2
13

11

4
8
9

WELL 3 field

14

Figure 2.23 Well location in Walloon Subgroup depth (mSS) structure map. Modified from
QGC Pty Limited (2012a) report.

67
Chapter 2

Field

Burunga
Comet ridge Fault

10 Trelinga
Anticline
Hutton- 12
Wallumbilla
Fault 7

13 2

11

Miles Fault Undulla nose

4
Kogan
8 Anticline
9

6
Leichhardt
fault
5

14

Roma Shelf Moonie fault

Seismic section (fig.2.25)

Tingan fault

Figure 2.24 Major tectonic elements of the Surat Basin and well location. Modified from
QGC Pty Limited (2012a) report.

68
Chapter 2

Se
Figure 2.25 Seismic section located as indicated in figure x. From (QGC Pty Limited, 2012b)

69
Chapter 2

Figure 2.26 Litho-stratigraphy of the Walloon Sub-group from (Martin et al., 2013).

In Western Australia CSG exploration has been growing at a fast pace since early

1990s. According to the Australian Government, this country came from no significant

CSG production in 1997 to reach around 6,000Mm3 in 2011, from its extensive

underground coal reserves estimated to be 900,000Mm3 (Bradshaw et al., 2012; Leather

et al., 2013; Queenslands coal seam gas overview, 2013, Figure 2.27). With the first

well for CSG only drilled in 1995, the Surat Basin alone is believed to have (2013

estimate) reserves of more than 650,000Mm3 of methane. Methane production started in

2005, by 2012 had already overcome the production of the Bowen Basin.

70
Chapter 2

The relatively shallow, quick, cheap and of easy completion of most of these wells

made the number of wells drilled per year reach a record number of 720 in 2011.

Moreover, there are three CSG-to-LNG terminals under construction in Curtis Island.

These will the first ever in the world to be built to export produced CSG.

Figure 2.27 Queensland Australia proved and probable reserves estimated in 2013 (Source
Queensland Government, Department of Natural Resources and Mines).

2.8. Studied Dataset

The data used in this study was provided by QGC/BG group operator. It provided a

reasonably good representative sampling of the Walloon Sub-group CSG exploration

NW-SE fairway. The coals from the Walloon Sub-group were targeted for this study

due to the complete dataset of what is todays standard measurement in CSG. The

dataset comprises 14 cored wells and 3 production wells, the latter being part of the

evaluation of the area where well 3 was first drilled.

The core dataset is summarised in Table 2.5 to Table 2.7. With most of the

individual cored well datasets having desorption, adsorption, maceral and reflectance

analysis, and also descriptions and photographs of the core. Critical to this work, was

acquiring the individual coal seam depth intervals; this work assumes these to be

71
Chapter 2

complete and correct. All cored and production well datasets were provided with DST

test results.

The wireline dataset was provided by Weatherford UK after permission was granted

by the operator. With the exception of well 3C, all the remaining wells have wireline

log data. Most of the wireline log dataset was acquired in the original raw, unfiltered,

unsmooth version. Wells 4, 7 and 9 logs were only available in plotted version and

therefore these were impossible to be reprocessed. All the individual well data have in

common caliper, natural gamma ray, bulk density, neutron photoelectric factor (Pe),

compressional sonic velocity measurements. Standard resistivity and image logs

availability are well dependent (Table 2.7).

Minimal depth matching between core and wireline logs was needed, in other words,

although many coal beds needed depth shifting this was only of the order of a few

centimetres. Quality analysis of wireline log data identified multiple washout zones in

multiple wells, especially in well 1. In terms of QC/QA, only MMR tool readings in

well 13 were considered completely unreliable due to a lack of character of the log and

mismatching another reading of the same tool in well 5.

72
Chapter 2

Unit/Well 1 2 3 4 5 6 7 8 10 11 12* 13 14
Springbok Sdt. - - - - - - - - - - - 2 -
Kogan - 2 - 2 2 6 1 - - 2 - 3 -
Macalister Upper - 3 4 4 7 7 8 12 - 4 11 6 -
Macalister Lower - - 2 5 3 2 13 4 - 12 6 5 3
Nangram - - - 3 1 2 1 5 3 1 - 3 2
Wambo 4 9 4 1 2 1 5 5 2 2 9 4 4
Iona 2 3 6 2 6 1 8 4 6 3 14 3 3
Argyle 7 6 7 4 9 9 5 2 7 8 5 9 14
Tangalooma Ss. - - - 2 2 1 - - - - 7 4 -
Auburn 2 1 4 1 7 5 - - - 2 7 4 6
Bulwer 2 2 3 2 11 1 7 1 4 5 5 3 3
Condamine 5 12 8 7 7 7 3 2 5 8 14 12 10
Well total 22 38 38 33 57 42 51 35 27 47 78 58 45
Total 571
Table 2.5 Number of core samples per coal seam from each well.
*The coal seam location of the samples in this well was estimated through correlation with well 2, the nearest.

Data/Well 1 2 3 4 5 6 7 8 9 10 11 12 13 14 3A 3B 3C
Desorption analysis             
Maceral analysis            
Adsorption analysis            
Reflectance analysis            
Petrology report 
Chemostratigraphy report 
Core description             
Coal Seam depths Intervals               
Core photos             
Core HD photos   
Well Completion report   
DST results                 
DST report 
Table 2.6 Summary of the core analysis and DST test results dataset from each well.

73
Chapter 2

Log Standard Resistivity Logs Imager


Well Well type Year
format MAI MFE MDL MMR CMI AS ATV
1 Cored 2006 Raw  
2 Cored 2009 Raw      
3 Cored 2008 Raw    
3A Production 2009 Raw  * 
3B Production 2009 Raw  * 
3C Production 2009 ----
4 Cored 2006 Plotted
5 Cored 2008 Plotted   
6 Cored 2009 Raw 
7 Cored 2008 Plotted   
8 Cored 2010 Raw  
9 Cored 2010 Raw  
10 Cored 2009 Raw  
11 Cored 2010 Raw  
12 Cored 2010 Raw   
13 Cored 2009 Raw   
14 Cored 2009 Raw  
Table 2.7 Summary wireline log data used in this study.

Where:
AS Acoustic Scanner
ATV Acoustic Televiewer
CMI Compact Micro Imager
MAI Compact Array Induction
MDL Compact Dual Laterolog
MFE High-resolution shallow focused electric; (*) with sleeve
MMR Micro Laterolog

74
Chapter 3

CHAPTER 3 Coal lithotype properties of

the Walloon Sub-group, in Queensland,

Australia

3.1. Introduction

CSG exploration targets coal seams with distinct properties from the ones mined or

from conventional gas reservoirs. These coals are generally thinner, more vertically

spaced and less continuous horizontally. To populate the attributes in CSG reservoir

simulation, the operator needs to support pre-drilling surveys with many more wells

compared to conventional gas resources (Sung and Ertekin, 1987; Weida et al., 2005).

CSG is highly dependent on exploratory cored wells. Special core analysis can clarify

fundamental parameters related to the gas yield in the coal beds. Nonetheless, for

economic reasons, only a limited number of wells are cored. Wireline logs are usually

the most abundant data set acquired, and to which core is related to, enabling

correlation between wells. In exploration and appraisal wells open hole pressure

transient tests, also called drill stem tests (DST) also contribute to data of high

importance, which is absolute permeability (Kabir et al., 2011).

In Humic coal, it is possible to differentiate macroscopic layers designated as

lithotypes. The alternation of these structures within a coal seam indicate different

physical and chemical characteristics reflecting changes of the original plant

communities and mineral content (Cao et al., 2011; Crosdale, 1995). These major

changes in coal are then expected to influence not only the capacity to store gas but also

the way the gas flows (Clarkson and Bustin, 1997). Furthermore, it is well established

75
Chapter 3

that coal rank is a major factor affecting coal characteristics, together with coal

composition and coal bedding (Bustin and Clarkson, 1998; Laxminarayana and

Crosdale, 1999). Lastly, as demonstrated in this study, the heterogeneity associated with

the coal seam nature, in the shape of coal facies distribution, is also reflected in the cleat

network development.

This chapter examines how core analysis, wireline measurements and DST test

results reflect the bulk properties and methane producibility of coal lithotypes from the

Walloon Sub-group, illustrating the degree of heterogeneity associated with coal beds

and discuss why they still constitute a challenge to reservoir evaluation.

3.2. Methodology

3.2.1. Original data set

The basis for this chapter is the examination of coal macroscopically and how it is

reflected in terms of composition, wireline log responses and DST permeability

estimation. This work was carried out by interpreting a previously produced data set

acquired from common procedures in CSG exploration by the operator. It comprises a

total of 13 wells, biased by what are todays common practices, not only on which

samples were selected to be analysed but also how the measurements were taken. The

entire Sub-group was cored totalling 5676.3 metres across all the wells, with 99.4%

recovery rate. To raise the core to the surface, a slimhole wireline coring system was

used to minimize core damage and gas loss. 491 core samples were cut into 0.3m to

1.0m long, 61.5mm diameter. These were then described and tested by Earth Data Pty

Ltd Geological & Earth Science consultants for rock type, and in the case of coal, coal

lithotype, cleat spacing, cleat length and cleat mineral filling. The same company also

measured desorbed gas and provided proximate analysis. Composite samples for

76
Chapter 3

maceral, reflectance and adsorption isotherm analysis were tested by Energy Resources

Consulting Pty Ltd. Also, analysed from the same wells were 42 individual DST

permeability tests conducted by the operator.

From the well dataset 12 wells were selected that had wireline log data and core

descriptions. The wireline logging data analysed were resistivity (induction, shallow

focused electric and laterolog), nuclear (gamma ray, density, Pe, neutron porosity),

spontaneous potential, caliper measurements, resistivity imager and acoustic scanner

logs.

A recent publication by Martin et al. (2013) distinguishes the Walloon Subgroup in

three main facies associations: primary channel, overbank and volcanic extrusives. The

coal facies, designated as part of the overbank facies association, are distinguished by

the depositional styles or classes and are part of Class I - decimetre stacked coal beds up

to 8m thick intercalated by sands, claystones and volcanic tuffs - or Class II - lacustrine

originated less than 1.0 metres thick beds coarsening upwards. For the purpose of this

study, such distinction was not made, as the focus is on coal bed lithotype

characteristics and not on the coal seams (group) they are part of. The coal seams

comprise a sequence of different coal lithotypes facies, with transitions to mudstones

with variable organic matter and volcanic tuff layers, boundary by very fine to medium

grained sandstones, siltstones and siderite. Transitions between the beds are sometimes

sharp.

3.2.2. Macroscopic description

For this study several characteristics of the coal beds were individually and

collectively analysed statistically. The procedure of testing samples regardless of their

coal lithotype composition created difficulties defining unique characteristics for the

77
Chapter 3

different types of coals. Furthermore, core samples tested in many cases are shown to

be interbedded with other rock types. The original data set was then reorganized so that

it could reflect the nature of each lithotype. Characteristics for each lithotype analysed

include distribution, thickness, maceral and proximate analysis composition, cleat

development and filling, gas yield and gas saturation.

Published research for coal lithotypes associates brighter coal with higher vitrinite

content, due to the generally low liptinite content found in coal samples. Classifying as

bright if the coal is vitrinite-rich or dull if inertinite-rich (Chalmers and Marc Bustin,

2007). In the Surat Basin, coal is dominated by vitrinite, however, it also contains a

high percentage of liptinite. The description (Table 3.1) of the coal samples was

performed following the Australian standard (AS 2916-2007) lithotype classification.

Coal Lithotype Description


Bright >90% Bright Coal, occurring mainly as thin bands
Bright banded 60% to 90% Bright coal, with Dull bands
Dull and bright Similar number of Bright and Dull bands
Dull Banded Dull coal with numerous Bright bands, 10% to 40% Bright coal
Dull minor bright Dull coal with rare Bright bands, 1% to 10% Bright coal
Dull Dull coal, <1% Bright Coal
Stony coal Impure coal
Table 3.1 Macroscopic description of coals based on the Australian Standard.

3.2.3. Sample selection

At a macroscopic scale it was possible to identify the heterogeneous nature of the

coal seams of the Walloon Sub-group. The vertical and horizontal variability of coal is

often associated with the sources of error in the Gas-in-place estimation (Lagendijk and

Ryan, 2010). The beds that constitute a coal seam have different physical properties, but

because of their limited thickness it is impractical to sample every single one separately.

Depending on the coal characteristic analysed a different source of data was interpreted.

For macroscopic characteristics of the lithotypes (frequency, thickness, cleat

78
Chapter 3

development, cleat filling) the entire extension of macroscopically described coal was

used; a total of 366.8 metres of cored coal, distributed over 2374 individual coal beds,

described with a spacing of 1 cm (APPENDIX A Seam and net coal thickness). To

investigate the properties determined through gas desorption, proximate and maceral

analysis characteristics were selected for non-broken single lithotype samples, totalling

136 core samples where each sample was analysed individually. The interpretations of

reflectance and isotherm analysis have an associated difficulty due to the fact that these

tests were carried out on crushed coal seam composite samples.

3.3. Coal bed Thickness

At the macroscopic scale, lithotypes can be seen as independent parts of a coal seam.

A sequence with different lithotypes is then associated with major changes not only in

the depositional system but also the physical and chemical characteristics of the coal

itself, including gas yield (Clarkson and Bustin, 1997; Holdgate et al., 1995; Mares and

Moore, 2008). The distribution of the observed lithotype thickness is displayed in

Figure 3.1. The histogram highlights the thinly bedded nature of the coal seams. Core

macroscopic description has indicated that each individual coal bed lithotype has a

thickness mean of 0.15m, ranging from coal laminae (<1cm) to a bed as thick as 2.3m.

Observations indicate that thickness increases slightly with depth (Table 3.2) having

nevertheless a wide range of variability indicated by the standard deviation.

79
Chapter 3

350 100%

300
80%
250
Frequency

60%
200

150
40%

100
20%
50

0 0%

Thickness (metres) Cumulative %

Figure 3.1 Coal lithotype thickness frequency and accumulated frequency.

Standard
Coal Seam Count Max (m) Mean (m)
Deviation (m)
Kogan 62 0.47 0.14 0.11
Macalister Upper 148 1.35 0.15 0.18
Macalister Lower 176 1.00 0.16 0.14
Nangram 94 0.58 0.12 0.10
Wambo 162 0.78 0.16 0.14
Iona 194 0.70 0.15 0.13
Argyle 368 0.90 0.17 0.16
Tangalooma Sandstone 55 0.38 0.14 0.10
Auburn 83 0.78 0.17 0.13
Bulwer 119 0.90 0.18 0.18
Condamine 346 1.05 0.17 0.17
Table 3.2 Coal bed thickness found in the 10 cored well in the Walloon Sub-group.

80
Chapter 3

In Figure 3.2, the thickness of each lithotype indicates similar bimodal distributions

in Stony and Dull lithotypes at 0.025m and 0.1m. Dull minor bright and Dull banded

show normal distributions centred between 0.1m and 0.2m. Dull & Bright coal has a

dominating mode between 0.1 and 0.2m. Bright banded coals have a particular curious

distribution, increasing from 0.8m to 0.1, dropping at 0.05m and reaching the highest

mode in samples under 0.025m. Bright lithotypes besides being rare are much thinner

relatively to other lithotypes.

Figure 3.2 Lithotype thickness histograms distributions from the description of core from 10
cored wells.

81
Chapter 3

3.4. Lithotype distribution

The Dull minor bright is the most common lithotype (Figure.3.3) dominant in all

Walloons coal measures, followed by dull-banded coals. Overall the Walloon Sub-

group analysed in this study indicates that these are prominently dull, generally having

limited bright components distributed in the shape of bands or dispersed fragments.

Furthermore, bright-dominated coal beds are rare making less than 3% of all net coal

found. Dull minor bright coals vary between 35% and 62% of the total coal beds found

in each seam, having two peaks in the Macalister Upper and Iona. Dull-banded coal is

accounted to be between 21% and 33% of all coal in Macalister Lower and underlying

coal seams. Coal in the Kogan and Macalister Upper seams is especially dull, where

besides dull minor bright lithotype, Dull and Stony coals make significant proportions.

Coals with brightness above that of the Dull & Bright coals are generally in low

abundance with numbers under 10% in all coal seams.

The thin nature of the Bright and Bright banded coals, together with the limited

number of these facies, constrains the number of samples analysed in special core

analysis. These lithotypes when found in core samples are associated with other facies

or are distributed throughout the well in the form of isolated, very thin coal beds (0.05

m). As a consequence, further analysis, now focussed on core analysis, is restricted to

the lithotypes that make up the bulk of the Walloon Sub-group, i.e., Dull minor bright,

Dull banded, Dull, Stony and Dull & Bright.

82
Chapter 3

Figure.3.3 Variation of lithotype thickness with depth for Walloon Sub-group.

3.5. Petrography and Geochemistry

3.5.1. Rank

Reflectance analysis was performed by Energy Resources Consulting Pty Ltd over

composite samples (i.e. a blend of individual samples taken from the same coal seam).

The system used was a Leitz MPV1 to determine maximum reflectance (Ro max). The

Walloon Subgroup coal seam sequences drilled and used for this work have a range in

rank from 0.375, found in the Macalister Upper Coal Seam, up to 0.65% in Condamine

coal beds. This, together with proximate analysis, moisture and volatile matter content

results, indicates that the rank classification of the Walloon Sub-group is between sub-

bituminous and high-volatile bituminous. Note that, similar to Hamilton et al (2012),

the vitrinite reflectance measured increases only generally with coal seam depth. This

83
Chapter 3

can be observed if the wells are analysed individually (Table 3.3) or all together

(Figure 3.4).

Coal Seam/Well 1 2 3 4 5 6 7 8 10 11 13 14
Kogan --- --- --- 0.51 0.49 0.43 0.49 --- --- 0.45 0.40 ---
Mac.Up. --- 0.37 0.46 0.51 0.45 0.44 0.48 0.5 --- 0.46 0.44 ---
Mac. Lo. --- 0.39 0.47 0.47 0.46 0.40 0.47 0.51 --- 0.44 0.42 ---
Nangram --- --- --- 0.49 0.45 0.43 0.46 0.55 0.42 0.42 0.45 0.45
Wambo 0.48 0.41 0.45 --- 0.45 0.44 0.46 0.56 0.45 0.44 0.44 0.41
Iona 0.49 --- 0.45 0.53 0.43 0.43 0.46 0.57 0.46 0.45 0.44 0.43
Argyle 0.5 0.40 0.46 0.54 0.47 0.46 0.45 0.58 0.48 0.46 0.47 0.42
Tang. Ss. --- 0.43 --- 0.52 0.44 0.47 --- --- --- --- 0.46 ---
Auburn 0.51 --- 0.52 --- 0.46 0.47 --- 0.63 --- 0.47 0.50 0.44
Burwer 0.52 0.45 0.48 0.56 0.46 0.48 0.48 --- 0.52 0.49 --- 0.47

Condamine 0.55 0.47 0.49 0.61 0.48 0.49 0.56 0.65 0.53 0.51 0.48 0.46
Table 3.3 Vitrinite reflectance in 12 wells in the Walloon Sub-group coal seams provided.

Figure 3.4 Mean vitrinite reflectance in 12 well in the Walloon Sub-group coal seams.

84
Chapter 3

3.5.2. Proximate Analysis

Proximate analysis was performed according to the Australian Standard (AS1038.3-

2000) (APPENDIX B proximate and desorption analysis). This procedure is

particularly sensitive to core selection bias, by inducing an increment in the values of

mineral content (i.e. ash yield). Therefore, the selected samples considered had less than

50% ash yield. Fixed carbon increases with brightness, while volatile matter having a

similar behaviour seems to reach a plateau in coals brighter than Dull & Bright coal

facies. Moisture decreases from duller coals until it stabilizes at about 6% (as received)

in coals brighter than Dull Banded. These results are show on Table 3.4, safe to

conclude that brightness in coal is associated with the increase of organic content, and

opposite over the ash yield.

Lithotype: Bright banded (%, ar) Lithotype: Dull & Bright (%, ar)
Proximate Standard Proximate Standard
Min Max Mean Min Max Mean
Analysis Deviation Analysis Deviation
Fixed Carbon 36.9 44.3 40.8 2.9 Fixed Carbon 30.0 44.2 39.4 5.2
Volatile Matter 40.2 43.6 42.0 1.5 Volatile Matter 33.5 46.3 42.0 4.5
Ash 8.3 14.5 11.4 2.3 Ash 6.0 31.4 12.5 9.5
Moisture 5.1 7.0 5.9 0.7 Moisture 5.0 7.7 6.1 1.0
Samples Count 6 Samples Count 7

Lithotype: Dull banded (%, ar) Lithotype: Dull minor Bright (%, ar)
Proximate Standard Proximate Standard
Min Max Mean Min Max Mean
Analysis Deviation Analysis Deviation
Fixed Carbon 25.2 44.3 36.0 4.7 Fixed Carbon 21.2 45.6 33.8 5.2
Volatile Matter 30.0 46.1 39.7 4.0 Volatile Matter 31.0 48.6 39.2 4.5
Ash 5.2 37.0 18.1 8.6 Ash 3.7 42.4 20.4 9.3
Moisture 2.1 9.7 6.2 2.1 Moisture 2.3 11.0 6.6 1.8
Samples Count 30 Samples Count 82

Lithotype: Dull (%, ar)


Standard
Proximate Analysis Min Max Mean
Deviation
Fixed Carbon 15.6 42.3 31.0 13.8
Volatile Matter 24.6 41.7 35.1 9.2
Ash 5.8 49.5 26.1 24.6
Moisture 6.2 10.2 7.8 2.1
Samples Count 3
Table 3.4 Proximate analysis data from selected coal samples dominated by one single
lithotype and < 50% ash yield from Walloon Sub-group coal beds.

85
Chapter 3

3.5.3. Petrographical Analysis

Reflectance analysis using the standard Australian maceral techniques (AS2856.2-

1998) and classification (AS2856-1986) indicates that the coal beds are consistently

high in vitrinite, with mean of 72% (mineral matter free) being predominantly

telocollinite (mean 39.4% m.m.f.), desmocollinite (mean 27% m.m.mf.) and

corpogelinite (mean 6% m.m.f.). Liptinite is the second most common maceral with

23% (m.m.f.), being mainly suberinite. Inertinite is generally rare with mean 4%

(m.m.f.). The patterns indentified in terms of maceral content are generally in

agreement with Scott et al.(2007).

Petrographic analysis in single lithotypes (Table 3.5) from selected samples show

that mean Vitrinite content increases with brightness, with a sharp decrease in Bright

banded coal. Mean Liptinite content also increases with brightness, with a minor

decrease in dull-banded coal facies. Inertinite has only been found as traces (<2%).

Detailed maceral composition also indicates that in mineral matter free basis (m.m.f.),

telocollinite is common in all lithotypes ranging from 40% to 46%. Desmocollinite is

common in Dull samples (35%), Dull & Banded, Dull banded and Dull minor bright

lithotypes have similar content (24% to 26%), while Bright banded is not as common

(11%). Corpogelinite is on mean approximately 13% in Bright banded coal and

decreases as the samples became duller. Mineral matter decreases significantly as the

samples became brighter and are usually of clay, and more rarely, of carbonaceous

composition.

86
Chapter 3

Lithotype: Bright banded (%) Lithotype: Dull banded (%)


Maceral Standard Maceral Standard
Min Max Mean Min Max Mean
Analysis Deviation Analysis Deviation
Vitrinite 44.4 72.2 56.6 11.5 Vitrinite 29.1 81.0 57.7 14.7
Liptinite 17.3 40.9 30.5 11.2 Liptinite 6.1 32.2 15.0 6.5
Inertinite 0.2 1.4 0.7 0.4 Inertinite 0.0 12.4 1.6 3.3
Mineral Matter 8.2 19.8 12.3 4.6 Mineral Matter 2.7 61.4 25.6 16.2
Samples Count 6 Samples Count 30

Lithotype: Dull & Bright (%) Lithotype: Dull minor Bright (%)
Maceral Standard Maceral Standard
Min Max Mean Min Max Mean
Analysis Deviation Analysis Deviation
Vitrinite 37.0 81.5 66.1 17.2 Vitrinite 15.8 86.4 54.1 13.9
Liptinite 12.9 29.9 20.1 7.3 Liptinite 4.0 45.9 18.9 8.9
Inertinite 0.0 1.2 0.7 0.6 Inertinite 0.0 18.1 1.2 2.9
Mineral Matter 4.2 33.1 15.8 15.8 Mineral Matter 1.0 60.6 25.8 15.6
Samples Count 7 Samples Count 82

Lithotype: Dull (%)


Maceral Standard
Min Max Mean
Analysis Deviation
Vitrinite 29.7 60.6 45.2 21.8
Liptinite 10.5 14.7 12.6 3.0
Inertinite 0.0 0.4 0.2 0.3
Mineral Matter 24.2 59.8 42.0 25.2
Samples Count 3

Table 3.5 Maceral analysis data from selected coal samples dominated by one single lithotype
from Walloon Sub-group coal beds.

3.6. Cleat system

When the reservoir pressure is lowered, within a gas saturated coal, diffusion is

triggered in the matrix. Following the concentration gradient, gas molecules move from

pore to pore until they flow out to the borehole through the fracture network (Levine,

1996; Wang and Ward, 2009). Fractures occur in coal with variable spacing, length and

filling. When present in coal, cleats can be of two types both perpendicular to the

bedding plain. The dominant facture set is designated as face cleat, and these strike

perpendicular to butt cleats. Free gas in the cleat/fracture system is usually disregarded,

as it is within the coal matrix that the gas is mainly stored. In contrast, the fracture

system itself has a fundamental role, as hydrocarbon and water flow to the well through

it.

87
Chapter 3

Description of the cleat patterns has identified a relationship with coal type. Cleats

may or may not be present in coal. If present, the spacing between them is commonly

uniform throughout the bed. Face cleats, which generally have the closest spacing, are

considered to be the first to develop followed by the butt cleat. Analysis indicated that

each individual coal bed has particular cleat development properties. Four

characteristics were analysed in relation to face and butt cleat development for each

lithotype: (1) the number of coal beds with cleats; (2) the number of cleats that each

sample contains; (3) the length of the cleat as percentage of penetration of the core; and

(4) percentage of cleat filling. The mean spacing shown in this analysis may or may not

be representative of the cleat development much beyond the wells radius, depending

on the beds horizontal continuity, although it has been recognized in previous studies

that patterns can be uniform at a regional scale. Furthermore, several studies have also

indicated relationships between cleat spacing and rank, coal lithotype and bed thickness

(Dawson and Esterle, 2010; Laubach and Tremain, 1991).

3.6.1. Lithotype cleat attributes

Cleats when present are very close together, in the order of millimetres, therefore it

is possible to distinguish hundreds at the coal core scale. Coal, if fractured, can have

cleat spacing as low as 0.8mm (mean 7mm) in face cleats and 0.9mm (mean 16mm) in

butt cleats. Coals with high mineral content have shown a trend of having fewer cleats

then others. Ash rich coals, i.e., Stony coals are shown to be less fractured, followed by

Dull lithotypes coal beds (Table 3.6). Interestingly, Dull lithotypes having more

samples with cleats, when present these occur lesser numbers than even Stony coals.

88
Chapter 3

Coal Samples Mean Cleat Spacing Mean Cleat Penetration Mean Cleat Filling
Coal Lithotype bed with cleats Face cleat Butt cleat Face cleat Butt cleat Face cleat Butt cleat
count (%) (cleats/m) (cleats/m) (%) (%) (%) (%)
Bright 25 84% 7.7 16.0 89.8 89.6 39.1 33.1
Bright banded 75 80% 6.6 12.5 75.6 73.0 32.6 28.3
Dull and bright 148 88% 4.0 6.9 64.0 53.6 27.8 20.5
Dull Banded 532 86% 3.9 6.6 46.6 37.8 26.9 24.8
Dull minor bright 868 82% 5.3 7.2 40.9 29.8 26.8 22.9
Dull 374 49% 14.1 23.9 71.3 72.9 20.3 18.0
Stony coal 252 25% 7.0 9.0 63.7 53.9 20.2 21.7
Table 3.6 Variation of lithotype cleat mean frequency, length and filling from 2374 Walloon
Sub-group coal beds.

No direct trend was found between cleat development and coal bed thickness, but a

trend was found between cleat development and coal lithotype and coal lithotype

banding. Banded coals are shown to be more fractured than more homogeneous coals.

Banded lithotypes
types have similar numbers of cleated coal beds, while Dull banded and

Dull and bright coal beds have the higher numbers of cleats, followed then by Dull

minor Bright.. Because the data were


w analysed by lithotype bed, it is possible to analyse

the number of cleats found as a function


fu of lithotype thickness.
ickness. As thickness increases

so does
es the number of cleats for any given lithotype (Figure 3.5).

1.0

0.8
Bed Thickness (metres)

0.6

Linear (Stony)
0.4 Linear (Dull)
Linear (Dull minor bright)
Linear (Dull Banded)
0.2 Linear (Dull & Bright)
Linear (Bright Banded)
Linear (Bright)

0.0
0 100 200 300 400 500 600 700
Cleat frequency

Figure 3.5 Trends


ends between bed thickness and cleat frequency in each lithotype from 2374
Walloon Sub-group coal beds.

89
Chapter 3

Cleat development seems to be independent of coal type. The relationship between

the number of cleats per bed with rank have shown very poor results, with no particular

trend identified, however cleats appear to increase with depth (Figure 3.6).

Cleat length appears to be associated with coal homogeneity. The fewer bands the

coal has, the longer the cleats are. Observations generally indicate that for face cleat

filling, as brightness increases, carbonaceous and clay mineral filling in the cleats

increases as well. Therefore the mineral filling could be associated with cleat length.

Cleats are prone to present carbonaceous or clay mineral filling. Of the coals analysed,

70% of the beds with cleats demonstrated some degree of filling in the cleats, although

only 2% were completely closed by mineralization.

200

150
Cleats/metre

100

50

0
0 100 200 300 400 500 600 700 800 900 1000
Depth (metres)
Face Cleats Butt Cleats
Log. (Face Cleats) Log. (Butt Cleats)

Figure 3.6 Cross-plot of depth versus mean cleats per metre from 2374 coal beds in
Walloon Sub-group coal seams.

90
Chapter 3

3.7. Methane producibility

The two main factors that define the economic value of a CSG production are gas

volume in place and permeability. Together with the net coal thickness these are the

parameters most difficult to predict while evaluating a CSG project (Lagendijk and

Ryan, 2010). In conventional reservoirs one of the risks of investing in drilling is that

the reservoir could be dry. This risk is not associated with CSG, but instead the risk is

if the reservoir can produce methane in commercial quantities in a specific period of

time (Nelson, 2000b).

3.7.1. Gas content and saturation

Samples were analysed for total gas desorbed (desorbed gas + lost gas + residual

gas; Figure 3.7), gas saturation and gas diffusivity. Gas composition throughout all the

wells was 98% methane, with remaining gases being nitrogen, carbon dioxide and

ethane. The total desorbed gas is shown to be highly variable, with a mean 3.7 (7.0)

m3/tonne, with a minimum of 0.15 (1.1) m3/tonne and maximum of 15.2 (44.9)

m3/tonne as analysed (bracketed numbers indicate d.a.f. figures). Analysis of the data

indicated that most of the values above 12 m3/tonne on a d.a.f. basis are the result of

high gas in a core sample with lower organic content. This explains why the huge

difference between as analysed and d.a.f. basis maximum value. The data analysed

here are in agreement with previous studies over the Walloon Coal Sub-group that gas

content trends increase with depth in the Juandah and decrease in the deepest coal of the

Taroom measures (Hamilton et al., 2012; Scott et al., 2007) (Figure 3.8). Hamilton et

al. (2012) were not able to derive a single gas trend but yet three possible explanations

as to the gas variability in the Walloon Coal Measures. These were gas migration

upwards from deeper Walloon coals or from Bowen basin; methane secondary

91
Chapter 3

biogenesis in Juandah Measures;


Measures and/or early differential thermogenic gas generation

from liptinite macerals.

81%

9%
10%

Q1 - Lost Gas Q2 - Desorbed Gas Q3 - Residual gas

Figure 3.7 Mean percentage of


o Lost, Desorbed and Residual of the total gas yield in all 491
samples.

Gas content (d.a.f.) (m3/tonne)


0 2 4 6 8 10 12 14 16 18 20
0
Upper Juandah samples
Lower Juandah samples
100 Tangalooma Sandstone
Taroom samples
200

300
Depth (metre)

400

500

600

700

800

900
Figure 3.8 Gas content (d.a.f.) versus depth.

92
Chapter 3

Gas is stored in coal by adsorption, which is highly sensitive to coal maturity,

available organic content and pressure. Gas-in-place is usually estimated by having coal

seam grouped in close vertical proximity, similar rank, reservoir pressure and sorption

capability (Calvert et al., 2011; Mavor and Nelson, 1997). Knowing the original

desorbed gas content and testing the maximum holding capacity make it possible to

estimate gas saturation. Adsorption isotherms were used to estimate the ability to

adsorb volumes of methane as composite samples taken from a single Walloon Coal

seam. Out of the 452 samples, 97 composite samples were created and tested to create

adsorption isotherms. 24.75% of these 452 samples exhibited oversaturation in their

respective pressures, clearly indicating that this test is significantly under-sampled.

Samples indicating oversaturation were then considered to be 100% saturated (results

are shown on Table 3.7 and Figure 3.9).

The equation 3-1 was used to calculate gas saturation (Mares et al., 2009a; Mavor

and Nelson, 1997).

3-1
 1  !  "#! 

Where:
g gas saturation, %
a maximum holding capacity in certain depth in d.a.f. basis
d desorbed gas volume of a sample in certain depth in d.a.f. basis

Gas content does not fit the pattern of gas saturation throughout the seams (Figure

3.9) and a question can be posed as to whether gas content estimation should be

independent from the calculation of coal seam gas saturation; a discussion of this is

further explored in CHAPTER 5. Furthermore, as noted in other CSG Basins in New

Zealand (Mares and Moore, 2008) and United States (Pashin, 2010) coal beds can have

significant saturation variations within a single coal seam (an example is shown in

Figure 3.10 Well 2 example of variation of estimated saturation on a single coal seam.).

93
Chapter 3

Since only a composite sample is tested from each coal seam, this test is thought to

represent the entire coal seam thickness. Cleary this test was under-sampled since the

Walloon coal beds are shown to have different coal properties and gas storage

capacities even within a single coal seam.

Total desorbed gas (m3/tonne) (d.a.f.) Gas saturation (%)


Coal Seam Count Min Max Mean Std. Dev. Min Max Mean Std. Dev.
Kogan 18 1.3 9.5 4.4 2.2 35% 100% 72% 22%
Mac.Upper 53 1.7 14.5 5.7 3.0 29% 100% 74% 19%
Mac.Lower 44 2.7 14.8 7.2 3.0 45% 100% 84% 16%
Nangram 21 1.1 19.0 6.9 4.1 33% 100% 75% 21%
Wambo 37 2.3 37.7 8.5 7.5 39% 100% 81% 18%
Iona 44 2.7 44.9 8.2 6.6 56% 100% 88% 13%
Argyle 79 2.8 16.8 7.3 2.5 43% 100% 80% 16%
Tangalooma Sst. 8 2.6 10.0 7.3 2.5 40% 100% 81% 22%
Auburn 30 2.6 10.8 7.0 2.1 37% 100% 79% 18%
Bulwer 38 2.5 35.9 7.7 5.9 28% 100% 73% 26%
Condamine 80 2.1 19.2 6.6 2.8 24% 100% 64% 18%
Table 3.7 Total desorbed gas (d.a.f.) and gas saturation from 12 cored wells.

94
Chapter 3

Figure 3.9 Mean total desorbed gas volume and gas saturation from 12 cored wells.

Saturation % (d.a.f. basis)


65% 70% 75% 80% 85% 90% 95%
359

360

361
Depth (metres)

362

363

364

365

Figure 3.10 Well 2 example of variation of estimated saturation on a single coal seam.

95
Chapter 3

3.7.2. Coal lithotype and composition effect

With the previous parameters in consideration, there is no direct relationship of

estimated gas content with the different lithotypes. Direct comparison between

lithotypes has not revealed, in a mineral matter free basis, a greater gas content or gas

storage capacity to be associated with a particular lithotype. Nonetheless, it can be

assumed that the gas content in coal for identical conditions (pressure/temperature in

part of the same seam) should increase with brightness. As previously mentioned, as

brightness increases so does organic content, and this is where the gas is mainly stored.

The relationship between gas content and total organic content (from proximate

analysis) is displayed in Figure 3.11. There is a clear organic content relation with gas

content in as analysed basis; fixed carbon plus volatile matter increases as does gas

content. In contrast, it is well documented that ash content reduces gas content and gas

holding ability in coal (Bustin and Clarkson, 1998; Lamberson and Bustin, 1993;

Laxminarayana and Crosdale, 2002, 1999). Moisture yield does not seem to be a

significant factor as it decreases steadily with depth.

Besides the composition of the coal itself, its vertical location is shown to be the key

to gas content and gas saturation. When estimating gas-in-place, by relating it with

organic content (normally through a correlation between organic content and density

measurements), care should be taken with the gas saturation used as it does not

necessary increase with coal seam depth.

96
Chapter 3

14

Gas content as analysed (m3/tonne)


12

10

0
0 10 20 30 40 50 60 70 80 90 100
Organic Content (% weight)

Figure 3.11 Measured (raw) gas content versus organic content measured in 491 core
samples from 12 cored wells.

Gamson et al. (1996) concluded that the gas flow does not depend only on a dual

porosity system constituted by the porous matrix and cleats, but on a complex banded

structure each with its own composition and cleat development characteristics. Also in a

study carried out by Mastalerz et al. (2008b) the different maceral and minerals found

in coal provide it with a variability in micro, meso and macro-porosity and significant

variations in pore surface area. The organic micro-porous matter is considered to be the

main structure characteristic influencing gas sorption in coal by providing a greater

surface area where the methane is primarily stored, while in contrast mineral matter

contributes to more macro-porous volumes in coal. Micro-pore distribution in coal

depends on two main factors, rank, or percentage of fixed carbon, and organic content,

both in quantity and maceral type (Gan et al., 1972; Levine, 1996). Considering that

both vitrinite reflectance and fixed carbon increases generally with depth and the main

macerals remain relatively constant with depth compared to gas content variability

(Tables 3.8), coal micro-porosity and surface area should increase with depth. This

97
Chapter 3

doesnt seem to be reflected in the gas content. Nevertheless it may be indicative of the

variations in saturation and lead to over-saturated samples. Different coal lithotypes

have different maceral components, with different porosities and inner structures

(banding and fracture patterns). The sampling of test gas storage capacity (adsorption

isotherms) is done through a crushed composite sample of the entire seam, therefore no

association can be tested relating gas storage capacity and coal lithotypes.

Fusinite (% m.m.f.) Semifusinite (% m.m.f.)


Coal seam Count
Min Max Mean Std.Dv. Min Max Mean Std. Dv.
Kogan 15 0.0 28.2 7.8 8.9 0.0 43.4 8.6 12.1
Mac.Upper 52 0.0 13.7 2.8 3.0 0.0 62.6 4.8 9.4
Mac.Lower 45 0.0 19.6 2.7 3.8 0.0 52.2 4.8 10.9
Nangram 20 0.0 4.6 0.5 1.1 0.0 2.9 0.4 0.9
Wambo 40 0.0 33.3 1.4 5.3 0.0 33.3 1.2 5.3
Iona 42 0.0 0.4 0.0 0.1 0.0 50.0 1.2 7.7
Argyle 85 0.0 1.1 0.0 0.1 0.0 1.4 0.1 0.3
Tangalooma Sst. 9 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Auburn 29 0.0 0.4 0.0 0.1 0.0 1.2 0.1 0.3
Bulwer 38 0.0 66.7 1.8 10.8 0.0 0.6 0.0 0.1
Condamine 82 0.0 2.1 0.2 0.5 0.0 5.4 0.4 0.8

98
Chapter 3

Telocollinite (% m.m.f.) Desmocollinite (% m.m.f.) Corpogelinite (% m.m.f.)


Coal seam Count
Min Max Mean Std.Dv. Min Max Mean Std. Dv. Min Max Mean Std.Dv.
Kogan 15 3.6 80.0 35.6 21.2 3.4 62.4 22.7 19.9 0.0 13.2 3.8 4.3
Mac.Upper 52 8.8 90.9 40.1 19.2 0.0 62.2 26.0 16.3 0.0 15.5 4.4 3.6
Mac.Lower 45 3.4 81.2 34.5 16.2 1.1 82.8 31.0 21.1 0.0 14.1 4.8 3.8
Nangram 20 22.5 77.8 48.2 15.1 0.0 49.5 21.1 14.6 0.0 15.6 5.4 4.6
Wambo 40 7.4 85.4 45.5 21.4 0.0 64.2 24.8 18.3 0.0 16.6 4.8 4.2
Iona 42 14.3 100.0 34.9 16.6 0.0 61.4 28.5 16.8 0.0 21.6 6.5 5.6
Argyle 85 11.0 83.3 42.5 17.6 2.1 54.2 26.4 14.7 0.0 14.2 5.4 4.0
Tangalooma Sst. 9 12.3 73.2 42.7 19.6 9.7 45.4 28.4 14.1 2.0 8.1 4.6 2.2
Auburn 29 5.0 100.0 40.4 18.2 0.0 50.5 26.6 14.2 0.0 15.3 5.5 4.0
Bulwer 38 0.6 86.6 32.1 19.8 0.0 49.6 25.8 14.4 0.0 20.6 7.9 5.1
Condamine 82 0.0 100.0 40.2 21.1 0.0 96.6 29.8 19.0 0.0 14.6 5.5 3.7

Sporinite (% m.m.f.) Cutinite (% m.m.f.) Suberinite (% m.m.f.)


Coal seam Count
Min Max Mean Std.Dv. Min Max Mean Std. Dv. Min Max Mean Std.Dv.
Kogan 15 0.0 9.1 2.0 2.9 0.0 5.8 2.0 1.9 0.0 35.1 12.8 9.7
Mac.Upper 52 0.0 13.2 2.7 3.3 0.0 8.2 1.7 1.9 0.0 40.8 14.9 8.6
Mac.Lower 45 0.0 8.6 2.0 2.0 0.0 7.6 1.9 2.1 1.8 53.6 16.6 9.9
Nangram 20 0.0 4.8 1.6 1.1 0.0 11.2 3.7 3.6 4.4 41.6 18.1 9.7
Wambo 40 0.0 7.9 2.6 1.8 0.0 33.3 2.4 5.4 0.0 34.0 15.8 9.1
Iona 42 0.0 7.6 2.7 1.8 0.0 10.2 1.9 2.3 0.0 53.4 23.3 12.8
Argyle 85 0.0 16.7 2.7 2.8 0.0 10.3 2.4 2.4 3.4 44.8 19.5 9.8
Tangalooma Sst. 9 0.7 6.6 3.2 1.9 0.0 4.9 1.9 1.6 8.1 30.3 18.1 7.9
Auburn 29 0.0 9.1 2.4 2.2 0.0 7.7 2.9 2.6 0.0 42.9 21.3 11.7
Bulwer 38 0.0 8.4 2.5 2.0 0.0 5.6 1.6 1.5 0.0 66.9 25.5 14.1
Condamine 82 0.0 5.1 1.6 1.2 0.0 7.4 1.3 1.8 0.0 47.1 19.7 9.8

Tables 3.8 Maceral petrographic composition of the Walloon Sub-group (in percentage mineral mater free basis).

99
Chapter 3

Ash content is negatively associated with micro-porosity, gas sorption and number

of cleats (Mastalerz et al., 2008). There is no clear relationship identified between any

specific maceral group and gas content (Figure 3.12 and Figure 3.13) or maceral

distribution and cleat development (Figure 3.14). Nonetheless, there is an association

between maceral in lithotypes and number of cleats. Research by Yao and Liu (2009)

concluded that telocollinite is prone to fracture development. As stated before,

compared to other lithotypes, the Dull banded coal beds have been shown to be more

fractured and these are also the ones who, on an m.m.f. basis, are richer in telocollinite

(Figure 3.15).

12
Gas Content (m3/tonne) (As Analysed)

10

0
0% 10% 20% 30% 40% 50% 60% 70%
Liptinite Content (%) (m.m.f.)

Figure 3.12 Measured (raw) gas content versus liptinite content in 454 samples from 12
cored wells.

100
Chapter 3

12

Gas Conent (m3/tonne) (As Analysed)


10

0
0% 20% 40% 60% 80% 100%
Vitrinite Content (%) (m.m.f.)

Figure 3.13 Measured (raw) gas content versus vitrinite content in 454 samples from 12
cored wells.

100%
90%
80%
70%
60%
% (m.m.f.)

50%
40%
30%
20%
10% Vitrinite
Liptinite
0%
50 100 150 200 250 300
Cleats per m3

Figure 3.14 Mean cleat development against mean maceral content in the Walloon Sub-
group.

101
Chapter 3

0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%

Bright Banded

Dull & Bright

Dull Banded

Dull minor Bright

Dull

Telocollinite Desmocollinite Corpogelinite Suberinite

Figure 3.15 Maceral Petrographic composition of the most common lithotypes of the
Walloon Sub-group.

3.8. Wireline tool responses

3.8.1. Standard logging reading


A total of 1777 coal beds, totalling
totalling 276 metres in thickness from 10 wells (wells 1, 2,

3, 6, 8, 10, 11, 12, 13 and 14), where a direct correlation between core description and

unfiltered wireline log data was possible,


possible were studied.

With the exception of laterolog and induction logs, the wireline logs are usually

provided to the operator with added smoothing filters. Examples of filtered and

unfiltered logs can be seen in Figure 3.16 and Figure 3.17. Unfiltered, unsmoothed,

raw format logs were used to ensure the best analysis of the density, gamma ray, Pe and

resistivity measurements. The high level of uncertainty of the neutron-porosity


neutron

measurement in formations
ons with high apparent porosity such as coal is due to low near-
near

spaced count rates;; consequently,


consequently the analysed NPRS (Sandstone Neutron Porosity) log

data was in the filtered format.

102
Chapter 3

The density dataset is the most important measurement currently acquired for coal

seam gas evaluation due to the relatively low density of the coal and the high-resolution

of this dataset (Kempton and Peeters, 1977; Rieke III et al., 1981). The density log is

commonly used for coal identification, coal thickness estimation and is also indirectly

used to provide estimates of gas content. The most useful density logs provided by

Weatherford for this study are the high-resolution density log (HDEN) (APPENDIX C

wireline density to organic content x-plots). The HDEN was preferred to the near

spaced density log due to the latter higher sensitivity to common washouts. Also the

HDEN was preferred to the conventional compensated density (DEN) due to its

improved resolution.

Figure 3.16 Example of wireline logs supplied to the operator from well 3.
Except SP, caliper and Induction logs, the logs are usually supplied with smoothing filters.
Track 1: the depth in metres. Track 2: SP, gamma ray and Caliper. Track 3: contains
neutron-porosity, density log, sonic and Pe. Track 4: displays the resistivity readings, shallow
focused electric and induction. Track 5: contains the core macroscopic lithological description.
Track 6: Core desorbed gas content (APPENDIX I tools and logs acronyms).

103
Chapter 3

Figure 3.17 An example of wireline data interpretation in well 3 using unfiltered density, Pe,
and FEFE logs.
Track 1: the depth in metres. Track 2: SP, gamma ray and caliper. Track 3: contains
neutron-porosity, density log, sonic and Pe. Track 4: displays the resistivity readings, shallow
focused electric and induction. Track 5: contains the core macroscopic lithological description.
Track 6: Core desorbed gas content (APPENDIX I tools and logs acronyms).

Another useful characteristic of coal is its naturally high resistivity, recorded in

wireline responses from laterolog, induction and high-resolution resistivity tools. For

the resistivity measurements the raw data format was also used to interpret log

responses. Laterolog and induction data are usually displayed with no smoothing filters

while these are added to high-resolution measurements such as the FEFE (shallow

focused electric) log. There is no apparent benefit to filtering such high-resolution data.

This is because the filters make the responses more constant through coal bed zones,

subduing any shoulder effects. However, this apparent benefit does not hold true where

the coal beds are thin and are interbedded with tuffaceous or low organic mudstones.

The sharp responses are clear indicators of bed boundaries between coal and other rock

104
Chapter 3

types with contrasting physical properties. Applying smoothing filters to the data only

acts to mask the boundaries and generally degrade the resolution of the measurements.

The high carbon content in coal also influences the response of neutron logs. The

very low counting rates of the near spaced detector of the neutron tool results in a log

response of apparent porosity above 50 porosity units. Natural gamma ray

measurements of coal usually yield low values that increase with the increasing mineral

content, especially clay minerals. Travel times in coal are typically high, although

dependent on the mechanical properties of the coal, between 120 and 140 us/ft.

Displayed in Figure 3.18 to Figure 3.25 are the responses of wireline log density,

neutron-porosity, natural gamma-ray, Pe, sonic velocity and induction, shallow focused

electric and laterolog resistivities. Through the analysis of this dataset, with the

correlation between centimetric core description and wireline logs, it was possible to

have a clearer notion of the benefits and limitations of interpreting these downhole

measurements over thin bedded coal seams. The integration of the core and wireline

datasets generally indicates values of high apparent porosity and travel times, low

density, gamma ray and Pe. A few coal units, however, show the opposite responses,

the main reason for which is the limited thickness of these coal beds combined with the

physical properties of the adjacent lithologies (mudstones, siltstones, sandstones and

tuffaceous interbeds). Besides issues with coal bed thickness, wireline measurements

also have some data quality issues that can affect the ability to identify thin beds. In

particular, density and Pe measurements require that the tools sensor is in direct

contact with the borehole wall which can be difficult in the event of an oversized

borehole (due to washout) which can commonly occur in these rock types. Gamma ray

responses become more relevant in these cases, although when combined with the

neutron, sonic and laterolog measurements they still do not provide sufficient

105
Chapter 3

quantitative data to allow the detection of thin coal beds. From the resistivity tools

analysed the focused electric log is the only one that seems useful in coal identification

due to its higher resolution compared to other resistivity measurements. Also, the

resistivity histograms (Figure 3.23 and Figure 3.24) indicate clearly that different

measurements of resistivity (using different downhole tools) over the same coal beds

can provided extremely different reading.

100

90

80

70
Thickness (metres)

60

50 35%
31%
40

30

20 15%

10 6% 6%
0.2% 2% 3%
0
1.0 - 1.2 1.2 - 1.4 1.4 -1.6 1.6 - 1.8 1.8 - 2.0 2.0 - 2.2 2.2 - 2.4 2.4 - 2.6
Wireline log Density (g/cm3)

Figure 3.18 Wireline density responses found in coal bed zones in 10 cored wells.
3
Density values above what would be expected in coal (>1.75g/cm ) with a total of 17.6% of
the total coal thickness logged in these wells.

80

70

60
Thickness (metres)

50

40
26% 25%
30
19%
20
11% 11%
10
2%
0.4% 3% 1% 0.1%
0
30 - 40 40 - 50 50 - 60 60 - 70 70 - 80 80 - 90 90 - 100 100 - 110 110 - 120 120 - 130
Wireline log Neutron-porosity (%)

Figure 3.19 Wireline neutron-porosity responses found in coal bed zones in 10 cored wells.

106
Chapter 3

100

90

80

70
Thickness (metres)

60

50
33%
40
27%
30 23%

20
11%
10
3% 3%
0
0 - 20 20 - 40 40 - 60 60 - 80 80 - 100 100 - 120
Wireline log natural Gamma-ray (api)

Figure 3.20 Wireline natural gamma-ray responses found in coal bed zones in 10 cored
wells.

100

90
80

70
Thickness (metres)

60

50 35%
40

30 20%
18%
20

10 6% 7% 6% 2%
0.1% 2% 3% 1%
0
0 - 0.3 0.3 - 0.6 0.6 - 0.9 0.9 - 1.2 1.2 - 1.5 1.5 - 1.8 1.8 - 2.1 2.1 - 2.4 2.4 - 2.7 2.7 - 3.0 > 3.0
Pe (b/e)

Figure 3.21 Wireline photoelectric responses found in coal bed zones in 10 cored wells.

107
Chapter 3

90

80

70

60
Thickness (metres)

50

40 29%
27%
30

20 15%
11%
10 8%
3% 4% 3%
0
0-110 110-115 115-120 120-125 125-130 130-135 135-140 >140
Travel time (us/ft)

Figure 3.22 Wireline sonic log responses found in coal bed zones in 10 cored wells.

40

35

30 Medium
Thickness (metres)

Deep
25

20
58% 60%
15

10
22% 24%
5 14%
13% 3%
8% 0% 0%
0
0-5 5-10 10-20 20-40 >40
Induction resistivity (ohm.m)

Figure 3.23 Wireline induction resistivity responses found in 3 cored wells in coal bed
zones.

108
Chapter 3

14

12

10
Thickness (metres)

6 19%
15%
14%
4 12% 12%

7%
2 5%
1% 4% 1% 4%
3% 1%
0

Shallow focused electric resistivity (ohm.m)

Figure 3.24 Wireline shallow focused electric resistivity responses found in 3 cored wells in
coal bed zones.

60

50
Shallow Deep
40
Thickness (metres)

30
25% 26% 25%
22% 22% 21%
20

10
7% 7% 7% 8%
5% 4%
3% 4% 4% 3% 3% 2%
0
0-5 5-10 10-20 20-40 40-60 60-80 80-100 100-200 >200
Lateralog resistivity (ohm.m)

Figure 3.25 Wireline lateralog resistivity responses found in 8 cored wells in coal bed zones.

109
Chapter 3

3.8.2. Imager and scanner readings


Resistivity and acoustic imaging tools have become common-place in coal seam

exploration due to the high-resolution measurements providing fine detail of the

borehole wall. Interpretation of such logs is useful to define boundaries between coal

beds and other rock types and between coal beds with distinct quantities of organic

matter. The level of detail of these measurements is exemplified in the relative easiness

to correlate them with individual core samples (Figure 3.26).

Borehole images are also extremely useful for the identification and measurement of

major factures. However, identification of coal cleats and banding using these images

within individual coal beds seems not possible, as it appears the scale of these cleats

and bands are too small to allow identification in a systematic way (

Figure 3.27 and Figure 3.28). Also, even in the optimal case of the coal beds to be

sub-horizontal, organic matter estimation is impractical as a result of the fragile nature

of the coal and its response in both tools. Furthermore, in the case of the resistivity

imager, this relies on direct contact with the rock measured. If the coal breaks and

fractures, not all the electrode buttons on the pads may stay in contact with the rock,

creating different resistivity colour contours that dont necessarily represent the real

organic content variation in the coal bed.

110
Chapter 3

Figure 3.26 Example taken from well 2 of imager and scanner responses in coal.
From left to right: 1) Core description; 2) Dynamically Normalized CMI tool image log; 3) ATV Acoustic Televiewer log; 4) Static CMI normalized image; 5)
Acoustic Scanner log.

111
Chapter 3

Cleats

Figure 3.27 (Left) Example of a scanning electron microscope picture showing a cleat with calcite filling.
Figure 3.28 (Right) Coal hand specimen showing face and butt cleats.

112
Chapter 3

3.8.3. Coal lithotype and permeability effect on wireline measurements

The question arose as to whether different lithotypes could generate different log

responses. What is observed is that lithotype undoubtedly affects the measurements

although it is difficult to distinguish them solely using the logs (Table 3.9). There are,

nevertheless, interesting responses and some trends identified. The nature of the

lithotypes influences the shallow log responses. This is not only the organic matter

associated with each lithotype but also its thickness and to certain degree its cleat

frequency and coal banding. In core analysis organic matter increases with brightness

which leads to lower relative densities from stony to dull banded coal beds.

Measurements downhole are affected by the limited thickness of the coal beds. Brighter

coal beds end up indicating higher densities due to their limited thickness. This can

also be observed in the other nuclear measurements where bright coals have lower

resistivity, lower neutron-porosity and higher Pe.

In comparison to other measurements the FEFE log is the one that better separates

different coal beds, even when they are on top of each other (Figure 3.17).

Interestingly, when it comes to resistivity, it is possible to see that besides thickness and

organic matter content the resistivity FEFE log is also to an extent affected by fracture

development and coal banding. Highly banded, highly fractured coal beds may have

their high-resolution resistivity lowered. Dull banded coal seams, which are the ones

showing more banding and more fractures, have lower resistivity than what may be

expected considering they have the lowest wireline density. Nevertheless fracture,

banding or friability cannot be quantified because the resistivity readings are

significantly affected by boundary beds. The interpretation of this measurement and its

importance in CSG gas-in-place estimation is discussed in more detail in CHAPTER 4.

113
Chapter 3

When analysing the CMI image tool it was not possible to tell lithotypes apart
beyond the organic content differences associated with them, which resistivity also
differentiates.

3.8.4. Reservoir permeability

Coal permeability depends on the interaction between the pore space (i.e. pores

within the matrix) and the fracture network (Close, 1993; Levine, 1996). In response to

a pressure drop, adsorbed methane starts diffusing within the coal matrix micropores

and mesopores, the gas then moves to the natural fractures/cleats and finally flows to

the wellbore. The intrinsic property of the coal to develop cleats in greater numbers

than other rock types, acts as a permeability enhancer in what would otherwise be a

tight bed. Cleat filling minimises the permeability of water filled cleats. Both the cleat

density, and cleat filling, together with gas content and gas saturation are shown to

change dramatically from coal bed to coal bed. With detailed core description, it was

possible to verify how many coal beds (number and thickness of coal beds), cleat

density and cleat filling relates to the actual well test performance.

In Coal seam gas exploration, permeability is usually estimated through a series of

DST tests done in exploratory cored wells and appraisal wells, where selected intervals

are targeted to evaluate production (Kabir et al., 2011). Independently of how thick

these coals are, the reservoir is made of a multi-layer, thinly-bedded coal beds

sequence. Each of these layers contributes in their own way to the two-phase flow (gas

and water). The results of these tests in Coal Seam Gas can be less than 1mD (poor

permeability) to more than several hundreds of millidarcies (permeable coal).

114
Chapter 3

Thickness Density PE Gamma Ray Neutron porosity FEFE DSLL RILM


Lithotype (m) (g/cm3) (b/e) (api) (%) (ohm.m) (ohm.m) (ohm.m)
Mean Stdev Mean Stdev Mean Stdev Mean Stdev Mean Stdev Mean Stdev Mean Stdev Mean Stdev
Stony 0.11 0.15 1.82 0.31 1.79 0.86 72 21.1 65 15.1 32 53.3 43 281 7 4.8

Dull 0.12 0.10 1.69 0.29 1.58 0.57 66 25.6 68 15.4 36 55.2 30 64 8 5.5
Dull minor bright 0.19 0.17 1.64 0.32 1.54 0.58 60 23.0 73 14.5 152 423.8 57 197 9 6.1
Dull Banded 0.17 0.14 1.58 0.27 1.44 0.47 59 21.6 73 14.3 116 310.2 55 207 8 3.7

Dull & Bright 0.15 0.14 1.62 0.35 1.5 0.64 55 23.3 72 15.0 188 358.5 55 122 9 4.5

Bright Banded 0.13 0.14 1.63 0.31 1.57 0.44 55 23.3 68 16.4 --- --- 25 38 --- ---

Bright 0.04 0.04 1.90 0.31 1.93 0.82 57 26.4 61 20.8 82 109.5 25 49 5 2.2
Table 3.9 Lithotype mean log responses

115
Chapter 3

Other sources of data that are related to permeability in the dataset provided by the

operator were sorption time and diffusivity measurements taken from individual

samples. As expected the values of sorption time and diffusivity vary widely on a core-

to-core sample basis. These two measurements are known by industry experience not to

be comparable to DST data, emphasising the highly heterogeneous nature of coal

permeability. Sorption time and diffusivity were disregarded as a source of permeability

estimation for the purpose of this study as they are associated with a significant number

of sources of uncertainty related and because they provide very different values of

permeability from the coal in an in situ state (Wold and Jeffrey, 1999).

The dataset received had calculated permeability using equation 3-2. The thickness

(
) used was through wireline log interpretation. Since the dataset also had core

descriptions, a more reliable source of the coal thickness, the net coal was calculated

from core and subsequently the permeability was recalculated. This was conducted by

substituting the coal thickness (


) in the following equation 3-2.

All of the intervals tested are not purely made of coal beds, but of several different

rock types. The permeability values presented are based on the assumption that only

coal beds are providing water and or gas to the main flow. This assumption is grounded

in the premise that non-coal beds are impermeable, with a low effective porosity matrix

composed of smectite- rich sandstones, siltstones and mudstones of illite and smectite

mineralogy.

3-2
$  $
/

Where:
K permeability, mD.
Kh flow capacity, mD.m.
h thickness, m.

116
Chapter 3

In Table 3.10 (see also x-plots in APPENDIX E Permeability and coal bed

characteristics), permeability estimated from 42 DST tests in 13 wells is compared

against coal properties within the test zones. As Mavor and Nelson (1997) investigated;

the individual coal beds work as individual reservoirs. The examples show a large

number of thin coal beds, each of them contributing (or not) with very different water

and gas flow to the borehole. The table also demonstrates that cleat density and

permeability do not have a straightforward relationship. Cleat development in the coal

within the test zones is also revealed to be highly variable, meaning that a highly

cleated (with a cleat spacing smaller than 1mm) coal could be in contact with a cleat-

less coal bed. There is also no indication that only coals with a high cleat density could

allow water or gas to flow in enough quantities to influence these permeability tests.

Cleat filling, as described in core analysis, does not seem to constrain the estimated

permeability. This is understandable since cleats are not completely closed by filling

and their width can changes with production.

117
Chapter 3

Mean bed
Net coal k Cleats Mean cleat Number
Coal Seam Ro max thickness
thickness (m) (mD) per m3 filling of beds
(m)
0.45 3.1 3.5 121 31% 16 0.19
0.44 3.9 1.3 419 40% 11 0.35
Macalister
0.47 1.7 28.8 289 47% 11 0.15
Upper
0.50 9.0 1221.2 486 40% 69 0.13
- 4.0 52.7 195 25% 23 0.17
0.48 7.7 36.5 139 30% 63 0.12
0.44 6.6 6.5 545 9% 33 0.20
Macalister
0.42 4.8 2.7 259 53% 24 0.20
Lower
0.47 1.7 0.5 95 22% 11 0.16
- 2.2 1.6 379 25% 23 0.10
0.43 2.7 28.2 123 11% 18 0.15
Nangram - 5.2 238.8 404 20% 48 0.11
0.42 1.4 0.1 238 22% 9 0.16
0.56 3.4 1419.4 643 12% 27 0.13
Wambo
4.3 13.9 197 39% 20 0.22
0.53 2.5 3.0 101 36% 18 0.14
0.46 3.1 438.7 160 17% 16 0.19
Iona
- 3.3 83.4 503 14% 21 0.15
0.45 4.2 19.7 130 81% 27 0.15
0.50 2.5 136.1 46 32% 17 0.15
0.54 2.4 135.1 117 15% 16 0.15
0.46 3.4 0.1 252 16% 21 0.16
0.45 3.2 11.3 229 43% 17 0.19
0.58 2.8 1318.3 340 6% 26 0.11
Argyle
0.40 3.6 4.6 123 30% 11 0.33
0.48 2.1 247 104 24% 9 0.23
0.46 2.4 33.6 515 17% 19 0.13
0.47 1.0 5.8 495 11% 7 0.15
0.42 4.8 5.7 297 29% 22 0.22
0.48 2.4 0.1 133 51% 13 0.18
0.63 1.8 418.1 169 7% 17 0.11
Bulwer
0.49 3.6 1.6 152 0% 12 0.30
0.48 1.9 31.6 271 10% 14 0.13
0.55 1.4 12.4 109 22% 8 0.15
0.61 1.8 113.9 74 0% 17 0.11
0.48 3.9 211 112 45% 15 0.26
0.49 2.7 0.0 183 19% 18 0.15
Condamine 0.47 4.8 2.2 152 24% 14 0.34
0.53 3.0 0.9 104 10% 15 0.20
- 4.1 30.7 320 30% 15 0.27
0.48 5.1 0.1 542 9% 36 0.14
0.44 6.6 2.9 239 26% 34 0.19
Table 3.10 Drill Stem Test results compared to coal bed characteristics in wells 1 to 8 and
10 to 14.

3.9. Discussion and Conclusions

Investigation into the sub-bituminous to high-volatile bituminous Walloon Sub-

group coal beds has revealed a predominately dull thinly-bedded nature, where the coal

beds are on (mean) average 15cm thick. Dull minor bright and Dull banded coal

lithotypes are the most common, together making 69% of all coal described, and also

forming the thickest coal beds. Bright coal is found in low percentages throughout the

13 wells analysed, mainly in the shape of bands or fragments inside a generally dull

118
Chapter 3

coal bed. The lithotypes could be a part of a relatively thick seam or isolated coal beds

interbedded by siltstones, sandstones and mudstones that may or not have coal laminae.

As core analysis was originally carried out regardless of the lithotype, samples were

selected to reflect the nature of each lithotype. Proximate analyses of selected samples

indicate that as brightness increases, so does fixed carbon and volatile matter against

decreasing ash and moisture content. The Vitrinite group macerals are the most

common. Interestingly, the coal with more bright bands proved to have the highest

concentration of liptinite. The maceral content together with the proximate analysis

results may explain why liptinite content was associated with higher gas content (Scott

et al., 2007).

Cleat development seems to be influenced by the coal type and depth. The number of

cleats in a coal seam is dependent positively on the thickness of banded coal beds of

which it is composed. The mineral content is clearly constraining the cleat development

as the detailed description of the stony and dull lithotypes has shown. Dull banded and

Dull & Bright lithotypes have the closest spaced distribution of face and butt cleats,

followed by Dull minor bright. Cleat length has a negative association with cleat

frequency. Mineral filling in cleats is present in the majority of the coals, especially in

the brightest coals, although rarely completely closing the natural fractures.

No stronger relationship was found between gas content and coal lithotypes

compared to those existing with organic content. Gas content increases with depth until

the Auburn coal measure and then decreases in deeper coals. The saturation variability

was striking on a sample scale. Saturation can be expected to vary also on a coal bed

scale, indicating that each bed constitutes effectively an isolated reservoir. Of the coal

beds analysed, coal type overshadows all other factors. The geochemical properties of

119
Chapter 3

each lithotype are therefore influencing cleat development, gas content and gas storage

capacity.

Wireline logs show non-typical responses for coal in thin beds due to the limitations

of the individual tools resolution. Furthermore, differences in wireline log readings in

distinct lithotypes were identified although it is difficult to separate them in a

systematic way. Resistivity seems to be influenced to a certain extent by the lithotype

banding and fractures. Interpretation of image logs was found to be very useful for coal

seam thickness determination, identification of major fractures and for core-log

correlation.

The levels of heterogeneity in the coal seams create difficulties in the identification

of influential factors on the permeability estimated through DST tests. Nevertheless, the

general trends show that the more permeable coals seem to (1) be thinly bedded, i.e.,

the coal beds within the test zone are thinner and more numerous; (2) are more mature;

and (3) have a higher number of cleats per m3. The results may be indicative of (1) the

alluvial and lacustrine nature of the Walloon Sub-group, which may lead to limited

horizontal continuity of the cleat development described from the core; (2) limited

continuity of the coal beds and/or coal seams far beyond the borehole radius; (3)

isolation of single coal beds, even in cases that are a part of a coal seam, which could be

due to the lack of cleat development between the beds preventing inter-bed water and

gas flow; and (4) gas saturation may be affecting the DST test results due to its high

variability on an individual coal bed basis.

120
Chapter 4

CHAPTER 4 Estimating coal thickness in

Thinly Bedded Coal Seam Gas (CSG)

4.1. Introduction
To assess the economic value of this unconventional reservoir various well data

inputs are needed and most of them are acquired from core analysis, wireline logs and

drill stem tests. When compared to other hydrocarbon accumulations, the vertical and

lateral variability found in CSG requires a greater number of wells to be drilled. For

economic reasons, the number of cored wells is usually limited and consequently

standard wireline measurements are seen as a more economical practice to evaluate

fundamental parameters including coal thickness, coal quality and indirectly, coal gas

content (Mullen, 1989; Rogers, 2007; Seidle, 2011).

The estimation of reservoir thickness, also known as net coal, is a fundamental

aspect of Coal Seam Gas resource assessment as it directly influences the estimation of

gas-in-place reserves and consequently the overall economic analysis (Dhir et al.,

1991). Due to the organic-rich nature of coal, the standard approach used to quantify net

coal is the analysis of high-resolution open-hole density log data. This estimate is

nevertheless associated with being one of the main sources of error in CSG reservoir

modelling, due to the usage of a defined wireline bulk-density maximum value of

typically a value of 1.8 g/cm3 which is used to estimate gross reservoir thickness

(Lagendijk and Ryan, 2010; Mavor and Nelson, 1997; Nelson, 1999). Field experience

and the use of modern high-resolution measuring tools has diminished the degree of

uncertainty, although as demonstrated in this study, it remains significant when the coal

121
Chapter 4

seams being measured are thinly bedded; this is normally the case in CSG worldwide

(Colson, 1991; Mavor et al., 1990a).

Coal seam gas wells have distinct characteristics with regards to drilling, wireline

logging operations and interpretation (Mavor and Nelson, 1997; Palmer, 2010; Rogers,

2007; Seidle, 2011). The wells are usually shallow (less than 1000m) where special

consideration is taken into borehole stability due to common washout problems. Coal

and associated shaly lithologies have a tendency to create cavings and borehole

rugosity, and this is detrimental not only to core recovery but also to wireline logging

quality, especially in tools dependent on good contact with the borehole wall.

In this study a workflow is proposed for use as part of CSG wireline log

interpretation. This is tested in three cored and two production wells. It combines

density with resistivity measurements to improve the estimation of thinly bedded net

coal thickness in CSG.

As an effective electrical insulator, coal has relatively high electrical resistivity.

Standard resistivity measuring tools provide medium to high-resolution measurements

with different vertical ranges. The most common resistivity tools being run in CSG

wells are a combination of induction and shallow focused electric tools. An

investigation into the resistivity log responses has indicated that different logging tools

provide very different responses across coal seams. The reliability of associating

downhole resistivity log responses to coal permeability (Hoyer, 1991; Yang et al.,

2006) is tested. This involved the correlation between log data and core analysis from

thinly bedded, highly-volatile bituminous coal seams, from the Walloon Sub-group,

Surat Basin, Queensland, Australia.

122
Chapter 4

4.2. Core and wireline data correlation


The data set for this study was acquired in five wells drilled in the Surat Basin in

Queensland Australia. These data are used to evaluate and test the production potential

of coal seam gas from the Walloons Sub-group Coal Measures. For this study the data

set comprises wells where high-resolution focused electric tools were run in cored wells

(9.6cm diameter borehole) and also in production test wells (21.6cm diameter

borehole). Core recovery in the wells 1, 2 and 3 was conducted through a slim-hole

wireline coring system to minimize core damage and gas loss. The full Walloon Coal

Measures section was described macroscopically (Figure 4.1) by Earth Data Pty Ltd

Geological & Earth Science consultants personnel. Coal type was identified following

the Australian standard (Australian Standard, AS 2916-2007. Symbols for graphical

representation of coal seams and associated strata, 2007) lithotype classification. Also,

with regards to coal, the cleat (face and butt) spacing and penetrability are also

described. In these wells eighty-nine core samples were retrieved from the cored

section for special core analysis.

Besides cored wells 1, 2 and 3, wireline log data from production wells 3A and 3B

were also examined. These wells are part of the same CSG exploration development

plan of cored well 3. Not only are they in geographical proximity of each other (within

a distance of 1.7km) but they also have similar depth intervals drilled (Table 4.1). The

logging suite included resistivity (induction, shallow focused electric and laterolog),

nuclear (gamma ray, density, Pe, neutron porosity), spontaneous potential and caliper

measurements.

123
Chapter 4

Total Depth Walloons coal Core desorption


Wells
(m) measures depth (m) samples
1 (cored) 760 153 - 566 13
2 (cored) 610 297 - 610 37
3 (cored) 698 346 - 670 39
3A 716 364 - 660 ---
3B 734 350 - 653 ---
Table 4.1 Total well depth and Walloons Sub-group depth intervals identified in core and
wireline logs. From cored wells 1, 2 and 3 selected samples were tested for special core
analysis.

Figure 4.1 Correlation between well 1, 2 and 3 based on core description. For location see
Figure 2.22 in chapter 2.

124
Chapter 4

The data acquired for this study vary considerably, with each measurement

technique possessing its own vertical resolution. In addition to each downhole tool

having its own specific resolution, the data from core analysis also have their own scale

of measurement. Core description was provided on a centimetre scale. Correlation

between core macroscopic description and log response was matched at 40

samples/metre intervals (an increment of 0.025m). With regards to the core analysis, the

samples had lengths between 0.3m and 1.0m depending on the bed thickness and

limited by the length of the desorption canisters. Therefore, to relate the special core

analysis with the log response, the core thickness was averaged over the depth interval

where the core sample was recovered.

4.3. Coal Properties


In order to interpret the wireline logs it is important to understand the nature of the in

coal bearing strata. Overall, the data provided are broadly in the same setting described

by previous studies focussing on the Walloon Sub-group (Hamilton et al., 2012; Martin

et al., 2013; Scott et al., 2007). They consist of an assemblage of sub-bituminous to

high-volatile bituminous coal lithotypes with vitrinite reflectance between 0.36% to

0.55% Ro max, dominated by Vitrinite (60%-80% average), variable Liptinite (40%-

5%), very low Inertinite (<5%) interbedded with carbonaceous mudstone, tuffaceous

and siltstones beds of variable thickness. Besides coal lithotypes coal was also found in

the shape of thinly bedded layers (less than 2.5cm coal laminae) in mudstones and

siltstones. As described by Martin et al., (2013), they can be identified as Class I or

Class II, forming thicker or thinner coal seams with different lithotypes stacked on

top of each other, separated by other lithologies. The total thickness of each lithotype

can be seen in Table 4.2. In the three cored wells, an overall total of 323 individual coal

125
Chapter 4

beds were identified. Coal bed mean thickness averages 0.19m, ranging from a

minimum thickness of 0.025m up to 0.925m. Well 1 and well 3 have similar coal seam

distributions, with coal seams having on average 0.20m thickness in both wells. While

well 2 has thicker coal seams, with an overall average thickness of 0.37m.

Total Thickness (metres)


Coal Lithotype
Well 1 Well 2 Well 3
Dull & Bright 1.8 0.6 0.6
Dull banded 3.3 4.3 5.1
Dull minor bright 0.6 18.6 13.8
Dull 0.7 3.4 2.6
Stony 0.4 3.9 3.2
Total coal 6.8 30.8 25.3
Total Coal Laminae 1.5 9.2 5.1
Table 4.2 Coal facies identified in the core description

4.4. Wireline logging for Coal Seam Gas


Along with nuclear and sonic tools, which are run in all wells, resistivity tools are

also run, though these may vary in measurement technique from well to well depending

on the salinity of the drilling mud. The drilling fluid used in CSG wells is usually water

based with an exception found in cases of wells extensively cored. In these wells the

main objective is to recover core in the best condition for special core analysis and core

description. Special core analysis is a key aspect of the overall CSG exploration. In

addition to providing unique analysis, core data works as a correlation anchor to what

could be, in the case of an economically viable project, a program with hundreds of

wells with proximity of less than 1km apart where wireline logs are always run, but

core is not recovered. Cored wells account for only the number of initial pilot wells in

the region. Therefore, due to the limited thickness of the coal beds and the limited

numbers of cored wells, net coal thickness determination comes in great part from the

analysis and interpretation of wireline data.

126
Chapter 4

As observed in the core description and indicated by Martin et al. (2013), coal may

or may not grade into shaly lithologies. The more obvious sharp contacts are found in

contacts between coal and volcanic tuff beds, which besides the contrast in macroscopic

characteristics have completely different compositions with tuff beds having no organic

content. A hypothetical log used for thin coal seam interpretation should indicate

contact between contrasting facies, not only because of its competence to identify

dramatic changes in composition of the formation, but also to separate those formations

that can be potential coal seam gas producers.

For some time tool resolution has been an issue when dealing with wireline logging

in coal exploration (Kempton and Peeters, 1977; Samworth and Cherrie, 1976). As

demonstrated in the previous chapter, to understand the nature of potential productive

methane coal beds, coals thin scale heterogeneity demands a similar scale in its

measurement. Interpretation of log responses in wells drilled in the Walloon Sub-group

suffers from the same issues identified in other Basins with low thickness of individual

coal beds.(Mavor et al., 1990a) analysed evaluation procedures in CSG wells drilled in

the San Juan Basin, ranking the log measurements in terms of importance. Primary

importance was assigned to the high-resolution density log and secondary or tertiary

value to logs such as neutron, gamma ray, Pe, dual induction and laterolog resistivities

for quantitative purposes. Colson (1991), analysing log responses over coal beds less

than 1ft thick in the Black Warrior Basin, considered gamma ray and standard-

resolution density logs to have wrong and inadequate responses. Examples of

intervals that could lead to such statements are also found in the data set described in

this thesis. Figure 4.2 reveals a correlation between a zone where core was recovered

against standard wireline log measurements. Both the bulk density and the shallow

focused electric (FEFE) resistivity log responses found in that image are designated by

127
Chapter 4

the service company as high-resolution with a vertical resolution of 15cm. This level

of resolution in density depends on the source-receiver spacing and has been considered

to be the best possible for more than 40 years for this type of tool based on a gamma ray

source of Cesium 137. As shown in the same figure, interpretation from other nuclear

measurements (gamma ray and neutron-porosity) suffers from their 61cm (2ft)

resolution. The example shown in Figure 4.2 also indicates that although the two last

log responses are being influenced by the coal beds, a quantitative interpretation would

result in a significant degree of uncertainty. The coal bed thickness would be

overestimated (neutron-porosity log) or underestimated (gamma ray). When estimating

the net coal in thinly bedded coal seams, it is not just the question of identifying the

coal seam but also distinguishing the individual beds from other lithologies.

128
Chapter 4

Figure 4.2 An example of wireline data correlated with core analysis from well 2.
Track 1: the depth in metres. Track 2: SP, gamma ray and caliper. Track 3: contains density
3
log (with less than 1.8 g/cm shading) and the neutron-porosity. Track 4: displays the resistivity
readings, shallow focused electric, induction and laterolog. Track 5: contains the core
macroscopic lithological description. Track 6: derived lithological log using density log together
with induction and shallow focused resistivity log measurements (APPENDIX I tools and logs
acronyms).

4.4.1. Net coal estimation using density log


The definition of a cut-off and its application as a policy are critical in CSG

exploration, not only to know the volume of coal that can be a potential methane

producer but also to define well completions and as part of permeability estimation.

Furthermore the net coal estimation that results from a cut-off workflow can directly

affect the ultimate permeability of the producing coal beds, as thick coal seams are

selected to be under-reamed in open-hole conditions, so gas flow rate can be enhanced

by increasing the wellbore radius and cleaning drilling damage zones (Palmer, 2010).

129
Chapter 4

The importance of the accurate estimation of coal quality and coal thickness led to

the high-resolution density log response being of primary importance. Due to its

organic rich nature, coal has a lower bulk density than other sedimentary lithologies.

Density log data is used to estimate coal seam thickness and ash content so that gas in

place can then be calculated by the following equation (Mavor and Nelson, 1997):

GIP=(A*h**Gobs)

Where GIP is the gas in place (m3), A is the drainage area (m2), h = net coal

thickness (m), is the bulk density of the net coal (g/cm3) or (tonne/m3) and Gobs is the

observed gas content (m3/tonne).

The gross thickness (h) used today is estimated by using an in situ open-hole density

log cut-off. The choice of this cut-off can lead to one of the major sources of error in

CSG reservoir gas-in-place estimation (Lagendijk and Ryan, 2010; Nelson, 1999).

100

75
Weight (%)

50

25

0
1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6
Core Density (g/cm3)
Ash Content Organic Content Content Moisture Content

Figure 4.3 Coal ash and organic content against core density cross plot of wells 1, 2 and 3.

In the Walloon Sub-group, core density can be associated easily with the basic

composition of coal determined in proximate analysis. Core density against percentage

of weight of moisture, organic content and ash content for wells 1, 2, and 3 is shown in

130
Chapter 4

Figure 4.3. According to the geological definition of coal, for a sedimentary rock to be

classified as coal, it must contain less than 50% of mineral content by weight and more

than 70% of organic content by volume. The relationship found between ash content

with density and the possibility of measuring density with a high-resolution tool

downhole, leads to the application of a density cut-off as the standard method for

identifying coal in wireline log analysis. A density cut-off is a maximum value that is

applied to the density log to differentiate coal from other formations. The choice of this

value has major significance in the estimation of Gas-in-place. There is an associated

error related to borehole cavings and sharp lithological contacts, and to thin bed coal

thickness that could be less than the tools resolution.

Figure 4.4 shows the average for the density measured by a high-resolution density

log against the density of core recovered in the same zone (APPENDIX D wireline

responses). Thus it can be seen that the quality of the downhole measurement is

affected by borehole washouts making the density values lower. Moreover, it can be

observed that measuring density downhole in thin bedded coals has a tendency to

underestimate the coals quality (i.e. the density measured in core tends to be less than

that downhole). Today this leads to most companies not to follow the geological

definition to identify coal when interpreting wireline density, which from Figure 4.3

would be close to 1.6 g/ cm3; instead a higher density maximum value is used.

Experience of working with wireline logs in coal has led log analysts to increase this

density cut-off to values between 1.75 g/cm3 and 1.85 g/ cm3.

131
Chapter 4

2.6

2.4

2.2
Core density (g/cm3)

2.0

1.8

Borehole diameter
1.6 <10cm
10 to 20cm
1.4 20 to 30cm
30 to 40cm
>40cm
1.2
1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6
High-Resolution wireline Density (g/cm3)

Figure 4.4 Wireline against core density with colour distribution of borehole enlargement of
wells 1, 2 and 3.

The question is thus, what value should be considered as a cut-off? The application

of too low a log density cut-off can remove non-producing higher density formations

although it can also easily remove organic bearing strata. These methane productive

sedimentary sequences are organic-rich mudstones, with coal layers and coal beds that

due to their higher mineral content have higher density log readings, such as that of

shaly or stony coal. Additionally, because coal seams are thin, the density log readings

are elevated by the neighbouring beds. However, if the policy is just increasing the

maximum density cut-off, then there is an associated risk of over-estimating the

reservoir size by introducing other lithologies that have negligible gas storage capacity.

Since wells 1, 2 and 3 recovered core over the entire sequence of the Walloon Sub-

group, it is possible to correlate it to the high-resolution density log and investigate how

much of the coal was estimated by using an increasing density cut-off. This is shown in

Figure 4.5. In this cross-plot the x-axis is the ratio between the reservoir rock and the

estimated reservoir rock using a cut-off, in the case of this figure, a maximum density

132
Chapter 4

cut-off. In the y-axis is the percentage of total reservoir rock detected, as identified in

core. In other words, the score at the horizontal indicates the quality of the estimation,

while the vertical is indicative of quantity. A good methodology should score high in

both.

The fragile nature of coal seams usually leads to washout effects, where the density

tool readings are unreliable, creating difficulties distinguishing coals beds from nearby

formations. A good example of this was detected in well 1, where several zones

suffered washouts. Figure 4.5 shows how this makes well 1 score less in both axis x

and y. Wells 2 and 3 did not have the washout issue that was detected in well 1.

Nevertheless, they do not demonstrate the same behaviour in Figure 4.5. This is related

to the general thickness of the coal beds. The frequency of thicker coal in well 2 is

significantly higher than well 3. The density tool is therefore less affected by boundary

formations in well 2 and consequently a greater part of the reservoir is characterised by

lower readings in the density log compared to well 3. This explains why in Figure 4.5

the well 2 shows better results.

133
Chapter 4

100%
2.6g/cm3 2.4 g/cm3 2.0 g/cm3 2.2
90% 2.0

1.8g/cm3
80%
1.8 1.8
Total core reservoir thickness

70%

60%
1.6g/cm3
1.6
50% 1.6

40%

30%
Well 1 1.4 g/cm3 1.4
20% 1.4
Well 2
10% Well 3

0% 1.2
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Core reservoir rock thickness/Estimated reservoir thickness

Figure 4.5 Study of density log cut-off value.


X-plot between the ratio of core reservoir rock thickness to estimated reservoir thickness
using a maximum wireline density cut-off against the total reservoir thickness.

4.5. Coal borehole resistivity

4.5.1. Standard downhole resistivity measurements


There are two broad factors to be considered when measuring resistivity downhole:

geological and technical. The matrix (e.g. minerals, cements) of most rocks are

electrical insulators, although sedimentary rocks are often porous and may contain salty

water where electrical current can flow with little resistance. The high concentration of

organic material found in coal constrains the pore space and reduces the pore

connectivity making this sedimentary rock highly resistive when compared to other

sedimentary rocks. Bed thickness, relative to the measurement scale, is also a

significant influence on the resistivity log response. Resistivity in conventional oil and

gas exploration has long been a source for identification of hydrocarbons through the

separation of shallow and deep measuring resistivity logs, for picking formation tops

134
Chapter 4

and bottoms, and for estimating formation hydrocarbon saturation using Archies

equation. However in coal logging, with the common use of water-based drilling muds

and the low salinity inside the borehole environment, the thin coal beds do not create

the invasion profile identifiable in more porous formations when invaded by mud

filtrate. To interpret an association of two measurements, one shallow and one deep, it

is fundamental to ensure that both are measuring the same formation vertically. This is

relevant because of the thin nature of the coal beds, in which the deep measuring log is

more affected by the boundary formation on the top and bottom of the coal than the

shallow log. This leads to the deep resistivity log showing a lower value than the

shallow log. This represents a typical result when measuring deep and shallow

resistivity through the induction tool in thinly bedded coal seams and cannot be

interpreted in the manner applicable to conventional reservoirs..

Downhole resistivity devices have different formats with different electrode sizes

and spacings. Depending on the device, coal can show a wide range of electrical

resistivities, measured downhole, over the same seam. This is demonstrated in Figure

4.2 where a section of well 2 is shown where both the Array Induction tool (AIT) and

Dual Laterolog tool (MDL) were run after the mud system was changed by adding

potassium chloride (KCl) to the water based fluid. The induction tool and the focused

electric tool are the resistivity measurements most often used in coal seam gas, at least

in the case of the Australian Walloon Sub-group, where most of the wells are drilled

with fresh water based mud. The induction tool samples the formation with several

depths of investigation (medium and deep) away from the borehole. The approximate

1.0 metre distance between the transmitter and the receiver coils defines the induction

resolution which is clearly impractical for interpreting coal seam layering with its

heterogeneities, with mudstone and tuff interbeds. The measurements are shown to be

135
Chapter 4

highly influenced by the inorganic beds bordering and interbedded in the coal seam, and

consequently the measurement does not provide the characteristic of high resistivity

normally associated with coal. In the same situation the laterolog response, because it

possesses a closer spaced electrode arrangement, has a shorter vertical investigation and

is less affected by conductive beds, thus providing resistivity values higher than an

induction survey as far as thin beds are concerned. Yet again the laterolog measurement

does not resolve shale or tuff beds within a coal seam.

4.5.2. MFE Response to coal thickness


The running of the AIT assembly tools is associated with another resistivity device,

the high-resolution shallow focused electric (MFE). Similar to the AIT and the MDL

described above, the MFE tool is centralised in the borehole. Comparison with core

data leads to the conclusion that this high-resolution shallow focused electric (MFE) log

has the best precision in thinly bedded coal seams. High-resolution resistivity shallow-

focussed devices show a clear difference between coal and other lithologies that is

distinct from induction and laterolog devices, and that are far more complicated to

interpret. When recording any non-contact resistivity measurement, the electrical

current will flow through the path of least resistance. The recorded resistivity can

therefore be significantly reduced by conductive fluid saturating more porous boundary

formations on either side of a coal bed or mineral-rich interval within the coal bed.

Correlation of the core samples macroscopic geological description and proximate

analysis with log responses suggests that the most influential factor affecting the MFE

reading is the organic content distribution, the relative volume of organic content it

contains and the thickness of the bed (Figure 4.6). The spiky response found in the

FEFE is thus associated with the existence of a single bed and a more massive irregular

136
Chapter 4

response with the packing of several independent beds. The organic content together

with the thickness of the coal seams controls the character of the readings. Therefore,

although a specific coal seam has indicated in proximate analysis to have high fixed

carbon and volatile matter, it may not show high resistivity in the FEFE log due to its

thickness. The high sensitivity of this tool to coal thickness, which does not produce a

quantifiable estimate of organic content, but if used in coordination with other logs, can

assist in the net pay estimation of thinly bedded coal seams. Figure 4.7 shows an

example in well 3 of how the macroscopic description of core relates to the MFE

response to coal beds, thinly bedded coal wisps and carbonaceous layers within

mudstones.

100
Organic Content (% by weight)

90
80
70
60
50
40
30
20
10
0
1 10 100 1000 10000
Shallow Focussed Electric Resistivity (ohm.m)

Coal Seam thickness < 15cm 15 to 30cm >30cm

Figure 4.6 Shallow Focused Electric Resistivity against Organic Content estimated from
proximate analysis as a function of coal seam thickness.

A thinly bedded coal seam in most cases consists of a variety of coal lithotypes with

variable organic content and inorganic interbedded intervals that vary in thickness.

Carbonaceous mudstones, inorganic mudstones, volcanic tuffaceous layers or siltstones

and sandstones, constitute a more conductive flow path then coal due to their porous,

often water saturated nature. In contrast to the induction readings, the shallow focused

137
Chapter 4

electric log is able to give an indication of the coal seam heterogeneity. The resistivity

measured by the MFE is sensitive to this organic content and to porosity changes. The

vertical continuity of highly rich coal lithotypes (i.e. dull & bright, dull banded, dull

minor bright and dull coal beds) is associated with high resistivity, whereas stony coal

and shaly coal are associated with lower resistivity. The alternation of these lithotypes

within a coal seam are evident in the FEFE readings; these vary between high values (

100.000 ohm.m) and low values slightly above the induction readings (as low 2

ohm.m).

Figure 4.7 An example of wireline data correlated with core analysis from well 3.
Track 1: the depth in metres. Track 2: SP, caliper and gamma ray. Track 3: contains density
3
log (with less than 1.8 g/cm shading) and the neutron-porosity. Track 4: displays the resistivity
readings, shallow focused electric and induction. Track 5: contains the core macroscopic
lithological description. Track 6: derived lithological log using density log together with induction
and shallow focused resistivity log measurements (APPENDIX I tools and logs acronyms).

138
Chapter 4

4.5.3. Effect of permeability on resistivity response


If present in coal beds, face cleats and butt cleats together with coal banding can

enhance the permeability of coal, creating a dual-porosity system together with the

microporosity found in the coal matrix. The cleat system, if well developed, can then be

invaded by the mud fluid injected downhole. Therefore in theory, when the borehole

fluid is more conductive then the formation fluid in a highly fractured coal, the deep

measurement can be higher than shallower resistivity, generating a separation between

deep and shallow resistivities. Downhole measurement of the permeability related to the

cleat development is nevertheless debatable, as several authors (Ahmed et al., 1991;

Gash et al., 1992; Nelson, 2000a; Purl et al., 1991) have considered that porosity

originated by cleats in coal beds can be considered to be smaller than 2%.

In this work the only proof that coal permeability is in a limited way related with

resistivity was indicated in the previous chapter, where the dull banded coal lithotype

beds had lower FEFE resistivity readings compared against other lithotypes (see Table

3.9). This suggests that resistivity may be influenced not only by the higher cleat

frequency of the dull-banded lithotypes but also related to the fact that these are banded.

Several publications have indicated an association between wireline resistivity,

particularly laterolog and microresistivity logs, and coal permeability (Hoyer, 1991;

Mavor et al., 1990a; Yang et al., 2006). Even though this is not the focus of the current

chapter, a few comments are necessary to address this point. If such an association

could be confirmed, a highly permeable coal could theoretically have a low resistivity,

consequently leading to underestimation of coal thickness using a measurement such as

the FEFE log. Although all of these referenced publications address coal permeability

as dependent on cleat development they do so referring to limited production data

results and do not directly relate to the log responses in the presence of cleats.

139
Chapter 4

In Hoyer (1991), the study of the laterolog response led to the assumption that it is

responding to cleat development. This premise, however, has several flaws. First, it is

based on the modelling study of fractured rocks of Sibbit and Faivre (1985), where in

the case of a thick, high-resistivity formation a model can be created to estimate

fracture porosity through the analysis of laterolog responses. The coal seams in CSG are

however usually associated with thin beds, where a resistivity tool such as the laterolog

has a resolution greater than the actual thickness of the coal bed it is measuring.

Secondly, there is a lack of attention given to the nature of the coal, and more

specifically to its composition. Coal resistivity is associated with its composition, which

could to a certain extent be derived from the density log. In coal, density and resistivity

logs broadly reflect the nature of coal (composition and thickness) demonstrated in

many cases by their similar characteristics if both tools are run in non-washout zones,

and both logs are of similar resolution. The electric current from a high-resolution tool

can nevertheless show a more spiky character than the density due to its sensitivity to

less organic-rich layers inside a coal seam zone. Variations of resistivity within a coal

seam are often associated with organic content variation and not so much with cleat

development. Illustrations of this statement are found in all of the logs shown in this

work, where examining the correlation between the core descriptions and the FEFE log

show many log readings are affected by the organic content and thickness of the bed.

Figure 4.9 shows a more direct demonstration of this. Note that if the resistivity

measured by the MFE was being significantly affected by cleat frequency, coal beds

with identical densities (i.e. similar organic content) would show a resistivity

dependency on the number of cleats, which was not confirmed.

When analysing wireline responses over coal beds in the Uinta Basin in the US,

Yang et al., (2006) assume that cleat porosity can be estimated through downhole

140
Chapter 4

wireline resistivity interpretation. Although the authors assume correctly that the coal

matrix porosity is minimal, they incorrectly considered that fluid-filled porosity existing

in a highly developed cleat system could influence a focused resistivity log. A

straightforward way to disprove such a theory is found in the geology of coal. Coals

with a distinct fracture development are in many cases found in close contact with each

other, within the same coal bed (see example in Figure 4.8). If the theory is that a

resistivity log measurement can be influenced by cleat derived permeability to a point

that it could be measured, then the log should indicate a dramatic variability from an

impermeable bed to a permeable one. No such behaviour, however, has been identified

in any of the wells interpreted for this study. Cleat spacing in coal is measured on the

order of millimetres, changing from coal bed to coal bed within a coal seam at a

centimetric scale. No standard resistivity tool has the capability to be influenced by a

changing fracture system on that scale in coal.

Cleats/m3
0 100 200 300 400 500 600
150

200

250

300

350
Depth (metres)

400

450

500

550

600

650

Figure 4.8 Cleat development against depth in all the coal bed found in well 2.

141
Chapter 4

High-resolution shallow resistivity (ohm.m)


1.0 10.0 100.0 1000.0
1.2

1.4
High-resolution density (g/cm3)

1.6

1.8

2.0

2.2 No cleats
<100 cleats/m3

2.4 >100cleats /m3

2.6

Figure 4.9 High-resolution wireline FEFE shallow resistivity against high-resolution wireline
density showing coloured variation on cleat development in each coal bed.

4.6. Net coal thickness estimation in washout zones


CSG wells are often associated with enlarged washout sections. A question arises as

to which log can be used to minimize the lack of reliability on the density log when

cavings occur. Figure 4.10 shows a zone in well 1 where three cavings are affecting at

least the density tool readings. These cavings can clearly be identified by the increased

caliper readings. The gamma ray and neutron logs are considered to be useful as backup

logs in washouts (Mavor et al., 1990a; Seidle, 2011), as they do not have the same

sensitivity to washouts as the density, although it can be observed that neither do they

have the precision of the MFE. The FEFE log is displayed with a yellow shading to

indicate which values are above the induction logs. Notice how the increased FEFE log,

compared to the induction medium and deep logs, corresponds to the presence of

organic material in coal or in carbonaceous mudstones (lithological log in track 5); all

seem to be unaffected by borehole enlargement.

142
Chapter 4

Research on using gamma ray, neutron-porosity and sonic (DT35) logs as net coal

quantifiers are shown in Figure 4.11 to Figure 4.13. There are two aspects that can

easily be identified. Firstly, the borehole cavings are also affecting the sonic, gamma

ray, and to a lesser degree the neutron. A good example of this is shown in Figure 4.10.

Notice how at 368.8m there is a caving indicated by the larger caliper log, where the

gamma ray is too low, and the neutron (80%) and the sonic (140us/f) are too high for a

siltstone. The only log indicating it is not a coal bed at that depth is the resistivity log.

Secondly, the thickness of the coal still affects the effectiveness of reservoir

identification as well 2 consistently remains the easiest to estimate net coal thickness

since it has the thickest coal seams. It can be concluded that the neutron-porosity is the

most effective back-up primary log in cases of non-reliable density measurement.

143
Chapter 4

Figure 4.10 An example of wireline data correlated with core analysis from well 1.
Track 1: the depth in metres. Track 2: SP, gamma ray and caliper. Track 3: contains density
3
(with less than 1.8 g/cm shading), neutron-porosity and sonic log. Track 4: displays the
resistivity readings, shallow focused electric and induction (shading place where the FEFE
resistivity is higher than the induction logs). Track 5: contains the core macroscopic lithological
description. Track 6: the log derived lithological log, shows particular good agreement with the
core beds and carbonaceous mudstones with coal wisps and closely spaced carbonaceous
laminae even in washout zones (high caliper) - (APPENDIX I tools and logs acronyms).

144
Chapter 4

100%
Well 1
90%
100
90 Well 2
90
80% Well 3
80 80
Total reservoir thickness

70% 90
70 70
60%
80
50% 60
60
40%
70

60 50
30% 50
20% 50 40
40
10% 40
30 30
0% 20 30 20 20
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Reservoir rock thickness/Estimated reservoir thickness

Figure 4.11 Study of gamma ray log cut-off value.


X-plot between the ratio of core reservoir rock thickness to estimated reservoir thickness
using a maximum gamma ray (in API) cut-off against the total reservoir thickness.

100% 40
30 50
90%
80% 60
Total reservoir thickness

70%
60%
70
50%
40%
30%
80
Well 1
20%
Well 2 90
10%
Well 3 100
0%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Reservoir rock thickness/Estimated reservoir thickness

Figure 4.12 Study of neutron porosity log cut-off value.


X-plot between the ratio of core reservoir rock thickness to estimated reservoir thickness
using a minimum neutron-porosity (percent) cut-off against the total reservoir thickness.

145
Chapter 4

100% 100 110 110


110
90%

80% 120
Well 1
Total reservoir thickness
120 120
70% Well 2
60% Well 3

50%

40%
130
130
30% 130
20%

10%
140 140
0% 150
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Reservoir rock thickness/Estimated reservoir thickness

Figure 4.13 Study of sonic log cut-off value.


X-plot between the ratio of core reservoir rock thickness to estimated reservoir thickness
using a minimum compensated sonic (microseconds/foot) cut-off against the total reservoir
thickness.

4.7. Using Resistivity in net coal estimation


In conventional oil and gas exploration, it is common for reservoirs to be thick and

restricted in numbers. In CSG, the target reservoirs are many thinly bedded coal beds.

Through core description, 90 individual coal beds were identified in well 1, with 115 in

well 2 and 160 in well 3. Direct correlation between core and the density log has

indicated that to maximize coal bed detection the cut-off for the high-resolution density

log should be 2.0g/cm3. Isolated thin coal beds have nevertheless shown considerably

higher density (>2.0g/cm3), due to their very limited thickness, making up to 10% of the

total coal drilled in these wells. Although still considered as part of the reservoir for this

study, these isolated thin coal beds may not be considered in the total net coal thickness

because their significance as a methane producer is limited. Such a cut-off, is illustrated

146
Chapter 4

in Figure 4.5, and is enough to add to the estimated 90% of the coal identified in core.

Alone a density cut-off has a significant margin of error, especially in wells with

washout sections, for which well 1 is a good example.

To identify the coal beds in the Wells 1, 2, 3, 3A and 3B the following the workflow

was used:

- Density < 2.0 g/cm3


- FEFE resistivity > Caliper/Bit *Induction resistivity

This method uses the relation between the induction logs and the focused electric

logs when inside a coal seam zone based on the high-resolution density. The density

cut-off is then improved and the net coal identification and effectiveness increased.

These two resistivity tools have distinct ways of measuring the same coal seam. First,

the induction is measuring the coal bed with a vertical resolution of approximately 1.0

metres by inducing current flow in a circular motion parallel to the tools position in the

well. The current is influenced by the conductivity of the formation. The second

resistivity measurement by the MFE tool is part of a family of tools that try to minimize

the influences of boundary formations and borehole fluids by directing the current

horizontally (perpendicular to the tools position) and reading primarily the resistivity of

the formation with a vertical resolution of 12 (with sleeve) to 15cms (sleeveless). Given

that the induction log is mainly responding to the conductivity of the boundary

formations and interbedded intervals, it was used as a base log to separate what is low

resistivity, as measured by the induction, against high resistivity organic-rich

formations (i.e. coal seams and carbonaceous mudstones where methane gas is stored)

measured by the MFE. This is done by plotting induction logs (both deep and shallow)

and FEFE logs on a logarithmic scale showing a clear separation between the typical

low values of the first against the high values of the second. To avoid carbonaceous

mudstones being wrongly considered to be coal as a result of an increase in the FEFE

147
Chapter 4

resistivity just above the induction in washout zones, a caliper/bit size ratio was added

to the cut-off. This ratio is multiplied by the induction resistivity to increase the

effectiveness of the resistivity cut-off in large caved zones. Applying this cut-off, in a

coal seam zone where the resistivity of the FEFE is higher than the induction logs, to a

high-resolution density tool, enables the non-organic beds to be filtered so that they will

not be part of the net coal estimation. Also, the FEFE log will pick values above the

induction logs in isolated coal beds that a low-density cut-off alone would not identify,

adding this correctly to the estimate as coal. In this last situation the FEFE log indicates

a minimal increase above the induction because of its limited thickness and isolated

vertical position.

4.7.1. Results in cored wells


The method was tested in well 1, well 2 (where the coal beds tend to be thicker),

well 3 and production wells 3A and 3B. Figure 4.14 is a similar figure to Figure 4.5 in

that it shows the same density cut-off but is now shown together with resistivity

responses. All the cored wells have their net coal estimation effectively increased. To

detect 90% net coal downhole the estimation effectiveness increases by 7% in well 2

and 16% in well 3. In wells with several washout zones, which is the case of well 1,

this method led to an increase in estimating effectiveness of detection of 80% of net

coal by 37% (Table 4.3).

148
Chapter 4

Reservoir/Log estimated reservoir


Reservoir
Well Gamma
described in core Neutron Sonic Density Density + Resistivity + Caliper
ray
1 80% 9% 56% 40% 49% 86%
2 90% 16% 63% 51% 84% 91%
3 90% 17% 65% 54% 77% 93%
Table 4.3 Examples of accuracy estimation of 80% (in well 1) and 90% (wells 2 and
3) of total reservoir thickness using different cut-offs.

100%
2.6 g/cm3 2.4 g/cm3 2.2g/cm3
90% 2.0g/cm3
80%
Total reservoir thickness

70%
1.8g/cm3

60%
1.6g/cm3
50%

40%

30%
Well 1 1.4 g/cm3
20% Well 2
Well 3
10%
1.2g/cm3
0%
0% 10% 20% 30% 40% 50% 60% 70% 80% 90% 100%
Reservoir rock thickness/Estimated reservoir thickness

Figure 4.14 Study of workflow cut-off value.


X-plot between the ratio of core reservoir rock thickness to estimated reservoir thickness
using a maximum wireline density cut-off together with resistivity and caliper measurements
filter against the total reservoir thickness.

Another way of showing how the resistivity coupled with the density log can

improve the estimation of coal bed reservoir sequences is shown in Figure 4.15. Notice

that by using resistivity as a filter to the density improves the resemblance of the true

coal bed thickness frequency compared to using the density cut-off alone.

149
Chapter 4

Well 1

14 14
Core Reservoir 2.0 g/cm3 Density Cut off
12 12

10 10
Frequency

Frequency
8 8

6 6

4 4

2 2
0 0

14 14
1.8 g/cm3 Density Cut off 2.0 g/cm3 Density (resistivity filter)
12 12

10 10
Frequency

Frequency
8 8

6 6

4 4

2 2

0 0

Thickness (metres) Thickness (metres)

150
Chapter 4

Well 2
18 18
Core Reservoir 2.0 g/cm3 Density Cut off
16 16
14 14
12 12

Frequency
Frequency

10 10

8 8
6
6
4
4
2
2
0
0

18 18
1.8 g/cm3 Density Cut off 2.0 g/cm3 Density (resistivity filter)
16 16
14 14
12 12

Frequency
Frequency

10 10
8 8
6 6
4 4
2 2
0 0

Thickness (metres) Thickness (metres)

151
Chapter 4

Well 3
40 Core Reservoir 40 2.0 g/cm3 Density Cut off
35 35
30 30
25
Frequency

25

Frequency
20 20
15 15
10 10
5 5
0 0

40 40
1.8 g/cm3 Density Cut off 2.0 g/cm3 Density (resistivity filter)
35 35
30 30
25
Frequency

Frequency
25
20 20
15 15
10 10
5 5
0 0

Thickness (metres) Thickness (metres)

Figure 4.15 Comparison between the entire reservoir rock sequences in well 1, 2 and 3 from core description against the estimated from 1.8 and 2.0g/cm3
density cut-off without and with resistivity filters.

152
Chapter 4

4.7.2. Results in production test wells


Since production wells were not cored, the effectiveness of using a resistivity filter

can only be tested by comparison with the nearest cored well measurements. DST

testing took place in wells 3A and 3B drilled 1,7km and 0.9km respectively from cored

well 3. In Figure 4.16 the high-resolution logs FEFE are plotted against the density

response in the Walloon Sub-group in well 3. The data points, representing intervals of

2.5cm, are coloured by what was described as coal (in black), organic rich mudstones

with more or less coal laminae (in green) and none or limited organic content

siliciclastic formation of sandstone, siltstone, volcanic tuff and mudstone (in red).

Based on the core description, black and green points can be considered as potential

methane gas sources. Note that more than 90% of coal and organic rich formations have

density readings of less than 2.0 g/cm3, but such a cut-off does not completely eliminate

red points (i.e. non-reservoir) from the estimation. Samples as low as 1.3g/cm3 may not

be coal or even organic rich mudstones, but may in fact be tuffaceous beds, or boundary

siltstones and sandstones. Industry practices today is to use deep or medium resistivity

as a backup reading in cases of borehole enlargement when the density is considered

unreliable (Mavor et al., 1990a). This is done by applying a resistivity cut-off where it

is interpreted that coal is above a fixed value. This would not be advisable. First,

because the values would need to change from tool to tool as different tools read

different values of resistivity in a coal seam zone. Second, because deep measuring

resistivity is less sensitive to thin coal beds and is less accurate to pin point its vertical

position. Lastly, the resistivity cut-off is not advisable because coal measured downhole

can show very low resistivity values due to its limited thickness and the influence of

boundary formations.

153
Chapter 4

Figure 4.17 shows the results of the proposed workflow. Basically Figure 4.17

displays the same plot as Figure 4.16 but instead uses the resistivity filters together

with a density cut-off of 2.0g/cm3. Notice how the zones selected as reservoir rock (in

black) in Figure 4.17 resemble the black and green data points in Figure 4.16. In

Figure 4.18, the density and FEFE log readings together with the interpretation of the

potential production sections from the production 3A and 3B wells show a good

resemblance to the associated cored well 3, demonstrating that the workflow generates

good results in both core and in production wells. Notice the more curvilinear shape the

data points make compared to well 3. This is another example of how different tools or

identical tools with different configurations can provide different measurements of the

formation resistivity. Comparing these two production wells, interpretation of resistivity

measurement in well 3B indicate that this well drilled thicker coal beds with less non-

coal interbeds.

In production wells, due to the borehole diameter, it was possible to run the MFE

tool with an enhancer sleeve. Comparing the readings in the cored and production well,

the tool run with the enhancer sleeve attached increased the resolution even though it

was run in a well with a larger diameter. This is also an indicator that the FEFE reading

is not affected by borehole enlargement.

154
Chapter 4

Figure 4.16 The FEFE resistivity log against the high-resolution density log responses
throughout the Walloon Coal Measures sequence identified in core description in well 3.
Each point represents 0.025m samples increments. Coal (black), carbonaceous mudstones
with coal wisps and closely spaced carbonaceous laminae (green); inorganic siliciclastic rocks
(red) - (APPENDIX I tools and logs acronyms).

Figure 4.17 The FEFE resistivity log against the high-resolution density log with estimated
net coal throughout the Walloon Coal Measures sequence by 0.025m sample increments in
well 3.
Estimated net coal (black); remaining inorganic siliciclastic rocks (red) - (APPENDIX I tools
and logs acronyms).

155
Chapter 4

Well 3A

Well 3B

Figure 4.18 The FEFE resistivity log against the high-resolution density log responses
throughout the Walloon Coal Measures sequence by 0.025m sample increments in wells 3A
and 3B respectively.
Estimated net coal (black); remaining inorganic siliciclastic rocks (red) - (APPENDIX I tools
and logs acronyms).

156
Chapter 4

4.8. Conclusions
The challenges found in estimating coal thickness in CSG using standard wireline

logs is related to the thin nature of the beds that are often less than the tools resolution,

together with common washouts .

This chapter focuses on three cored wells, one of which has significant washouts,

and two production test wells. The aim is to establish a log analysis method that could

increase the estimation accuracy of net coal in thinly bedded coal seams. This was

performed by calibrating standard log responses with core descriptions.

The estimation of net coal relying uniquely on nuclear tools results in

underestimation, if the accepted definition of coal is followed, or overestimation, if the

cut-off is broadened.

Using the contrast between the induction and the shallow focused electric resistivity

logs applied to the high-resolution density log, net coal estimation benefits from a

sharpening of the identification and definition of top and bottom boundaries of both

thick and isolated thin coal beds, while being helpful in the identification of coal in

washout zones.

This study shows that a significant optimization of the CSG net coal estimation in

both cored and production test wells can be achieved by applying a new cut-off

workflow to commonly run wireline logs in CSG exploration and production wells.

157
Chapter 5

CHAPTER 5 Evaluation of in situ Coal

Seam Gas content using wireline logs


5.1. Introduction
Chapter 4 illustrated how the interpretation of wireline logs in coal seam gas (CSG)

can be used to estimate net coal thickness. This chapter investigates the next natural

step in CSG evaluation, the estimation of gas content in CSG.

Estimation of gas content in CSG today centres on detailed analysis of core material

recovered from wells. Wireline logging of all wells drilled provides an economical

method of evaluation, allowing correlation of in situ log responses with laboratory

measurements on core recovered from only a subset of the wells. Unfortunately, gas

content in coal has no direct significant effect on any of the downhole measurements

currently used that would allow a gas charged coal to be distinguished from an under-

saturated one. Therefore there is a necessity to establish a relationship between

properties measured in core analysis and in wireline logging that are somehow

indirectly associated with gas content in CSG. Various published studies over the last

three decades have demonstrated that the best approach is to estimate gas as a function

of coal quality. This approach has led to gas content being estimated as a function of

density, not only because the core density can be accurately measured, but also because

it can be measured at high-resolution by logging tools (Calvert et al., 2011; Kim, 1977;

Mavor and Nelson, 1997; Mavor et al., 1994; Mullen, 1989, 1988; Rogers, 2007).

This chapter describes an investigation aimed at establishing a means for estimating

in situ gas content using wireline logs. To do so we consider which of the core

analytical test results can be used to correlate with wireline logs with the aim of

158
Chapter 5

generating a gas content estimate. In doing so it is important to distinguish between

those core data types that can be used quantitatively and those that should only be

treated qualitatively.

Two well data sets were used to calculate the accuracy of reservoir thickness,

organic and/or inorganic fraction, and gas content estimation, using previously

published techniques. This study then suggests where these methodologies might be

improved, so that gas-in-place results from different interpretation procedures can be

evaluated and compared.

5.2. Estimation of Coal quality


Different special core analysis tests were investigated to determine the best gas

estimation practices. Beyond the estimation of reservoir thickness, which is considered

in chapter 4, there are two other factor influencing gas content estimation: coal quality

and gas saturation as a function of gas storage capacity and pressure. The next section

reviews how these are measured and how these data are used to create a gas log from

wireline log interpretation.

After estimating the reservoir thickness the next step to obtain gas content is to

estimate reservoir quality. When referring to quality of the reservoir, due to the

intrinsic properties of coal, the tendency is to refer to the organic yield of this rock. This

quality can be calculated by considering that organic matter, in the form of macerals,

has a significant lower density than most common rock forming minerals. Estimation of

coal quality can thus be estimated by weighing samples and estimating the relative

density during routine core analysis, or estimating it from wireline density log

interpretation.

159
Chapter 5

To evaluate the constitution of coal and how it can be estimated in wireline log data

interpretation, the density of the organic and inorganic fractions though proximate and

maceral analysis testing was investigated.

5.2.1. Density of organic content

The coal macerals have their origin different vegetation tissues and it is this that is

responsible for them having different physical and chemical properties that allow them

to be differentiated. One such property is density (Choi et al., 1987; Dyrkacz and

Bloomquist, 1992; Dyrkacz and Horwitz, 1982; Stankiewicz et al., 1994; Taulbee et al.,

1989). Since organic matter is the main contributor to the total volume of coal,

variations in maceral density influence the bulk density measured in coal bearing zones.

Maceral analysis indicates that the coal seams organic composition changes (Table

5.1). Overall the Walloon Sub-group has high vitrinite content with mean percentages

per seam being above 70 %, with the exception of the Kogan and Bulwer seams.

Liptinite is the second most abundant maceral group with a mean per seam of between

17 and 31 % content by weight. Inertinite makes up only a minor portion of the maceral

composition in all but the shallowest seams, which have mean values ranging from ~8

% to 20%. Usually inertinite constitutes one of the top 2 most abundant macerals in

coal seams along with vitrinite. Compared to other worldwide Basins, the maceral

composition of the coal beds of the Walloon Sub-group is more heterogeneous with its

higher than normal liptinite content (Berbesi et al., 2009; Colson, 1991; Crosdale et al.,

1998; Mastalerz et al., 2008a; Pan et al., 2013; Saikia et al., 2014; Strpo et al., 2007;

Rao and Wolff, 1980).

160
Chapter 5

Vitrinite (% m.m.f.) Liptinite (% m.m.f.) Inertinite (% m.m.f.)


Coal seam Count
Min Max Mean Std.Dv. Min Max Mean Std. Dv. Min Max Mean Std.Dv.
Kogan 15 23.1 100 62.1 24.2 0.0 38.6 17.7 10.0 0.0 67.7 20.1 22.9

Mac.Upper 52 27.5 97.8 70.5 13.2 0.0 45.2 20.0 10.3 0.0 68.1 9.5 11.6

Mac.Lower 45 21.8 90.1 70.4 14.8 5.8 59.0 21.0 11.2 0.0 70.7 8.7 14.7

Nangram 20 52.3 91.5 74.7 11.0 6.8 47.7 24.0 11.6 0.0 8.6 1.2 2.4

Wambo 40 33.3 93.0 75.1 12.0 0.0 55.5 22.1 11.4 0.0 66.6 2.9 10.6

Iona 42 40.1 100 70.0 13.7 0.0 59.9 28.7 14.1 0.0 50.0 1.4 7.7

Argyle 85 47.5 90.9 74.3 10.8 9.0 51.7 25.4 10.7 0.0 2.0 0.3 0.5

Tangalooma Ss. 9 61.2 88.2 75.7 8.5 11.8 38.8 24.2 8.6 0.0 0.5 0.1 0.2

Auburn 29 51.6 100 72.5 13.6 0.0 48.3 27.3 13.6 0.0 1.5 0.2 0.4

Bulwer 38 17.8 100 66.7 16.7 0.0 82.2 31.4 17.0 0.0 66.7 1.9 10.8

Condamine 82 38.4 100 76.0 11.8 0.0 47.1 23.2 10.9 0.0 5.4 0.8 1.1

Table 5.1 core maceral group analysis mineral matter free results from 12 cored wells (in percentage mineral mater free basis)..

161
Chapter 5

To know to what extent the maceral composition influences the sample density in

each coal seam, the sum of ash and moisture components was extrapolated to 0% in

each well (APPENDIX F organic and inorganic content density estimation). This

provides the mean reciprocal density (1/) of the organic matter of the samples in the

coal seam. An example of this is shown in Figure 5.1.

Table 5.2 presents the density of organic matter estimated from core samples in each unit of
the Walloon Sub-group across 12 wells. As Calvert et al., (2011) concluded, if the entire data
set is extrapolated, as in Figure 5.2, the results indicate that pure organic matter has a density
3
of 1.2g/cm . Nevertheless if the data set is analysed per well the organic matter show slightly
different density values and if analysed per coal seam in each well the variation becomes even
more significant. The organic composition in shallower units have an estimated density as high
3 3
1.32g/cm , while the deeper units can have densities as low as 1.15g/cm . A reason for this
variation that was posed but not confirmed was that the fixed carbon to volatile matter ratio
found in the organic matter could be leading to the density variations due to higher volatile
matter content in lower density macerals. This ratio was also calculated in each coal seam in
each well but no clear relation was found with either the density or the respective organic
content (
Table 5.3). Another possible explanation for this significant organic content density

variation may be found in the maceral composition. Inertinite rich coal beds in the

upper units may lead to a higher density than the remaining deeper coal seams.

Therefore in CSG formation evaluation, if it is found necessary to input organic content

density during wireline log interpretation, it should be considered how the maceral

constitution is varying and how this may reflect in the bulk density readings, thus

enabling a better relationship with gas content to be achieved.

162
Chapter 5

0.85

0.80

0.75
Reciprocal core Density (g/cc)

0.70

0.65

0.60

0.55
Argyle

0.50
0 10 20 30 40 50 60 70 80
Ash + Moisture (% of weight)

0.85

0.80

0.75
Reciprocal core Density (g/cc)

0.70

0.65

0.60

0.55

0.50

0.45 Condamine

0.40
0 10 20 30 40 50 60 70 80 90
Ash + Moisture (% of weight)

Figure 5.1 An example of the extrapolation and mean absolute error of the estimate bars
(Willmott and Matsuura, 2005) of dry ash and inherent moisture percentage by weight against
reciprocal density one well in Argyle and Condamine coal seams from 24 samples from well 14.

Well 1 2 3 4 5 6 7 8 10 11 13 14 All

163
Chapter 5

wells
Kogan 1.28 1.20 1.19 1.15 1.20 1.21
Macalister
1.32 1.24 1.23 1.21 1.19 1.19 1.20 1.21 1.16 1.22
Upper
Macalister
1.30 1.20 1.23 1.17 1.21 1.20 1.20 1.21 1.21
Lower
Nangram 1.19 1.21 1.20 1.22 1.24 1.19 1.21
Coal Seams

Wambo 1.18 1.19 1.14 1.17 1.16 1.18 1.18 1.17 1.16 1.20 1.18
Iona 1.18 1.17 1.19 1.20 1.22 1.17 1.21 1.18 1.17 1.20 1.26 1.19
Argyle 1.19 1.17 1.19 1.19 1.18 1.20 1.17 1.20 1.19 1.19 1.20 1.19

Tang. Ss. 1.16 1.20 1.19 1.20


Auburn 1.20 1.15 1.20 1.18 1.20 1.21 1.17 1.19
Burwer 1.22 1.21 1.31 1.20 1.18 1.20 1.22 1.19 1.19 1.24 1.20

Condamine 1.21 1.18 1.18 1.21 1.20 1.21 1.24 1.18 1.22 1.19 1.22 1.21 1.21
Upper
1.27 1.27
Measures

Juandah
Coal

Lower
1.18 1.18
Juandah
Taroom 1.19 1.18

Mean 1.20 1.19 1.19 1.19 1.19 1.20 1.19 1.19 1.20 1.19 1.20 1.21

Table 5.2 Estimated core organic matter density in each coal seam unit in 490 samples
from 12 wells. For mean absolute error of the estimate and sample count see APPENDIX G
Error estimation.

Coal Seam / Well 1 2 3 4 5 6 7 8 10 11 13 14


Kogan 1.019 1.094 0.611 0.973 0.496
Mac.Up. 1.054 0.892 0.938 0.608 0.757 0.742 0.839 0.720 0.775
Mac. Lo. 0.807 0.919 0.755 0.644 0.711 0.880 0.850 0.673
Nangram 0.802 0.789 0.824 0.835 0.538 0.966
Wambo 0.843 0.828 0.494 0.741 0.651 0.859 0.819 0.356 0.677 0.638
Iona 0.714 0.901 0.774 0.777 0.679 0.693 0.842 0.809 0.545 0.632 0.535
Argyle 0.840 0.782 0.652 0.851 0.574 0.767 0.689 0.852 0.816 0.794 0.763
Tang. Ss. 0.814 0.900 0.925
Auburn 0.874 0.749 0.683 0.761 0.844 0.754 0.541
Burwer 0.812 0.875 0.697 0.842 0.710 0.552 1.027 0.607 0.574 0.767
Condamine 0.841 0.863 0.823 0.908 0.854 0.829 0.880 0.986 0.952 0.675 0.854 0.759

All Coal Seams 0.830 0.866 0.738 0.892 0.699 0.952 0.677 0.842 0.987 0.731 0.759 0.711

Table 5.3 Fixed carbon to volatile matter ratio in each coal seam unit in 490 samples from
12 wells. For standard deviation of the estimate and sample count see APPENDIX G Error
estimation

164
Chapter 5

5.2.2. Density of inorganic content

There are to two inorganic components in coal: inorganic mineral content and pore

fluid (moisture).

Moisture content has a very weak (r=0.25) relationship with the reciprocal density

(Figure 5.2). As expected, moisture content tends to decrease with depth due to

compaction and maturation of the coal beds (Figure 5.3).

The inorganic mineral content in the core samples can be found as either discrete

thin layers separating coal beds or within the coal beds. In the coal seams in the

Walloon Sub-group, the mineral content of the coal is largely clay (mainly kaolinite,

illite, smectite and montmorillonite) and quartz, but also with minor percentages of

plagioclase and potassium feldspar. Minor concentrations of calcite are found as cleat

filling in association with authigenic kaolinite and smectite. Volcanic tuff usually forms

very thin interbeds within a coal seam, composed mainly of smectite minerals.

Considering ash content as indicative of mineral content, Figure 5.2 shows a strong

negative correlation with reciprocal density. The extrapolation of ash density to 100%

produces a matrix density of 2.68g/cm3. Because of the variable mineralogy together

with the lower fraction found in the core samples compared to organic content, the

estimated mineral density can vary significantly from coal seam to coal seam; examples

of this for two well is shown in Table 5.4.

165
Chapter 5

Organic content Moisture Ash


0.9

0.8
Reciprocal density (cm3/g)

0.7

0.6

0.5

0.4

0.3
0 10 20 30 40 50 60 70 80 90 100
% by weight

Figure 5.2 Dry organic matter, moisture and ash from proximate core analysis against
reciprocal density from 491 core samples taken from 12 cored wells.

Moisture (% by weight)
0 2 4 6 8 10 12 14
0

100

200

300
Depth (metres)

400

500

600

700
Upper Juandah samples
800 Lower Juandah samples
Tangalooma Sandstone
Taroom samples
900
Figure 5.3 Moisture content estimated in proximate analysis against depth from 491 core
samples taken from 12 cored wells.

166
Chapter 5

2 3
Coal Seam / Well Count Density MAE Count Density MAE

Mac.Up. 2 2.54 0.012 4 2.59 0.007


Mac. Lo. 2 2.10 0.003
Wambo 9 2.68 0.011 4 2.86 0.007
Iona 3 2.74 0.002 6 2.68 0.009
Argyle 6 2.77 0.008 7 2.65 0.023
Burwer 2 2.81 0.001 3 2.21 0.010
Condamine 8 2.76 0.017
Table 5.4 Examples of the estimated ash density and mean absolute error of the estimate
from 57 core samples in each coal seam unit in wells 2 and 3.

5.2.3. Wireline density readings

Calvert et al. (2011), referring to these same Walloon coal seams, state that they had

an average density of 1.5 g/cm3. From the same extensive wireline log dataset analysed

in chapter 3, the mean wireline density of the coal beds found in the 12 wells analysed

was 1.66 g/cm3. This estimation determined here is believed to be more credible

because it is calculated through wireline log to core correlation, while the estimate from

Calvert et al., (2011) was determined by considering that coal is only found in units

with a wireline density less than 1.8 g/cm3.

Densities of less than 1.22 g/cm3 are not expected in coal samples because coals are

not composed of pure organic matter, however they can be found in a small percentage

in wireline density data in coal zones (two examples can be found in the logs in Figure

5.30 and Figure 5.31). This is explained by the higher resolution measurement read by

high-resolution density tool downhole compared to measuring, and averaging, over a

core sample. Calvert et al. (2011) also identified these lower density coals in wireline

log readings, dismissing them as an error related with the Z/A correction (Samworth,

1992). However, because coal seams are made mainly of organic matter, with densities

167
Chapter 5

that, as demonstrated here, can be less than 1.2 g/cm3, it is feasible that very pure thin

coal beds can have such low densities.

5.3. Gas content and saturation


The main objective of recovering and testing core samples in CSG exploration is to

estimate gas content and the gas storage capacity of the coal, and to determine how

these relate to the physical properties of coal as measured by wireline logs at a regional

scale.

Core analysis has determined that all the core samples studied here are gas charged.

The measured core sample density was between 1.22g/cm3 and 2.66 g/cm3. The higher

density values indicate the presence of rock types including carbonaceous mudstones,

tuffaceous beds and, more rarely, other siliciclastic inorganic interbeds within, or

bordering, the coal seam. The gas content is associated with the presence of organic

matter in the formation, therefore although the carbonaceous mudstone has less gas

storage capacity, it may also be gas charged due to the presence, of organic compounds.

Ultimately, gas content depends on storage capacity and that is dependent on the

available organic matter, reservoir pressure and the degree of gas saturation in the coal

seams.

An Adsorption Langmuir Isotherm Analysis Test is conducted over a composite of

samples taken from the same seam. This analysis is a fundamental procedure to relate

gas storage capacity to pressure and a method for estimation of gas saturation (Calvert

et al., 2011; Mavor and Nelson, 1997). Sorption isotherm analysis is conducted with

equilibrium of moisture. Analysis of the core dataset indicates the samples are

commonly undersaturated, i.e., under 100% (Table 5.5).

168
Chapter 5

Coal Seam / Well 2 3 4 5 6 7 8 10 11 13 14


Kogan 44% 73% 99% 63% 224% 80% 84%
Mac.Up. 59% 57% 103% 92% 65% 158% 69% 98% 63%
Mac. Lo. 62% 94% 97% 75% 187% 83% 80% 65%
Nangram 110% 87% 60% 123% 76% 50% 93% 86% 50%
Wambo 83% 83% 100% 82% 311% 109% 65% 280% 71% 58%
Iona 93% 102% 225% 81% 69% 207% 96% 78% 107% 86% 80%
Argyle 73% 92% 81% 99% 75% 91% 75% 74% 73% 99% 92%

Tangalooma Ss. 103% 40% 82% 81%


Auburn 69% 101% 64% 67% 58% 87% 82% 101%
Burwer 103% 96% 142% 42% 33% 120% 72% 155% 77%
Condamine 80% 73% 78% 38% 65% 77% 54% 60% 77% 54% 63%
Table 5.5 estimated gas saturation from core data from 452 from 11 wells. For standard
deviation of the estimate and sample count see APPENDIX G Error estimation.

The values greater than 100 %, indicate that a sample has desorbed more gas during

the desorption testing than the estimated gas storage capacity from the Adsorption

Langmuir Isotherm Analysis Test. Such overestimation is a clear indication of under-

testing of Adsorption Langmuir Isotherm Analysis, a common practice in coal seam gas

exploration (MacLennan et al., 1995; Mares et al., 2009a). For example, most of the

samples from Well 7 have more gas than the estimated storage capacity.

The Adsorption Langmuir Isotherm Analysis Test procedure has three main flaws

related to the fact that the samples are made of a composite of all samples taken from

the coal seam they are a part of. First, the sample is crushed, destroying, at least in part,

the original structural pore format that the gas is adsorbed onto. Second, the samples are

a mixture of different maceral types that may, or may not, be together under in situ

conditions and may, or may not, be in contact with the mineral matter. Therefore

saturation estimation can be over- or under-estimated due to the composite sample not

being representative of the in situ coal composition. For example, a sample made of a

coal interbedded with organic-free macroporous tuff layers results in adsorption

169
Chapter 5

capability higher than an organic and inorganic composite sample as this translates into

coal matrix surface area where the gas is adsorbed.

Finally, the equilibrium of moisture used during the test may not be representative of

the moisture in situ, limiting the gas adsorption of the sample. Equilibrium of moisture,

estimated during Adsorption Langmuir Isotherm Analysis is believed to provide a better

estimate of the true bed moisture (Crosdale et al., 2008; Langenberg et al., 1990; Mavor

et al., 1990a; Nelson, 1999; Rogers, 2007). Samples may have lost formation fluid

during coring and therefore this may be less than in situ moisture; in fact comparing

Figure 5.3 with Table 5.6 moisture estimated from proximate analysis is shown to be

lower than the estimated equilibrium of moisture. Nevertheless, Mavor et al. (1994)

suggested that possible flaw in the Langmuir Isotherm Analysis is the risk of making

the test with a equilibrium moisture content superior to the moisture in situ. Since

increasing moisture diminishes the gas storage capacity of coal, gas content storage

capacity could therefore be underestimated. Moreover, one of the motives that could

lead to the sample being tested with an equilibrium of moisture greater than the

moisture found in situ is the fact the sample is tested as a composite of samples and a

mixture of rock types that may not be together in situ.

Coal Seam / Well 2 3 4 5 6 7 8 10 11 13 14 Average


Kogan 10.3 13.8 7.3 14.5 9.6 9.2 10.8
Mac.Up. 13.0 8.5 10.2 11.5 8.6 8.8 4.5 8.4 9.7 9.2
Mac. Lo. 12.2 10.1 9.3 13.1 8.4 10.6 5.5 9.4 6.8 9.5
Nangram 9.6 11.9 8.5 11.2 3.7 9.9 4.3 6.8 5.8 8.0
Wambo 9.7 7.9 11.5 7.0 11.4 4.1 7.4 7.9 9.5 6.5 8.3
Iona 8.6 11.9 12.1 5.9 8.6 5.0 7.6 8.4 6.8 6.9 8.2
Argyle 9.2 8.0 8.4 9.8 7.2 7.8 3.5 7.6 7.5 7.0 6.8 7.5

Tangalooma Ss. 9.2 8.0 10.1 8.9 7.0 8.6


Auburn 8.5 9.1 6.4 3.9 7.0 5.7 5.8 6.6
Burwer 7.2 12.6 9.1 8.5 5.7 9.2 7.9 5.2 9.2 8.3
Condamine 7.9 8.8 7.2 7.9 8.0 9.3 4.6 6.8 6.0 5.9 7.3 7.2
Table 5.6 Equilibrium of moisture from Adsorption Langmuir Isotherm Analysis core data.

170
Chapter 5

For these reasons the estimation of in situ gas content relies on the observed gas and

not on Adsorption Langmuir Isotherm Analysis. The results of this procedure are

therefore only largely qualitative (saturated or non saturated coal seam), though they are

still useful. The identification of undersaturated coal seams, besides being undesirable

in terms of CSG productivity, is also important in gas content estimation. The closer the

gas content is to gas storage capacity the easier it will be to estimate gas-in-place,

because it will be easier to associate it with measurable coal physical properties (coal

thickness, quality and coal depth). This is demonstrated in the next section where the

results show that for deeper coals that are undersaturated or do not have increasing gas

content with depth, it is clearly more difficult to estimate gas content from wireline log

interpretation.

5.4. Testing gas content estimation methodologies


Several authors (Calvert et al., 2011; Hawkins et al., 1992a; Kim, 1977; Mavor and

Nelson, 1997; Mavor et al., 1994; Mullen, 1989, 1988; Rogers, 2007) have defended

methods to interpret wireline measurements in coal to assess the coal thickness, quality

and gas content.

A review of Kim's (1977) methodology was found not to be applicable to the coal

seams analysed in this work. Gas content as a function of depth and the ratio between

fixed carbon and volatile matter is not applicable to the coal seams studied. As

discussed in Chapter 3, coal rank, although associated with gas storage capacity is not

related to gas content due to the deeper coal seams of the Taroom Measures not being

saturated. Also there is no apparent relationship between the maceral content, or fixed

carbon to volatile matter ratio, and the gas content. What is apparent is that the amount

of organic matter is important in terms of gas content estimates. Also, for these reasons,

171
Chapter 5

plus the observation that it is unreliable to use Langmuir Isotherm Analysis in a

quantitative way, the methodology defended by Hawkins et al. (1992) does not fit the

data set analysed. Methodologies of Mavor et al., (1994) or Mullen, (1988) that

defended a linear relationship between bulk density of all coal seams with gas were

disregarded due to the variable gas saturation of the Walloon Sub-group coal seams and

the different densities of the organic compounds that form them.

After a literature review, the methodologies that were considered to be applicable to

the Walloon Sub-group coal seams are Calvert et al. (2011), Mavor and Nelson (1997)

and Rogers (2007). These are briefly explained and results interpreted in sections 5.4.1

to 5.4.3 . Furthermore, after testing these methodologies, and considering that a new

way to estimate coal thickness was demonstrated in chapter 4, new methods are

presented and tested in section 5.4.4.

The methodologies investigated and the ones tested considered that these coal seams

are thinly bedded, with variable gas saturation, and therefore the views expressed here

are adapted to this type of resource. Due to significant differences between coal seams

in terms of coal composition, gas content and permeability it is common practice to

evaluate coal seam gas in separate reservoirs. Typically these reservoirs are coal seam

groups and in the geological sense represent sedimentary cycles where coal is

interbedded with sandstones, siltstones and mudstones. In the study by Calvert et al.

(2011) these reservoir groups are the Walloon coal measures. Mavor and Nelson,

(1997) and Rogers (2007) also separated the coal seams into reservoirs.

Wells 2 and 3 were selected due to them having a complete dataset, enabling the

application of any of methods tested. Only wells 2 and 3 in the dataset have core data

together with wireline log data that includes the shallow focused electric and induction

172
Chapter 5

resistivity logs; these are fundamental to coal thickness estimation as previously

explained in chapter 4.

For each methodology the following 4 aspects were analysed and compared:

1) Reservoir thickness estimation

The first step of formation evaluation using wireline measurements in coal seams is

to define the reservoir. The methodologies may be differentiated not only on the basis

of which rock types are considered to be gas producible but also which cut-offs are used

to identify them. Having the description of the core it was possible to compare which

rock types were selected depending on the cut-off used in each methodology. The

results are show in Table 5.9 and Table 5.10.

2) Reservoir quality estimation

In this study when mentioning reservoir quality we refer to organic content of the

coal, in percentage of weight or percentage of volume. Depending on the methodology,

the estimated reservoir quality from wireline log interpretation may or not be influential

in the gas-in-place calculation, although in some form all the methodologies present

ways to estimate coal quality. It is common practice to create linear equations relating

core bulk density and the different components of proximate analysis. To determine the

accuracy of estimation of reservoir quality the estimations were compared with

proximate analysis provided from core analysis.

173
Chapter 5

3) Gas content estimation

Gas content analysis was compared with desorbed gas from core samples and the

estimate from wireline logs. Two gas logs were created. The first considering that all

rock types can contain gas (total gas content), and a second where only the rock types

selected after the application of the cut-off contribute to gas content (producible gas).

The comparison between these two gas logs indicates the influence of the cut-off used

in the gas-in-place estimation (APPENDIX H density and gas content equations).

Considering that the core samples tested in wells 2 and 3 were mainly of coal lithology,

the coal gas content plots should show minimal differences to the total gas content plots

if the cut-offs that each methodology has defined to identify coal are accurate. The

general trend in gas content cross-plots estimation in coal are of course

underestimations compared to total gas estimation due to the fact that gas charged

carbonaceous mudstones and coal are to some extent cut out of the calculation.

4) Gas in place

Finally gas-in-place estimation results are compared using the different

methodologies. After reviewing the literature of the formation evaluation in CSG it is

expected that the gas-in-place estimation will be significantly different between the

methodologies, due partly to different interpretations of this resource, but also due to

the different interpretations of the well data.

The results of this analysis in the shape of tables, figures, examples of interpretation

plots and a resume of methodologies and the general results are displayed in pages 188

to 206 together with discussion below.

174
Chapter 5

5.4.1. Mavor and Nelson (1997)


The Mavor and Nelson (1997) methodology is based on the estimation of ash

content. Core ash and organic content reciprocal density values are determined initially

through proximate analysis extrapolation to 100% ash content and 100% ash plus

moisture content in each coal seam. These are then applied to the density log using

equation 5-1 where it is also added equilibrium of moisture, estimated from core

analysis.

Reservoir rock thickness can be any rock with a value less than the ash density

considered. In this method reservoir rock is not only coal but also sequences of coal

with carbonaceous shale. This is a clear difference compared to the following

methodologies explained here. Other methods choose to identify coal as the only source

of gas, with the argument that only this rock type has enough permeability, supported

by cleats, to be commercially viable in thinly bedded coal seams. The estimation of gas

content using the Mavor and Nelson (1997) methodology shows only total gas as this

methodology targets literally the gas in place.

After the reservoir rock is identified, the gas content log is then calculated through a

negative relationship with total inorganic matter following equations 5-2 and 5-3.

175
Chapter 5

5-1
1 1 1 1
( +  &, (  +
) *  )
&' 
1 1

) '

5-2
-  -) 1  &'  &, 

5-3
  1359.7
c c

Where:
Wa Ash content in the coal, % by weight
Wwe Water in equilibrium of moisture % by weight
o Organic matter density g/cm3
b Bulk density g/cm3
w Water density g/cm3
a Ash content density g/cm3
Gc Gas content as received m3/ton
Gco Gas content dry ash free m3/ton
G Gas-in-Place, scf (*1359.7 for scf )
A Reservoir Area, acres
h Thickness, feet
c Average in situ coal density, g/cm3
 c Average in situ gas content, scf/ton

Results

Mavor and Nelsons methodology of estimating gas content using wireline log

interpretation is based on the negative relationship between total inorganic matter and

gas content. This method stretches the reservoir thickness considerably in both wells to

a point that well 2 has almost no effective cut-off by considering that the net-coal

thickness is of 390.73 meters when in the core description it is only 40.03m (Table

5.9).

The results using this method show it to overestimate inorganic matter compared to

other methodologies (Figure 5.6 the estimated inorganic content is many times higher

than the actual core proximate analysis data), therefore in the next step this estimate

leads to an under-estimation of gas content in individual beds. This is clearly observable

176
Chapter 5

in well 2 (Figure 5.16 - the estimated gas content is many times lower than the actual

core data). Nonetheless, in spite of this methods tendency to underestimate gas content

in coal beds individually, the gas-in-place volumes estimated are much greater than any

other method due to the point of view that all organic rich beds may be considered as

part of the reservoir and not just coal. An example of the gas log created is shown in

Figure 5.29 and the gas-in-place estimated in Table 5.12 and Table 5.13 Rogerss,

Mavors, Calverts and the new workflow results from in well 3.

5.4.2. Rogers (2007)


In Rogers (2007) coal seam method the author first defines strict coal identification

cut-offs. These were applied to the high-resolution density log (<2.0 g/cm3), natural

gamma ray log (< 60 API), neutron-porosity log (> 50%), sonic log (> 80 s/ft) and

shallow focused electric log (> 50m2/m). This methodology for estimating gas

reserves has some similarity to Mavor and Nelson (1997) which by using equation 5-2,

derives an estimated ash content together with the equilibrium of moisture to estimate

gas content. Besides differences in the cut-off, this method is distinct when it comes to

estimating ash content through a relationship with density, compared with Mavor and

Nelson (1997) reciprocal density. Furthermore Rogers (2007), referring to Mullen

(1989), advocates an independent relationship between density and the proximate

analysis components. For comparison purposes both proximate analysis and gas content

equations were created in each Well 2 and 3 and are shown in Figures 5.4 and 5.5.

177
Chapter 5

The gas reserves are estimated using the following equations:

5-4
./!0 1/2! 3 435 67!53 3135  835! 3 70 9326:1;  <=  ? ./!0 @
:<2366  ?

5498382.88 (coal tonnage per C / <= 

5-5
!6 :2 40!<3  </!0 1/22! 3 435 67!53 3135 ? !6 </21321  <= 

100 100

Estimated proximate analysis (% by weight)


Estimated proximate analysis (% by weight)

90 90
80 80
70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8 1.2 1.4 1.6 1.8 2 2.2 2.4 2.6 2.8
Core relative density (g/cm3) Core relative density (g/cm3)

Organic content Ash Moisture Organic content Ash Moisture

Figures 5.4 and 5.5 Relationship between core relative density and core proximate analysis
in all samples from well 2 and 3.
The plots represent linear trendlines used to estimate proximate analysis through Rogers
(2007) methodology.

Results

In the net coal estimation, the effect of Rogers (2007) cut-off is particularly negative

in terms of coal identification due to the fact the coal seams are thinly bedded. As

Table 5.9 and Table 5.10 show, following this methodology the application of the cut-

off clearly underestimates the net-coal thickness. Rogerss interpretation only identifies

19.2 metres (out of 40.03m) and 9.58m (out of 30.45m) of net-coal from wells 2 and 3.

Coal quality estimation is an improvement compared to Mavor and Nelson (1997)

(Figure 5.7).

178
Chapter 5

Figure 5.17 shows gas content estimation following Rogers (2007) methodology

against the gas content found in core samples. Using a cut-off in this methodology aims

to find producible gas and therefore it disregards rocks containing minor volumes of

gas. In this case the estimation of gas content results in margins of error relatively low,

although after the application of the cut-offs the estimation clearly underestimates the

CSG resource due to its cut-off parameters; this effect is striking in well 3 (Figure

5.23).

5.4.3. Calvert et al. (2011)


This 2011 publication has particular interest since it focuses on the same coal seams

analysed in this study. The reservoir rock is coal and to identify it Calvert et al. (2011)

establishe an upper limit for coal density at 1.8 g/cm3. As the authors state, this

methodology is highly influenced by Mavor and Nelson (1997) although difference are

clear in the coal thickness and how the gas content is estimated. Additionally, to

estimate proximate analysis using wireline logs the authors used a linear equation for

ash content and ash together with moisture content with the variable being reciprocal

density (equations 5-6 to 5-8); these were applied to all coal seams in both wells.

5-6
&'  226.8 ? D1# E  182.3

5-7
&'  &   223.1 ? D1# E  185.5

5-8
&  &'  &   &'
Where:
bulk density, g/cm3
Wa weight of ash, %
Ww weight of water, %

179
Chapter 5

To estimate gas content through Calvert et al.'s (2011) method a relationship

between reciprocal density and gas content is created in each cored well and in each

Walloon coal measure. This is due to the fact the relationship changes on a well to well

basis and in each Walloon coal measure, examples of this are shown on Figure 5.11.

Results

The methodology of Calvert et al. (2011) is shown to fit the Walloon coal measures

more than previous ones, although proximate analysis does not seem not be an

significant improvement compared to other methodologies (Figure 5.8). Interestingly

again, well 2 has a higher margin of error than well 3 when estimating gas content,

indicating that there are differences in gas content and/or in the nature of coal that

density, or reciprocal density, does not reflect clearly.

Another interesting effect is found regarding the effect of the cut-off in gas content

estimation. The Calvert et al. (2011) method sees minimal differences between the total

gas and coal gas content estimated in well 2 while there is a significant difference

between these two gas estimations in well 3 (Figure 5.18 and Figure 5.24). This is due

to the application of the cut-off in the estimation of gas content in coal only, by

influencing the accuracy more in well 3 than in well 2. This indicates that although the

estimate of gas content can be accurate the identification of productive zones could be

improved.

5.4.4. New methods tested


The new methodologies tested here, as in previous chapters, follow the hypothesis

that the reservoir rock is both coal and carbonaceous mudstones rich in coal laminae.

Only rock formations with high organic matter can sustain economically producible gas

180
Chapter 5

in thinly bedded formations. These rocks contain gas but may also have natural

permeability, provided by natural fractures or cleats and planar surfaces as in coal

banding or coal laminae in carbonaceous mudstones.

From core analysis it was possible to conclude that maceral content, gas content and

gas saturation vary from seam to seam in the Walloon Sub-group coals. Furthermore,

the densities of the organic matter and the mineral matter also change. Therefore the

workflow should take this into account by separating the well coal sequences into

separate reservoirs.

At this stage, through the process of creating a new methodology, two questions still

remained. Firstly, what is the best estimate of coal quality (i.e. organic content) that

could be estimated in wireline log analysis that could be better related with gas content?

And, secondly, should it be taken into account that gas content increases with pressure

in the Juandah measures and decreases in the deeper seams? After the net coal

estimation step is taken, to answer these two questions four methods were tested. Two,

named W (for weight) and WP (for weight & pressure) provide gas content logs in

m3/tonne. V (for volume) and VP (for volume & pressure) methods were created

estimating first organic matter volume previously to gas content in v/v units. New

methodologies tested are explained in more detail below:

Method W: Similar to Calvert et al. (2011), a direct relationship was calculated

between reciprocal core density and gas content (equation 5-9; APPENDIX H density

and gas content equations) with the distinction that in this case the separation of

individual reservoirs was processed to find the best fit depending of changes in the

maceral content density and gas content, this may require separating the Walloon Sub-

group in Measures or in coal seams.

181
Chapter 5

5-9

-  ! ? D1# E  F

Where:

- In situ gas content, m3/tonne


bulk density, g/cm3

Method V: this methodology uses an estimation of organic matter volume before

estimating gas content volume. The estimation of organic volume was obtained by first

calculating organic matter density by using similar plots to that shown in Figure 5.1 for

each coal seam in each well from core proximate analysis (Table 5.2). With these

organic densities calculated from each seam it is possible to estimate the organic

content volume (i.e. coal quality equation 5-10), after which it is correlated with gas

volume from core desorption tests so they could be applied to create a gas volume log

(equation 5-11; APPENDIX H density and gas content equations). To create a gas

volume log a linear relationship is first defined between reciprocal core density and

organic volume fraction, again with core data, so that afterwards the same relationship

could be applied to wireline density.

The purpose of the V methodology is to test if a better relationship can be achieved

by using a percentage of volume, instead of m3/tonne, and also by relating organic

matter, which is the main component in coal and not ash, directly to gas content. In

cases where the wireline density measurement is lower than the extrapolated core

organic matter (when the density tool measurement quality is considered to be reliable)

it was considered that this is due to a very low maceral content density and to close to

no mineral content. In these cases G)HI' JK- was considered to be close to 100%, where

the coal is only formed by organic matter and moisture.

182
Chapter 5

5-10

H
G)HI' JK-  &)
)

5-11

-  ! ? D G)HI' JK- -)JL,JL E  F

G)HI' JK- bulk density, v/v


&) Weigh percentage of fixed carbon plus volatile mater, dec
H Core relative density, g/cm3
H Organic density, g/cm3
- In situ gas content, v/v

WP and VP methods: these are almost identical to the previous W and V methods

respectively with the distinction that the gas content also takes into account the pressure

(M) of the coal in situ using the following the equations 5-12 to 5-15 (APPENDIX H

density and gas content equations). The introduction of pressure is simply due to the

idea that two identical coal seams with similar saturation values can have different gas

content due to different relationships with pressure. The usage of reciprocal pressure in

the deeper coal beds could be considered counter intuitive, since gas adsorption

increases with pressure, but the goal here is to estimate in situ gas content, not gas

adsorption capacity, which the data show actually decreases with pressure in deeper

coals in Walloon Sub-group.

183
Chapter 5

Method WP

5-12 Juandah coal measures:

-  ! ? ND1# E ? MO  F

5-13 Taroom coal measure & Tangalooma Sandstone coal seam:

-  ! ? P1# ? MQ  F

Method VP

5-14 Juandah coal measures:

-  ! ? N G)HI' JK- -)JL,JL ? MO  F

5-15 Taroom coal measure & Tangalooma Sandstone coal seam:

-  ! ? N G)HI' JK- -)JL,JL ? D1#M EO  F

Results

Estimation of coal quality is shown to be advantageous if it is calculated by

volume compared to any other method calculated in percentage by weight

(Figure 5.9 and Figure 5.10).

It seems to be beneficial to establish different relationships between coal

quality to in situ gas within each single Walloon coal measures. Even within

one Walloon coal measure gas saturation can vary significantly. An example

of this can be seen in the Lower Juandah in well 2. It was beneficial in well 2

to separate the Juandah measure into upper and lower (red and blue points

respectively in (Figure 5.12 to Figure 5.15) to create the best gas content

correlation. Curiously, in method VP this was not required and a good trend

was found for the entire measure (Figure 5.15).

184
Chapter 5

The Taroom measure is, as expected, the most difficult to estimate gas

content for, due to the variable gas coal saturation found in this measure and

it does not relate well with coal quality.

It seems clear that it is beneficial to estimate the organic matter volume of the

reservoir rock (Figure 5.20; Figure 5.22; Figure 5.26; Figure 5.28). The

variability of the nature, particularly the density, of the maceral content

seems to affect the relationship between the density and the gas content. The

estimation of the maceral density as a step to estimating the gas content

seems to benefit the accuracy of gas-in-place calculations.

Pressure helped to create better fit trendlines in both the WP and VP methods

indicating that this is a factor to consider when estimating gas content in situ

and should be used when creating gas content logs (Figure 5.21; Figure

5.22; Figure 5.27; Figure 5.28).

The following two tables (Table 5.7Table 5.8) show the mean absolute

percentage error for each estimation (Hyndman and Koehler, 2006). Notice

how the VP method is shown to have the lowest error when estimating total

gas content. Also the coal quality in V/V used in both V and VP methods is

shown to have the lowest mean absolute percentage error.

185
Chapter 5

WELL 2
Mavor Rogers Calvert W V WP VP
Coal quality
18.4% 9.2% 10.3% 9.4% 5.5% --- ---
MAPE
Total Gas content
23% 24.1% 26.1% 23.9% 14.6% 25.4% 13.2%
MAPE
Coal gas content
--- 31.4% 27.5% 22.1% 14.2% 24.1% 13.7%
MAPE
Table 5.7 Estimated Mean Absolute Percentage Error for each method for coal quality, total
gas content and coal gas content for well 2.
WELL 3
Mavor Rogers Calvert W V WP VP
Coal quality
15.1% 24.6% 18.5% 17.9% 12.6% --- ---
MAPE
Total Gas content
26.6% 21.6% 29.6% 29.5% 19.9% 33.5% 16.3%
MAPE
Coal gas content
--- 64.2% 26.6% 26.9% 20.0% 25.6% 18.7%
MAPE
Table 5.8 Estimated Mean Absolute Percentage Error for each method for coal quality, total
gas content and coal gas content for well 3.

From the different new methodologies tested the one that has best results was VP,

leading to an improved in situ gas volume calculation. A resume of the Mavor and

Nelson (1997), Calvert et al. (2011), Rogers (2007) and the new workflow is shown in

Table 5.11 and the resulting Gas-in-place in well 2 and 3 in Table 5.12 and Table 5.13

Rogerss, Mavors, Calverts and the new workflow results from in well 3.

The new method does not provide the highest gas in place volumes of all the

workflows discussed here, but it provides nonetheless the most reliable workflow to

estimate producible gas in place compared to others. The workflow methodology

created has two critical steps. Firstly, the definition of net coal and, secondly, a

following step to estimate gas content by creating an in situ gas log, all of this using

wireline logs and supported by core description and core analysis. The basic workflow

to create this advised workflow is show in Figure 5.32. This workflow is meant to be

the initial process used in a cored exploration well, so that afterwards it can be applied

to nearby production wells without necessarily needing core and/or core analysis input.

186
Chapter 5

RESERVOIR THICKNESS ESTIMATION TABLES

The two following tables display the results of net-coal thickness using the different

methodologies. In the Core column the total thickness of the several lithologies and

coal lithotypes found in core in wells 2 and 3 are shown. In the columns labelled

Rogers, Mavor, Calvert and New the results of cut-offs for each of these different

methodologies are displayed. For example, in the case of Dull minor bright coal total

thickness intercepted by the well 2 was 18.58 metres, and the application of Rogerss

method to wireline log interpretation considers that only 12.95m of these would be coal,

while Mavor would consider all to be coal, Calvert 16.83m and the New method

18.20m. The tables also display the total coal and net-coal estimation for the entire well.

For example, the estimated net-coal using Mavors in well 3 is of 65.65meters but

analysing the core description it is known that only 30.38m actually are coal. It can then

be concluded that Mavors methodology in this well, although identifying almost all

coal beds, also overestimates the reservoir thickness. In other words, the optimal

methodology should have both the highest total coal (for example well 2 this is 40.03

meters) and the lowest difference between estimated net-coal and total coal.

187
Chapter 5

Well 2 lithologies and net-coal cut-off results (metres)


Rock type Core Rogers Mavor Calvert New
Bright Coal 0 0 0 0 0
Bright banded Coal 0 0.00 0.00 0.00 0.00
Dull & Bright Coal 0.65 0.28 0.65 0.63 0.65
Dull banded Coal 4.28 2.85 4.28 3.95 4.25
Dull minor bright Coal 18.58 12.95 18.58 16.83 18.20
Dull Coal 3.40 1.18 3.40 2.15 3.33
Stony Coal 3.9 1.13 3.90 2.30 3.55
Coal Laminae 9.23 0.83 9.23 3.15 6.63
Carbonaceous Mudstone 5.38 0.28 5.38 1.20 1.93
Mudstone 23.30 0.40 23.08 0.63 0.95
Tuff 0.23 0.00 0.13 0.00 0.00
Siltstone 109.15 0.03 108.25 0.55 1.03
Sandstone 219.68 0.00 213.63 0.23 0.15
Siderite 1.08 0.00 0.25 0.00 0.08

TOTAL Coal 40.03 19.20 40.03 29.00 36.60

ESTIMATED net-coal ---- 19.90 390.73 31.60 40.73


Table 5.9 398.8 meters of core description against the results of reservoir rock cut-off for
each methodology in well 2.

Well 3 lithologies and net-coal cut-off results (metres)


Rock type Core Rogers Mavor Calvert New
Bright Coal 0 0 0 0 0
Bright banded Coal 0.13 0.05 0.13 0.10 0.13
Dull & Bright Coal 0.63 0.18 0.60 0.53 0.60
Dull banded Coal 5.00 1.20 5.00 4.30 4.95

Dull minor bright Coal 13.78 6.90 13.78 12.38 13.50


Dull Coal 2.60 0.85 2.60 1.73 2.50
Stony Coal 3.18 0.23 3.13 1.68 2.95
Coal Laminae 5.15 0.18 5.15 1.68 3.83
Carbonaceous Mudstone 5.00 0.18 4.95 1.53 1.48
Mudstone 45.83 0.05 13.30 1.40 0.98
Tuff 3.43 0.43 2.88 1.48 0.28
Siltstone 67.40 0.00 5.53 0.03 0.00
Sandstone 147.83 0.03 8.58 0.00 0.15
Siderite 3.15 0.00 0.05 0.00 0.00

TOTAL Coal 30.45 9.58 30.38 22.38 28.45

ESTIMATED net-coal ---- 10.25 65.65 26.80 31.33


Table 5.10 Core description against the results of reservoir rock cut-off for each
methodology in well 3.

188
Chapter 5

RESERVOIR QUALITY ESTIMATION

Mavor and Nelson (1997)

100 100
Estimated proximate analysis (% by weight)

Estimated proximate analysis (% by weight)


90 Well 2 90 Well 3
80 80
70 70
60 60
50 50
40 40
30 30
20
20
10
10
0
0
0 10 20 30 40 50 60 70 80 90 100
0 10 20 30 40 50 60 70 80 90 100
Core proximate analysis (% by weight)
Core proximate analysis (% by weight)

Ash + Moisture Ash + Moisture

Figure 5.6 Comparison between inorganic content from core proximate analysis against
Mavor and Nelson (1997) wireline log estimation.

Rogers (2007)

100 100
Estimated proximate analysis (% by weight)

Estimated proximate analysis (% by weight)

90 Well 2 90 Well 3
80 80
70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Core proximate analysis (% of weight) Core proximate analysis (% by weight)

Organic content Ash Moisture Organic content Ash Moisture

Figure 5.7 Comparison between core proximate analysis against Rogers (2007) wireline log
estimation.

189
Chapter 5

Calvert et al. (2011)

100 100
Estimated proximate analysis (% by weight)

Estimated proximate analysis (% by weight)


90 Well 2 90 Well 3
80 80
70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Core proximate analysis (% by weight) Core proximate analysis (% by weight)

Organic content Ash Moisture Organic content Ash Moisture

Figure 5.8 Comparison between core proximate analysis against Calvert et al. (2011)
wireline log estimation.

190
Chapter 5

W method

100 100
Estimated proximate analysis (% by weight)

Estimated proximate analysis (% by weight)


90 Well 2 90 Well 3
80 80
70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100

Core proximate analysis (% by weight) Core proximate analysis (% by weight)


Organic content Ash Moisture Organic content Ash Moisture

Figure 5.9 Comparison between core proximate analysis against W method wireline log
estimation.

V method

100 100
Estimated proximate analysis ( v/v)

90 Well 2 90 Well 3
Estimated proximate analysis (v/v)

80 80
70 70
60 60
50 50
40 40
30 30
20 20
10 10
0 0
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Core proximate analysis (v/v) Core proximate analysis (v/v)

Organic content Ash Moisture Organic content Ash Moisture

Figure 5.10 Comparison between core proximate analysis against V method wireline log
estimation.

191
Chapter 5

GAS CONTENT CORRELATIONS

The examples given of the correlated desorbed core gas content are from the same

measure.

Calvert et al. (2011)

7 7

Well 2 Well 3
As analysed gas content (m3/tonne)

As analysed gas content (m3/tonne)


6 6

5 5

4 4

3 3

2 2

1 1

0 0
0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Reciprocal core density (g/cm3) Reciprocal core density (g/cm3)

Figure 5.11 Core reciprocal core density against as analysed gas content in the Lower
Juandah.
The plots represent the trendlines used to estimate gas content using the Calvert et al.
(2011).

192
Chapter 5

W method

7 7

Well 2 Well 3
As analysed gas content (m3/tonne)

As analysed gas content (m3/tonne)


6 6

5 5

4 4

3 3

2 2

1 1

0 0
0.3 0.4 0.5 0.6 0.7 0.8 0.9 0.3 0.4 0.5 0.6 0.7 0.8
Reciprocal core density (g/cm3) Reciprocal core density (g/cm3)

Figure 5.12 Core reciprocal density against as analysed gas content and pressure in the
Lower Juandah used in method W for gas estimation.
The plots represent the trendlines used to estimate gas content used in W method. In well 2
the measure was divided in upper and lower (red and blue points respectively).

V method

5 8

Well 2 Well 3
As analysed gas content (v/v)

7
As Analysed Gas Content (v/v)

4
6

5
3

2
3

2
1
1

0 0
0 10 20 30 40 50 60 70 80 90 100 0 10 20 30 40 50 60 70 80 90 100
Organic content (v/v) Organic Content (v/v)

Figure 5.13 Organic matter in v/v against as analysed gas content used in the Lower
Juandah in method V for gas estimation.
The plots represent the trendlines used to estimate gas content used in V method. In well 2
the measure was divided in upper and lower (red and blue points respectively).

193
Chapter 5

WP method

7 7

Well 2 Well 3

As analysed gas content (m3/tonne)


As analysed gas content (m3/tonne)

6 6

5 5

4 4

3 3

2 2

1 1

0 0
1.0 1.4 1.8 2.2 2.6 3.0 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Reciprocal core density (g/cm3) * Pressure (MPa) Reciprocal core density (g/cm3) * Pressure (MPa)

Figure 5.14 Reciprocal density and pressure against as analysed gas content in the Lower
Juandah used in method WP for gas estimation.
The plots represent the trendlines used to estimate gas content used in WP method. In well
2 the measure was divided in upper and lower (red and blue points respectively).

VP method

5 8

Well 2 Well 3
As analysed gas content (v/v)
As analysed gas content (v/v)

7
4
6

3 5

4
2
3

2
1
1

0 0
150 200 250 300 350 0 100 200 300 400 500
Organic content (v/v) * Pressure (MPa) Organic content (v/v) * Pressure (MPa)

Figure 5.15 Organic matter and pressure against as analysed gas content (v/v) in the Lower
Juandah used in method VP for gas estimation.
The plots represent the trendlines used to estimate gas content used in VP method.

194
Chapter 5

TOTAL GAS CONTENT ESTIMATION

Mavor and Nelson (1997)

10 10
9 Well 2 9 Well 3

Estimated gas content (m3/tonne)


Estimated gas content (m3/tonne)

8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content Core gas content as received content
(m3/tonne) (m3/tonne)
Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.16 Plots of gas content from core analysis against estimated total gas content
estimated using the Mavor and Nelson (1997).

Rogers (2007)

10 10
9 Well 2 9 Well 3
Estimated gas content (m3/tonne)
Estimated gas content (m3/tonne)

8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content Core gas content as received content
(m3/tonne) (m3/tonne)
Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.17 Plots of gas content from core analysis against estimated total gas content
estimated using the Rogers (2007).

195
Chapter 5

Calvert et al. (2011)

10 10

9 Well 2 9 Well 3

Estimated gas content (m3/tonne)


Estimated gas content (m3/tonne)

8 8

7 7

6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content Core gas content as received content
(m3/tonne) (m3/tonne)
Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.18 Plots of gas content from core analysis against estimated total gas content
estimated using the Calvert et al. (2011).

196
Chapter 5

W method

10 10

9 Well 2 9 Well 3
Estimated gas content (m3/tonne)

Estimated gas content (m3/tonne)


8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content Core gas content as received content (m3/tonne)
(m3/tonne)
Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.19 Plots of gas content from core analysis against estimated total gas content
estimated using W method.

V method

10 10
9 Well 2 9 Well 3
Estimated gas content (m3/tonne)
Estimated gas content (m3/tonne)

8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content Core gas content as received content (m3/tonne)
(m3/tonne)
Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.20 Plots of gas content from core analysis against estimated total gas content
estimated using V method.

197
Chapter 5

WP method

10 10
9 Well 2 9 Well 3
Estimated gas content (m3/tonne)

Estimated gas content (m3/tonne)


8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content Core gas content as received content
(m3/tonne) (m3/tonne)
Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.21 Plots of gas content from core analysis against estimated total gas content
estimated using WP method.

VP method

10 10

9 Well 2 9 Well 3
Estimated gas content (m3/tonne)
Estimated gas content (m3/tonne)

8 8

7 7

6 6

5 5

4 4

3 3

2 2

1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content Core gas content as received content (m3/tonne)
(m3/tonne)

Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.22 Plots of gas content from core analysis against estimated total gas content
estimated using VP method.

198
Chapter 5

ESTIMATED GAS CONTENT FROM COAL

Rogers (2007)

10 10
9 Well 2 9 Well 3
Estimated gas content (m3/tonne)

Estimated gas content (m3/tonne)


8 8
7 7
6 6
5 5
4 4
3 3
2 2

1 1

0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content Core gas content as received content
(m3/tonne) (m3/tonne)
Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.23 Plots of gas content from core analysis against estimated coal gas content
estimated using the Rogers (2007).

Calvert et al. (2011)

10 10
Well 2 Well 3
Estimated gas content (m3/tonne)

9 9
Estimated gas content (m3/tonne)

8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content (m3/tonne) Core gas content as received content (m3/tonne)

Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.24 Plots of gas content from core analysis against estimated coal gas content
estimated using the Calvert et al. (2011).

199
Chapter 5

W method

10 10
9
9 Well 2 Well 3
Estimated gas content (m3/tonne)

Estimated gas content (m3/tonne)


8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content (m3/tonne) Core gas content as received content (m3/tonne)

Upper Juandah Lower Juandah Taroon


Upper Juandah Lower Juandah Taroon

Figure 5.25 Plots of gas content from core analysis against estimated coal gas content
estimated using W method.

V method

10 10
Estimated gas content (m3/tonne)

9
9 Well 2 Well 3
Estimated gas content (m3/tonne)

8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content (m3/tonne) Core gas content as received content
(m3/tonne)
Upper Juandah Lower Juandah Taroon
Upper Juandah Lower Juandah Taroon

Figure 5.26 Plots of gas content from core analysis against estimated coal gas content
estimated using V method.

200
Chapter 5

WP method

10 10
9 Well 2 9 Well 3
Estimated gas content (m3/tonne)

Estimated gas content (m3/tonne)


8 8
7 7
6 6
5 5
4 4
3 3
2 2
1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content Core gas content as received content
(m3/tonne) (m3/tonne)
Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.27 Plots of gas content from core analysis against estimated coal gas content
estimated using WP method.

VP method

10 10

9 Well 2 9 Well 3
Estimated gas content (m3/tonne)
Estimated gas content (m3/tonne)

8 8

7 7

6 6

5 5

4 4

3 3

2 2

1 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
Core gas content as received content (m3/tonne) Core gas content as received content (m3/tonne)

Upper Juandah Lower Juandah Taroon Upper Juandah Lower Juandah Taroon

Figure 5.28 Plots of gas content from core analysis against estimated coal gas content
estimated using VP method.

201
Chapter 5

ROGER MAVOR CALVERT


METHOD METHOD METHOD

Figure 5.29 An example of wireline data Rogerss, Mavors and Calverts interpretation in a
multiple coal bed bearing zone in well 3.
Track 1: the depth in metres. Track 2: Sp, gamma ray and caliper. Track 3: contains density
log, neutron-porosity, sonic and Pe Track 4: displays the resistivity readings, shallow focused
electric and induction. Track 5: contains the core macroscopic lithological description. Track 6:
Core desorbed gas content. Track 7: Mullen/Rogers reservoir rock. Track 8: Mullen/Rogers
estimated gas content log. Track 9: Mavors reservoir rock. Track 10: Mavors estimated gas
content log. Track 11: Calverts reservoir rock. Track 12: Calverts estimated gas log
(APPENDIX I tools and logs acronyms).

202
Chapter 5

Figure 5.30 An example of wireline data interpretation in a multiple coal bed bearing zone
in well 3 using W and WP methods.
Track 1: the depth in metres. Track 2: SP, gamma ray and caliper. Track 3: contains density
3
log, neutron-porosity, sonic and Pe; Example log showing bulk density <1.2g/cm in coals
(circled) with good wellbore conditions. Track 4: displays the resistivity readings, shallow
focused electric and induction. Track 5: contains the core macroscopic lithological description.
Track 6: Core desorbed gas content. Track 7: reservoir rock. Track 8: Estimated gas content
log using the W method. Track 9: Estimated gas log using WP method (APPENDIX I tools
and logs acronyms).

203
Chapter 5

Figure 5.31 An example of wireline data interpretation in a multiple coal bed bearing zone
in well 3 using V and VP methods.
Track 1: the depth in metres. Track 2: SP, gamma ray and caliper. Track 3: contains density
3
log, neutron-porosity, sonic and Pe; Example log showing bulk density <1.2g/cm in coals
(circled) with good wellbore conditions. Track 4: displays the resistivity readings, shallow
focused electric and Induction. Track 5: contains the core macroscopic lithological description.
Track 6: Core desorbed gas content. Track 7: reservoir rock. Track 8: Estimated gas content
log using the V method. Track 9: Estimated gas log using VP method (APPENDIX I tools and
logs acronyms).

204
Chapter 5

Reservoir
Method Reservoir analysis Reservoir cut-off(s) Proximate Analysis estimation In situ gas estimation
rock(s)
Coal beds Separated in < (Ash+water) density Ash percentage of weight linear Negative relationship with
Mavor Carbonaceous reservoirs relationship with reciprocal ash and moisture percentage
Mudstones core density of weight
Coal beds No distinction Density <2.0g/cm3 Organic content, ash and Negative relationship with
Natural gamma ray log < 60 api moisture percentage of weight ash and moisture percentage
Neutron-porosity log > 50% independent linear relationship of weight
Rogers with core density
Sonic log > 80 s/ft
Shallow focused electric log >
50.m
Coal beds Separated in coal Density <1.8g/cm3 Ash and Ash+Water percentage As a function of reciprocal
Calvert measures of weight linear relationship core density
with reciprocal core density
Coal beds Separated in coal Density <2.0g/cm3 Organic content and ash As a function of organic
Coal laminae seams Shallow focused electric log > percentage of volume linear content volume and in situ
New (VP)
rich induction resistivity relationship with reciprocal pressure
mudstones Caliper core density

Table 5.11 Resume of the different coal seam gas well interpretation workflows.

205
Chapter 5

Gas in Place
Method Net reservoir
estimation
Total: 390.7m
(which is) 8% coal bed (detected) 100% coal bed
2% coal laminae 100% coal laminae 1,071,530
Mavor
1% carbonaceous mudstone 100% carbonaceous mudstone m3/Acre
6% mudstone 99% mudstone
82% Sandstone/Siltstone/Tuff 98% Sandstone/Siltstone/Tuff
Total: 19.9m
(which is) 92% coal bed (detected) 60% coal bed
Rogers 4% coal laminae 9% coal laminae 440,128 m3/Acre
1% carbonaceous mudstone 5% carbonaceous mudstone
2% mudstone 2% mudstone
Total: 31.6m
(which is) 82% coal bed (detected) 84% coal bed
10% coal laminae 34% coal laminae
Calvert 662,592 m3/Acre
4% carbonaceous mudstone 22% carbonaceous mudstone
2% mudstone 3% mudstone
2% Sandstone/Siltstone/Tuff <1% Sandstone/Siltstone/Tuff
Total: 40.7m
(which is) 74% coal (detected) 97% coal bed
16% coal laminae 72% coal laminae
New (VP) 563,857 m3/Acre
5% carbonaceous mudstone 36% carbonaceous mudstone
2% mudstone 4% mudstone
3% Sandstone/Siltstone/Tuff <1% Sandstone/Siltstone/Tuff
Table 5.12 Rogerss, Mavors, Calverts and the new workflow results from Well 2.

Gas in Place
Method Net reservoir
estimation
Total: 65.7m
(which is) 38% coal bed (detected) 100% coal bed
8% coal laminae 100% coal laminae 820,306
Mavor
8% carbonaceous mudstone 99% carbonaceous mudstone m3/Acre
20% mudstone 29% mudstone
26% Sandstone/Siltstone/Tuff 8% Sandstone/Siltstone/Tuff
Total: 10.3m
(which is) 92% coal bed (detected) 37% coal bed
226,665
Rogers 2% coal laminae 3% coal laminae
m3/Acre
2% carbonaceous mudstone 4% carbonaceous mudstone
4% Sandstone/Siltstone/Tuff <1% Sandstone/Siltstone/Tuff
Total: 26.8m
(which is) 77% coal bed (detected) 82% coal bed
6% coal laminae 33% coal laminae 627,923
Calvert
6% carbonaceous mudstone 31% carbonaceous mudstone m3/Acre
5% mudstone 3% mudstone
6% Sandstone/Siltstone/Tuff 1% Sandstone/Siltstone/Tuff
Total: 31.3m
(which is) 79% coal (detected) 97% coal bed
12% coal laminae 74% coal laminae 650,496
New (VP)
5% carbonaceous mudstone 30% carbonaceous mudstone m3/Acre
3% mudstone 2% mudstone
1% Sandstone/Siltstone/Tuff <1% Sandstone/Siltstone/Tuff

Table 5.13 Rogerss, Mavors, Calverts and the new workflow results from in well 3.

206
Chapter 5

Figure 5.32 Basic two step workflow methodology to estimate net coal (as debated in
chapter 4) and in situ gas content with input of wireline logs and core and core analysis data.
After this methodology is applied to an initial core well it can be correlated to production
wells without necessary needing the core input.

207
Chapter 5

5.5. Conclusions

This chapter focuses on how the interpretation of wireline log readings can be used

in the estimation of gas content in coal seam gas wells.

For this, core analysis was investigated as the best approach to relate gas content

with wireline log responses. It was observed that coal beds in the Walloon Sub-group

have a variable maceral composition, influencing the density of the organic matter and

the estimation of coal quality. Desorbed gas analysis was found to be a fundamental

measurement for in situ gas calculation purposes, while saturation estimation of the coal

seams is important, although only qualitatively.

Previous methodologies to estimate gas content were reviewed and discussed. The

estimation of gas in place results using different methodologies is related not only with

distinct interpretation of the well data but also to how the core to log correlation is

established.

New methodologies were created and tested, in which the VP method is shown to

achieve the best results amongst all the methodologies considered. This new workflow,

following the conclusions in chapter 4, includes an estimation of in situ gas content as a

function of organic matter volume together with in situ pressure. This new workflow is

shown to be more accurate when estimating coal seam thickness and has an associated

lower mean absolute percentage error when calculating reservoir rock quality and in

situ producible gas content compared to other methods.

208
Appendices

CHAPTER 6 Conclusions
This work investigates the nature of coal, and specifically the Surat Basin Walloon

Sub-group coal beds, by analysing well data comprising core analysis, wireline logs and

DST test results. The overall objective of the study is to determine the best strategy for

the petrophysical analysis of coal and coal seam gas by raising key scientific questions,.

The current understanding of coal beds is presented and based on this knowledge a

novel way is demonstrated to evaluate wireline logs. This new workflow can

significantly improve the estimation of gas content and reserves in CSG from

petrophysical interpretation.

6.1. Conclusions to initial questions posed

Question 1: What is the nature of the coal beds found in the Walloon Sub-group
regarding their composition, thickness and fracture system?

As a step to enable the interpretation of the physical properties such as coal thickness

and gas content chapter 3 identifies the main features found in core samples that in later

chapters are estimated through wireline log interpretation. Chapter 3 considers the

traditional macroscopic coal lithotypes terminology and discusses possible links to

critical differences in coal composition, inner structure and gas flow behaviour that

ultimately affect not only core analysis results and downhole measurements but also the

reservoir performance.

The Walloon coal seams are shown to be thinly bedded predominately dull with

more or less bright bands. Therefore most of the beds were classified as Dull minor

bright or Dull banded, and these two lithotypes together dominate all the sub-groups

individual coal seams.

209
Appendices

Composite sample reflectance analysis indicated that these are sub-bituminous

to high-volatile bituminous rank with vitrinite rich coals, predominately telocollinite

and desmocollinite. A literature search revealed that suberinite is also a common

maceral in these coals, but this are relatively rare, as many of the worldwide coal seam

gas reservoirs are usually dominated by Vitrinite, Inertinite or by both.

Analysis of the core descriptions of the 13 cored wells shows that both face and

butt cleat distribution is a function of coal type, coal quality and depth. Mineral content

clearly constrains cleat develop, nevertheless coal beds are variable when it comes to

cleat spacing, having no macroscopically visible cleats up to a cleat spacing of less than

1mm. Most of the cleats are partially filled with carbonaceous or clay minerals,

therefore restricting but not necessarily stopping gas and water flow.

Desorbed gas from the Walloon Sub-group coal is predominately methane. Gas

content increases from the shallowest coal beds to the Lower Juandah measure due to

the increased ability of the organic content to adsorb gas with each pressure increment.

The gas content within the Lower Juandah coal seams steadily decreases with depth.

Estimated saturation has been shown to be highly variable and this can be in part

explained by the limited number of samples in a heterogenous rock type.

Question 2: How are the different coal lithotypes reflected in petrophysical


properties?

This research demonstrates that the identification of independent lithotypes within a

coal seam can be associated with specific petrophysical characteristics. The bright

bands in coal seams suggest fixed carbon increase in the samples while volatiles only

increase with brightness until reaching Dull & Bright description. Moreover, duller

lithotypes are richer in ash (usually clay minerals) and moisture. As organic content

210
Appendices

increases with brightness so does the vitrinite; interestingly the Dull banded facies show

a slight decrease of liptinite in what would otherwise be a trend identical to vitrinite.

One of the features that can be associated with each lithotype is cleat

development. Dull and Stony coal beds have more samples without cleats while other

lithotypes are shown to have some degree of fracture development in most cases (80%

beds). Probably the most interesting cleat to lithotype association is the trend that coal

with more bands has more frequent cleats; in contrast the more homogenous the bed is

then the more penetrative is the cleat. In the Walloon Sub-group coal beds cleat

frequency increases with bed thickness, but is dependent on the lithotype,.

This study was unable to establish a direct relationship between coal lithotypes

and gas content except that gas content increases as the organic content increases

against ash content; in other words, as coal quality increases under identical

pressure/temperature conditions in each lithotype so does the ability to adsorb more

methane. The brightest coal beds, due to their higher organic content, may be indicative

of a higher gas content compared to other bordering coals.

Wireline log responses, especially unfiltered ones, are influenced by the nature

of the coal lithotypes, namely coal quality and thickness, even though it seems difficult

to distinguish them from log analysis in a systematic way. Additionally, highly

fractured banded coal beds are shown to have lower resistivity responses compared to

others, although permeability estimation seems impractical.

No specific characteristics in coal were found that clearly relate to the

permeability estimated from DSTs. Difficulties in establishing any relationship could be

associated with the very heterogeneous nature of lacustrine Walloon Sub-group, limited

lateral continuity of the bed(s) and/or of the cleats found in the DST test zones. Gas

saturation can also be a factor leading to difficulty in estimation of permeability due to

211
Appendices

its high variability. Nonetheless, more permeable coal zones generally are characterized

by being thinly banded coal, with a greater number of differentiated beds than low

permeability coals, and being more mature with a higher number of cleats.

Question 3: What is the accuracy of estimating net coal thickness in thinly


bedded coal seams by interpreting density logs and other standard wireline
measurements downhole?

The interpretation of logs over thinly bedded coal seams requires precision not only

from the tools themselves but also from the interpreter, which if neglected can result in

a flawed estimation of net coal, computed permeability and, consequently, the gas-in-

place reserves calculation. Chapter 3 focussed on the nature of the Walloon Sub-group

lithotypes coal bed, while chapter 4 considered wireline measurements and how net

coal can be calculated from wireline log analysis.

Interpretation of wireline responses in 5 wells (3 cored and 2 production)

compared against core analysis indicates that the calculation has a significant margin of

error associated with the limited thickness and common borehole washouts found in

coal seam gas exploration and development wells. Sharp borehole enlargements affect

more or less, all nuclear and sonic measurements. Even in good borehole conditions the

wireline nuclear measurements are not as sensitive to sharp changes in coal facies when

these are thinly bedded and exhibit dramatic differences in organic content.

The usage of a cut-off following the definition of coal as an organic rich

lithology, with low density, low natural gamma ray, high apparent neutron-porosity and

high travel time readings, leads to highly underestimated reservoir thickness. On the

other hand, relaxing the cut-offs and interpreting sections with close proximity of thin

coal bed with highly inorganic tuffaceous layers, mudstones and siltstones leads to the

overestimation of total coal net thickness. To demonstrate this, cross-plots Figure 4.5

212
Appendices

(page 134), Figure 4.11 (p145), Figure 4.12 (p145), Figure 4.13 and Figure 4.14

(p149) were created comparing what was classified as reservoir rock (coal and coal

laminae rich mudstones) against what is estimated using each standard wireline

measurement analysed.

Question 4: Can resistivity measurements be used to improve the accuracy of


net coal estimation both in cored and production wells?

Downhole resistivity measurements over the same coal seam can provide very

different results depending on the tool configuration. The signal detected by the Array

Induction tool receiver is associated with the conductive fluid filled paths found in the

formation, consequently giving more weight to any conductive formations within its

investigation range.

The response of the high-resolution shallow focused electric log in the 5 wells

analysed is mainly a function of composition and geometrical configuration of coal bed

bodies and therefore the study has indicated it is advantageous to look to resistivity

measurements to increase the definition of what can be considered as a potential

methane producer when drilling for CSG. The contrast between these two

measurements is beneficial to formation evaluation of coal seam stratigraphy and coal

bed thickness calculation in cases of thicker coal seams with non-coal interbeds or

isolated coal beds.

Adding shallow focused electric and the induction resistivity log measurements

to nuclear measurements, preferably the high-resolution density log, can help

substantially with the estimation of net coal by improving the definition not only of

organic-rich coal beds but also stony coal and carbonaceous mudstones with coal wisps

(Table 4.3 in page 149).

213
Appendices

The level of accuracy improvement using this cut-off in net coal thickness

estimation depends on borehole conditions, coal seam thickness and the number of non-

coal interbeds within the coal seam. It is shown to be useful in cases where zones of the

well have suffered washouts and other logs are less reliable. It has also proved its worth

in production wells with borehole sizes larger than cored wells that make up the large

number of wells in commercially successful CSG fields.

Question 5: What is the level of error in the estimation of in situ gas content
using published methodologies?

Three workflows were tested to understand the level of error in gas content

estimation in CSG through the analysis of well data. Due to the particularities of the

Walloon Sub-group coal and the observed gas content in core samples the workflows

tested were Mavor and Nelson (1997), Rogers (2007) and Calvert et al. (2011).

Following Mavor and Nelson (1997) the estimate of gas content provides

results of gas-in-place regardless of whether the gas is in coal or other rock types.

In Rogers (2007) even though the total gas content estimation is an

improvement compared to Mavor and Nelson (1997) this methodology applied a strict

cut-off to identify coal bed that results in significant underestimation of the resource.

The Calvert et al. (2011) methodology benefits in having a similar approach to

gas content estimation to Mavor and Nelson (1997) but compared to Rogers (2007)

defines a cut-off to select coal that is much less strict and more accurate.

The Mean Absolute percentage error in total gas content estimation was as high

as 26.6% in Mavor and Nelson (1997) and as low as 21.6% in Rogers (2007).

214
Appendices

Question 6: Can a methodology be created to optimize this estimation?

Since it was found that the major factor influencing gas content is coal quality,

Chapter 5 considered how organic content density changes from coal seam to coal

seam. The maceral content of the Walloon Sub-group coal beds influences the organic

content density. The extrapolation of reciprocal density to 100% organic content

indicated that the maceral composition of the organic material produces a density of

1.32g/cm3 and 1.15g/cm3. This observation led to the question as to whether estimating

organic content volume first, before estimating gas content in coal, could improve the

estimation of gas-in-place in CSG. The identification of distinctions between Walloon

Sub-group coal seams with regards to gas saturation and the influence that pressure has

in coal gas adsorption led to testing if this last parameter could help constrain the error

of gas content estimation from wireline logs.

Out of four new methodologies tested it was possible to create one, designated

as VP, that show the best results, even when compared to published methods, which

provides a Mean standard percentage error in total gas content estimation of 13.2% in

well 2 and 16.2% in well 3. This was the lowest error when compared to other

methodologies (Table 5.7 and Table 5.8 in page 186). This methodology uses the coal

thickness estimation methodology advised in CHAPTER 4 together with estimation of

organic content volume and in situ pressure outlined in CHAPTER 5 to estimate gas-in-

place in CSG, providing what is concluded to be the optimum strategy of coal seam gas

wells petrophysical analysis.

215
Appendices

6.2. General suggestions for further work

Additional questions surfaced during the interpretation of this dataset that could only

be addressed if further analysis could be supported by new data and a longer time

frame.

As coal quality is one of the main factors attending gas storage capacity and gas

content in coal, its estimation could benefit from more detailed information regarding

the nature of maceral content, especially maceral density. Measurements of different

maceral concentrates have been tested before but not for macerals from the Surat Basin

(Choi et al., 1987; Dyrkacz and Bloomquist, 1992; Dyrkacz and Horwitz, 1982;

Stankiewicz et al., 1994; Taulbee et al., 1989). A statistical evaluation of the different

maceral density as a function of depth and maturation could possibly help the definition

of a more precise estimation of organic content volume (i.e., coal quality) during CSG

formation evaluation.

Another parameter that could be investigated more thoroughly is the gas storage

capacity of the Walloon Sub-group. The Langmuir isotherm analysis dataset used in

this work is composed of the results of tests of composite samples of the same seams

from 11 wells. As explained in this study,, the common practice in CSG exploration to

test composite samples diminishes confidence in the representability of this test. It

would be interesting to evaluate the gas storage capacity while testing core samples as

they were recovered after the in situ gas was desorbed from them.

Following the conclusions of chapters 4 and 5, it seems advantageous to apply the

techniques developed and extend them into multiple wells and to incorporate seismic

data thus enabling an evaluation of how the new workflow would affect the estimated

CSG reserves. Additionally it would be interesting to compare the conclusions from this

216
Appendices

work, which focusses on the Walloon Sub-group coal seams in the Surat Basin, against

analysis of other Basins where the nature of the coal beds is different.

217
Appendices

APPENDIX A Seam and net coal thickness

The following table provides the total thickness and the net coal in each coal seam,

through the analysis of 11 wells core description calculated in 1cm intervals provided

by the operator.

Wells
Coal seam Thickness 1 2 3 4 5 6 7 10 11 13 14

Total 10.4 27.1 16.7 23.8 14.9 41.4 6.6 - 22.3 1.3 -
Kogan
Coal - 1.1 0.3 1.8 1.1 2.4 1.4 - 1.7 1.0 -
Upper Juandah

Macalister Total 16.1 20.5 11.2 11.9 22.9 22.9 29.5 7.9 28.1 30.4 -
Upper Coal - 2.5 2.4 2.6 3.4 3.9 3.1 0.5 1.7 2.9 -

Macalister Total 16.3 17.2 13.9 6.8 13.1 14.5 44.0 6.8 25.7 17.8 23.2
Lower Coal - 0.0 1.7 3.7 0.5 3.1 4.9 0.4 6.9 4.5 2.6
Total 41.2 16.5 25.3 24.4 15.5 28.9 12.9 27.4 47.5 21.1 12.2
Nangram
Coal 0.1 0.0 0.9 1.6 0.6 2.8 0.5 2.2 0.2 1.8 1.5

Total 23.4 51.0 24.1 27.6 13.5 13.9 60.0 51.4 33.4 42.2 30.1
Wambo
Lower Juandah

Coal 2.0 6.8 1.9 0.7 1.2 0.9 4.1 2.4 0.8 2.0 2.6
Total 21.6 43.3 34.0 29.7 17.8 25.3 56.9 75.8 29.1 50.2 33.0
Iona
Coal 0.8 3.4 4.2 0.7 2.5 1.3 4.4 6.4 1.8 2.3 2.2
Total 50.7 48.3 57.6 50.0 33.4 25.6 40.0 75.0 42.7 75.9 56.9
Argyle
Coal 4.0 6.5 3.9 2.6 2.5 4.7 4.6 4.6 5.3 6.2 8.3

Total 32.2 46.8 30.2 55.0 64.1 47.1 49.4 71.2 77.5 44.2 28.7
Tangalooma Ss.
Coal 0.4 1.1 0.2 1.5 2.9 1.1 0.4 0.6 0.5 1.8 0.5

Total 11.2 27.0 23.3 12.7 30.0 26.5 23.7 - 36.5 29.6 35.8
Auburn
Coal 1.2 0.5 2.0 0.8 3.9 2.9 0.4 - 1.6 3.0 1.9
Taroom

Total 44.5 51.1 27.6 32.7 38.7 15.4 88.4 39.8 13.2 73.2 32.7
Bulwer
Coal 2.3 1.1 2.2 0.8 7.6 - 4.0 2.8 3.5 2.3 1.7

Total 44.1 64.6 47.4 36.3 25.5 56.8 46.9 33.8 54.6 68.8 20.7
Condamine
Coal 2.2 8.0 5.7 3.4 6.4 5.5 3.3 3.1 4.4 7.6 6.6

Total Thickness 312 413 311 311 289 318 458 389 410 455 273

Total Coal Thickness 13 31 25 25 32 28 31 23 28 35 28

218
Appendices

APPENDIX B proximate and desorption

analysis
The following tables display the proximate analysis, desorption analysis gas content

(as analysed) and sorption time minimum, maximum, mean and standard deviation

estimated from core samples retrieved from wells 1 to 8 and 10 to 14.

Volatile matter (wt.%) Volatile matter (wt.%) (d.a.f.)


Count Min Max Mean St.dev. Min Max Mean St.dev.
Kogan 18 4.8 38.1 22.3 9.8 47.8 98.9 59.6 15.9
Macalister Upper 55 4.9 45 28.8 11.5 45.1 97.5 57.8 13.1
Macalister Lower 48 5.7 43.7 29.1 10.5 48.6 99.6 57.9 11.3
Nangram 21 2 42.5 31.5 11.7 48.9 95.2 58.7 11.7
Wambo 43 1.3 48.4 28.8 11.7 1.7 98.1 58.5 14.9
Iona 47 5.7 47.5 29.7 10.8 45.2 98.3 59.1 9.2
Argyle 88 4.6 48.3 34.1 10.5 49.7 94.2 57.4 7.2
Tangalooma Ss. 9 5.5 38.9 29.3 11.4 38.8 100.0 58.6 16.7
Auburn 28 3.2 45.1 27.5 10.7 49.8 97.0 59.2 8.8
Bulwer 48 2.8 48.6 30.8 11.4 47.7 98.1 60.4 12.2
Condamine 84 3.4 45.8 34.1 9.0 47.8 98.5 55.7 9.2

Fixed Carbon (wt.%) Fixed Carbon (wt.%) (d.a.f.)


Count Min Max Mean St.dev. Min Max Mean St.dev.
Kogan 18 0.1 40.9 19.3 12.3 1.1 52.2 40.4 15.9
Macalister Upper 55 0.2 40.8 25.0 12.3 2.5 54.9 42.2 13.1
Macalister Lower 48 0.1 40.3 24.1 11.6 0.4 51.4 42.1 11.3
Nangram 21 0.1 42.4 26.0 12.2 4.8 51.1 41.3 11.7
Wambo 43 0.1 74.2 24.7 13.9 1.9 98.3 41.5 14.9
Iona 47 0.1 43 23.0 11.0 1.7 54.8 40.9 9.2
Argyle 88 0.5 44.2 27.2 10.5 5.8 50.3 42.6 7.2
Tangalooma Ss. 9 13.3 42.9 29.2 10.1 0.0 61.2 41.4 16.7
Auburn 28 0.1 44.3 21.3 11.1 3.0 50.2 40.8 8.8
Bulwer 48 0.1 44.3 24.0 11.6 1.9 52.3 39.6 12.2
Condamine 84 0.1 45.3 29.6 10.3 1.5 52.2 44.3 9.2

Ash Content (wt.%) Relative Density (g/cm3)


Count Min Max Mean St.dev. Min Max Mean St.dev.
Kogan 18 11.6 88.3 51.3 22.6 1.35 2.56 1.8 0.3
Macalister Upper 55 11.3 90 39.7 23.8 1.31 2.46 1.7 0.3
Macalister Lower 48 9.4 88.4 40.1 22.2 1.3 3.06 1.7 0.3
Nangram 21 8.3 94.9 36.8 24.1 1.33 2.36 1.6 0.3
Wambo 43 3.7 91.5 40.2 23.5 1.22 2.56 1.6 0.3
Iona 47 3.8 89.8 41.4 22.5 1.27 2.45 1.6 0.3
Argyle 88 4.4 93.4 33.3 21.4 1.27 2.60 1.5 0.3
Tangalooma Ss. 9 4.6 92.1 37.8 27.1 1.26 2.54 1.6 0.4
Auburn 28 5.2 95.8 46.4 22.3 1.27 2.66 1.7 0.3
Bulwer 48 8.3 94.4 40.0 23.0 1.29 2.51 1.6 0.3
Condamine 84 6.1 92.5 31.2 19.4 1.28 2.54 1.5 0.3

219
Appendices

Gas content (m3/tonne)


Count Min Max Mean St.dev.
Kogan 18 0.17 5.68 1.8 1.3
Macalister Upper 53 0.25 11.86 3.0 1.9
Macalister Lower 47 0.39 7.09 3.5 1.6
Nangram 21 0.39 7.32 3.6 2.4
Wambo 42 0.34 15.23 4.0 2.6
Iona 46 0.64 9.15 3.6 1.8
Argyle 86 0.55 9.86 4.4 2.1
Tangalooma Ss. 8 1.95 6.77 4.7 1.6
Auburn 28 0.34 6.50 3.5 1.9
Bulwer 48 0.16 10.08 3.7 2.4
Condamine 84 0.15 10.99 4.1 1.9

Sorption Time (days)


Count Min Max Mean St.dev.
Kogan 18 0.41 108.83 12.3 25.0
Macalister Upper 53 0.26 78.60 16.3 20.5
Macalister Lower 47 0.34 27.93 5.7 6.6
Nangram 21 1.16 120.73 13.1 27.5
Wambo 42 1.17 82.73 12.2 15.2
Iona 46 0.23 33.52 8.0 9.0
Argyle 86 0.15 34.46 7.8 8.2
Tangalooma Ss. 8 0.82 16.31 4.2 5.2
Auburn 28 0.16 23.23 5.9 6.3
Bulwer 48 0.18 43.84 7.8 9.6
Condamine 84 0.22 40.25 10.4 9.0

220
Appendices

APPENDIX C wireline density to organic

content x-plots
The following x-plots demonstrate that best wireline density to organic content
match is found in the unfiltered high-resolution density log (HDEN) against the filtered
version and the compensated log (DEN).

100

90 NO Filters r = 0.91
Organic Content (% by weight)

80

70

60

50

40

30

20

10

0
1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6
High-resolution wireline density (g/cm3)

100
90 WITH Filters r = 0.90
Organic Content (% by weight)

80
70
60
50
40
30
20
10
0
1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6
High-resolution wireline density (g/cm3)

221
Appendices

100

90 WITH Filters r = 0.88


80

Organic Content (% by weight) 70

60

50

40

30

20

10

0
1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6
Compensated wireline density (g/cm3)

222
Appendices

APPENDIX D wireline responses


The following tables show the correlation between core samples with the mean

wireline log responses (gamma ray, density, Pe, neutron-porosity, sonic and resistivity)

provided by Weatherford UK from samples from 10 well (1,2,3,6,8,10,11,12,13 and 14

well) from the Walloon Sub-group coal seams.

223
Appendices

Well 1 - Samples 1 2 3 4 5 6 7 8 9 10 11 12 13
Gamma ray (API) 53 60 75 56 74 64 57 60 67 43 82 44 38
Compensated Density (g/cm3) 1.83 1.51 1.75 1.44 1.56 1.36 1.52 1.59 1.60 1.42 1.89 1.55 1.51
Mean wireline responses

Vetar Density (g/cm3) 1.80 1.46 1.72 1.44 1.56 1.33 1.53 1.46 1.57 1.43 1.88 1.48 1.47
NEAR Density (g/cm3) 1.84 1.49 1.75 1.53 1.49 1.35 1.49 1.56 1.61 1.49 1.86 1.56 1.56
FAR Density (g/cm3) 1.83 1.49 1.75 1.46 1.53 1.35 1.50 1.60 1.60 1.43 1.87 1.56 1.51
Neutron Porosity (percent) 55 63 67 84 75 83 82 80 77 86 63 79 75
NEAR Neutron Raw (snu) 6292 5248 5890 4814 5051 4311 5186 5180 5327 4933 7093 5929 5254
FAR Neutron Raw (snu) 293 218 250 169 176 151 178 189 200 165 298 227 207
PE (barns/electron) 1.83 1.38 1.49 1.29 1.54 1.58 1.46 1.25 1.47 1.28 1.53 1.24 0.97
FEFE (ohm.m) 39 66 40 171 14 14 152 33 58 408 27 134 1117
Medium Induction (ohm.m) 5 8 13 14 5 6 9 9 10 14 7 10 12
Deep Induction (ohm.m) 5 8 12 13 6 5 8 8 9 12 7 9 11
Sample Length (m) 0.47 0.41 0.82 0.82 0.87 0.39 0.72 0.42 0.82 1.00 0.35 0.50 0.50
Depth (m) 338.15 341.93 349.70 354.76 366.60 368.58 383.60 390.48 402.73 406.00 411.99 421.13 428.06
SEAM Wambo Iona Argyle

224
Appendices

Well 2 - Samples 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Gamma ray (API) 69 48 71 65 49 42 70 40 47 35 53 67 38 46
Compensated Density (g/cm3) 1.76 1.53 1.89 1.72 1.69 1.62 1.79 1.47 1.61 1.52 1.57 1.58 1.45 1.64
Mean wireline responses

Vetar Density (g/cm3) 1.75 1.53 1.88 1.73 1.69 1.59 1.76 1.41 1.60 1.46 1.49 1.49 1.43 1.61
NEAR Density (g/cm3) 1.39 1.39 1.69 1.53 1.54 1.44 1.63 1.28 1.45 1.31 1.33 1.42 1.33 1.48
FAR Density (g/cm3) 1.63 1.49 1.82 1.65 1.64 1.59 1.74 1.42 1.58 1.48 1.52 1.56 1.42 1.62
(unfiltered)

Neutron Porosity (percent) 77 81 66 71 84 65 63 82 78 77 72 70 89 76


NEAR Neutron Raw (snu) 5436 4952 6716 5727 5679 5823 6417 5306 5526 5430 5439 5360 5033 5323
FAR Neutron Raw (snu) 198 175 271 218 198 255 278 198 201 200 228 229 166 204
PE (barns/electron) 1.83 1.68 1.97 1.97 1.90 1.69 1.83 1.58 1.70 1.35 1.61 1.46 1.67 1.91
FEFE (ohm.m) 16 30 6 27 19 56 17 93 33 108 30 31 41 24
Medium Induction (ohm.m) 6 9 2 4 6 8 5 8 10 13 10 9 11 4
Deep Induction (ohm.m) 6 9 2 4 6 6 5 7 9 12 9 9 8 3
Sample Length (m) 0.720 0.650 0.750 0.860 0.550 0.690 0.720 0.550 0.300 0.430 0.370 0.530 0.700 0.410
Depth (m) 169.03 169.74 190.36 191.63 192.28 236.18 243.85 246.28 252.15 258.82 261.89 267.86 271.10 272.21
SEAM Kogan Macalister Upper Wambo

Well 2 - Samples 15 16 17 18 19 20 21 22 23 24 25 26
Gamma ray (API) 53 43 38 50 44 43 74 55 69 49 41 53
Compensated Density (g/cm3) 1.48 1.39 1.46 1.53 1.46 1.39 1.72 1.53 1.68 1.77 1.53 1.52
Mean wireline responses

Vetar Density (g/cm3) 1.46 1.37 1.36 1.47 1.41 1.39 1.70 1.49 1.67 1.71 1.49 1.41
NEAR Density (g/cm3) 1.36 1.21 1.25 1.35 1.24 1.20 1.55 1.31 1.49 1.39 1.34 1.28
FAR Density (g/cm3) 1.45 1.34 1.43 1.49 1.41 1.33 1.67 1.47 1.63 1.69 1.49 1.49
(unfiltered)

Neutron Porosity (percent) 88 90 83 75 80 86 76 85 73 57 85 71


NEAR Neutron Raw (snu) 5148 4930 5309 5454 5201 5015 5781 5270 5649 6386 5420 5430
FAR Neutron Raw (snu) 178 170 207 216 199 189 218 195 220 311 207 230
PE (barns/electron) 1.61 1.57 1.24 1.54 1.54 1.54 1.75 1.44 2.01 1.78 1.69 1.56
FEFE (ohm.m) 95 116 250 37 102 258 17 109 34 27 184 39
Medium Induction (ohm.m) 9 15 8 7 9 12 7 8 7 6 6 6
Deep Induction (ohm.m) 9 12 7 8 9 10 7 7 6 5 5 6
Sample Length (m) 0.800 0.700 0.430 0.740 0.550 0.630 0.790 0.450 0.540 0.340 0.490 0.360
Depth (m) 290.05 293.63 304.09 349.84 353.38 364.68 367.91 371.04 374.99 424.11 452.57 475.51
SEAM Iona Argyle Auburn Bulwer

225
Appendices

Well 2 Samples 27 28 29 30 31 32 33 34 35 36 37 38
Gamma ray (API) 39 23 55 58 49 36 42 55 50 56 65 38
Mean wireline responses (unfiltered)

Compensated Density (g/cm3) 1.39 1.29 1.41 1.59 1.48 1.36 1.59 1.45 1.73 1.47 1.44 1.44
Vetar Density (g/cm3) 1.38 1.27 1.32 1.57 1.46 1.33 1.48 1.37 1.68 1.44 1.45 1.42
NEAR Density (g/cm3) 1.28 1.14 1.24 1.45 1.34 1.22 1.29 1.25 1.56 1.34 1.30 1.11
FAR Density (g/cm3) 1.35 1.25 1.38 1.54 1.43 1.33 1.52 1.39 1.69 1.43 1.39 1.33
Neutron Porosity (percent) 90 93 99 80 88 90 78 84 71 89 90 78
NEAR Neutron Raw (snu) 5124 4904 5042 5705 5301 5092 5383 5223 6625 5487 5214 5153
FAR Neutron Raw (snu) 168 159 166 208 180 172 211 201 267 197 171 185
PE (barns/electron) 1.61 1.49 1.47 1.60 1.52 1.52 1.47 1.61 1.76 1.52 1.57 1.55
FEFE (ohm.m) 178 311 253 627 172 2259 3793 170 90 1102 661 600
Medium Induction (ohm.m) 27 26 10 12 15 21 16 10 6 12 15 10
Deep Induction (ohm.m) 23 16 9 11 12 14 15 8 6 11 12 8
Shallow Laterolog (ohm.m) 151 530 57 28 41 520 50 63 17 113 62 21
Deep Laterolog (ohm.m) 130 480 65 25 35 432 47 62 15 94 46 16
Sample Length (m) 0.760 0.600 0.320 0.610 0.760 0.790 0.560 0.350 0.870 0.620 0.760 0.730
Depth (m) 503.07 503.77 506.84 521.24 521.92 522.70 528.98 536.93 548.53 554.71 555.48 556.23
SEAM Condamine

226
Appendices

Well 3 - Samples 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Gamma ray (API) 33 47 51 71 76 52 58 80 136 82 115 100 57 97
Compensated Density (g/cm3) 1.42 1.43 1.53 1.96 1.71 1.55 2.37 1.60 2.23 1.63 1.88 1.60 1.49 1.81
Mean wireline responses

Vetar Density (g/cm3) 1.42 1.41 1.53 1.95 1.69 1.53 2.37 1.57 2.25 1.60 1.88 1.59 1.46 1.80
NEAR Density (g/cm3) 1.50 1.49 1.60 1.94 1.75 1.57 2.24 1.45 2.04 1.55 1.66 1.60 1.43 1.73
FAR Density (g/cm3) 1.45 1.46 1.55 1.95 1.74 1.56 2.32 1.56 2.16 1.61 1.81 1.60 1.47 1.78
(unfiltered)

Neutron Porosity (percent) 94 96 79 60 68 85 30 77 39 76 64 81 83 65


NEAR Neutron Raw (snu) 5063 5037 5424 7355 6003 5223 10914 5179 8632 5337 5922 5119 4888 6223
FAR Neutron Raw (snu) 160 160 188 319 250 179 734 195 506 207 248 184 172 281
PE (barns/electron) 1.2 1.0 1.1 1.6 1.4 1.3 2 1.4 2 1.4 1.7 1.4 1.3 1.63
FEFE (ohm.m) 562 490 168 33 21 39 4 57 8 36 12 24 188 18
Medium Induction (ohm.m) 31 24 17 8 6 6 3 10 6 9 7 9 12 8
Deep Induction (ohm.m) 25 19 13 8 6 6 4 9 6 8 7 8 12 8
Sample Length (m) 0.69 0.75 0.48 0.79 0.78 0.8 0.74 0.7 0.79 0.59 0.83 0.8 0.81 0.63
Depth (m) 359.07 359.79 360.40 361.04 369.84 370.92 397.43 409.28 411.23 414.67 415.60 434.91 435.72 437.44
SEAM Macalister Upper Macalister Lower Nangram Wambo Iona

Well 3 - Samples 15 16 17 18 19 20 21 22 23 24 25 26 27 28
Gamma ray (API) 61 87 49 66 80 131 69 98 57 84 86 140 70 55
Compensated Density (g/cm3) 1.6 1.7 1.4 1.5 1.6 2.3 1.6 1.8 1.5 1.9 1.7 2.0 1.7 1.5
Mean wireline responses

Vetar Density (g/cm3) 1.56 1.62 1.30 1.40 1.65 2.26 1.56 1.73 1.47 1.91 1.59 1.93 1.64 1.47
NEAR Density (g/cm3) 1.56 1.63 1.39 1.43 1.45 2.16 1.56 1.71 1.48 1.88 1.65 1.91 1.70 1.59
FAR Density (g/cm3) 1.61 1.71 1.44 1.48 1.58 2.22 1.63 1.79 1.51 1.90 1.69 1.96 1.68 1.55
(unfiltered)

Neutron Porosity (percent) 71 66 69 77 74 52 82 62 79 64 71 66 80 91


NEAR Neutron Raw (snu) 5127 5579 4917 5060 5721 7886 5342 5911 5050 6830 5449 6320 5922 5303
FAR Neutron Raw (snu) 210 234 209 208 228 379 198 268 196 277 227 263 215 188
PE (barns/electron) 1.21 1.34 0.92 1.14 1.44 2.12 1.08 1.49 1.31 1.63 1.31 1.58 1.26 1.13
FEFE (ohm.m) 183 30 704 92 34 5 246 35 376 58 46 14 120 245
Medium Induction (ohm.m) 9 8 9 7 7 3 6 5 12 8 8 7 9 8
Deep Induction (ohm.m) 9 8 8 7 7 3 6 5 11 8 8 7 9 8
Sample Length (m) 0.38 0.45 0.4 0.35 0.8 0.8 0.41 0.37 0.79 0.74 0.4 0.44 0.76 0.79
Depth (m) 438.39 442.66 445.27 467.10 469.33 474.19 476.00 476.93 478.48 479.24 555.07 562.66 563.26 564.49
SEAM Iona Argyle Auburn

227
Appendices

Well 3 - Samples 29 30 31 32 33 34 35 36 37 38 39
Gamma ray (API) 58 68 73 87 50 44 49 100 42 40 71
Compensated Density (g/cm3) 1.7 1.7 1.5 1.6 1.4 1.5 1.5 1.6 1.4 1.4 1.6
Mean wireline responses

Vetar Density (g/cm3) 1.64 1.67 1.48 1.54 1.41 1.49 1.46 1.62 1.38 1.36 1.53
NEAR Density (g/cm3) 1.71 1.56 1.60 1.68 1.52 1.57 1.52 1.69 1.47 1.47 1.63
FAR Density (g/cm3) 1.71 1.63 1.53 1.61 1.48 1.55 1.49 1.63 1.41 1.40 1.60
(unfiltered)

Neutron Porosity (percent) 79 80 91 84 82 77 89 90 94 87 89


NEAR Neutron Raw (snu) 5976 5298 5068 5435 5087 5520 4975 5709 4959 4897 5355
FAR Neutron Raw (snu) 243 190 166 208 181 225 164 180 163 162 185
PE (barns/electron) 1.38 1.47 1.22 1.21 0.99 1.23 1.17 1.73 1.20 0.98 1.48
FEFE (ohm.m) 47 48 139 64 93 38 106 13 202 186 89
Medium Induction (ohm.m) 4 3 9 13 16 10 9 6 21 26 13
Deep Induction (ohm.m) 4 3 9 10 13 10 9 6 16 17 10
Sample Length (m) 0.71 0.8 0.81 0.49 0.83 0.53 0.8 0.41 0.81 0.8 0.8
Depth (m) 584.11 588.50 589.31 605.87 606.53 617.91 618.57 619.18 619.79 620.59 621.39
SEAM Bulwer Condamine

228
Appendices

Well 6 - Samples 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Gamma ray (API) 105 63 40 103 80 66 60 39 27 31 24 65 89 36
Compensated Density (g/cm3) 2.41 1.72 1.41 2.17 1.79 1.93 1.65 1.42 1.36 1.45 1.36 1.45 2.21 1.45
Mean wireline responses

Vetar Density (g/cm3) 2.41 1.69 1.44 2.18 1.76 1.91 1.63 1.42 1.34 1.46 1.36 1.47 2.22 1.41
NEAR Density (g/cm3) 2.40 1.67 1.44 2.15 1.77 1.89 1.60 1.41 1.31 1.48 1.37 1.37 2.03 1.43
(unfiltered)

FAR Density (g/cm3) 2.41 1.71 1.41 2.16 1.79 1.92 1.63 1.42 1.34 1.46 1.36 1.41 2.14 1.45
Neutron Porosity (percent) 36 72 88 46 71 56 72 90 89 88 95 80 41 83
NEAR Neutron Raw (snu) 8940 5668 4947 7989 6169 7078 5831 5016 4885 5132 4877 4831 8575 5117
FAR Neutron Raw (snu) 524 216 167 396 237 321 226 165 160 168 152 172 467 176
PE (barns/electron) 3.1 1.8 1.3 2.2 1.5 1.7 1.6 1.4 1.3 1.3 1.2 1.5 2.0 1.3
Shallow Laterolog (ohm.m) 8 15 31 10 25 15 20 37 198 105 413 6 5 48
Deep Laterolog (ohm.m) 9 14 28 10 23 14 17 28 156 88 348 7 6 36
Sample Length (m) 0.80 0.62 0.52 0.82 0.80 0.79 0.80 0.79 0.80 0.79 0.79 0.67 0.78 0.80
Depth (m) 334.9 337.2 337.8 347.6 348.5 349.5 371.1 372.2 373.0 373.8 374.6 380.2 383.5 395.1
SEAM Kogan Macalister Upper M.Lower

Well 6 - Samples 15 16 17 18 19 20 21 22 23 24 25 26 27 28
Gamma ray (API) 59 60 63 89 67 62 42 66 71 87 48 92 93 55
Compensated Density (g/cm3) 1.61 1.61 1.63 1.73 1.65 1.62 1.50 1.48 1.63 1.74 1.41 1.80 2.22 1.57
Mean wireline responses

Vetar Density (g/cm3) 1.57 1.45 1.62 1.70 1.59 1.59 1.42 1.40 1.54 1.72 1.39 1.78 2.29 1.43
NEAR Density (g/cm3) 1.58 1.48 1.61 1.69 1.61 1.49 1.47 1.36 1.49 1.48 1.39 1.50 2.23 1.46
(unfiltered)

FAR Density (g/cm3) 1.61 1.61 1.63 1.72 1.65 1.58 1.51 1.46 1.61 1.67 1.41 1.69 2.21 1.57
Neutron Porosity (percent) 80 66 78 71 73 70 79 68 69 63 87 50 43 68
NEAR Neutron Raw (snu) 5507 5386 5593 5591 5509 5705 5255 5011 5696 5866 4888 6366 8962 5218
FAR Neutron Raw (snu) 207 234 197 218 224 233 203 207 232 248 168 303 480 224
PE (barns/electron) 1.5 1.3 1.6 1.7 1.6 1.4 1.3 1.3 1.5 1.5 1.3 1.9 2.0 1.3
Shallow Laterolog (ohm.m) 16 11 10 11 18 21 64 14 15 9 47 6 10 26
Deep Laterolog (ohm.m) 15 14 10 9 16 18 55 14 14 8 38 5 10 22
Sample Length (m) 0.62 0.33 0.48 0.78 0.55 0.54 0.6 0.4 0.35 0.42 0.43 0.8 0.34 0.36
Depth (m) 404.0 407.7 413.7 439.1 456.2 476.3 477.8 481.2 483.0 484.2 484.7 491.4 492.9 493.2
SEAM M.Lower Nangram Wambo Iona Argyle

229
Appendices

Well 6 - Samples 29 30 31 32 33 34 35 36 37 38 39 40 41 42
Gamma ray (API) 54 44 40 49 48 78 44 43 61 44 98 46 50
Compensated Density (g/cm3) 1.64 1.58 1.39 1.57 1.45 1.66 2.31 1.42 1.65 1.46 1.86 1.41 1.40 2.40
Mean wireline responses

Vetar Density (g/cm3) 1.49 1.52 1.35 1.52 1.41 1.64 2.34 1.41 1.65 1.36 1.83 1.35 1.38 2.40
NEAR Density (g/cm3) 1.52 1.52 1.39 1.55 1.37 1.62 2.30 1.42 1.63 1.38 1.84 1.37 1.42 2.37
(unfiltered)

FAR Density (g/cm3) 1.64 1.58 1.40 1.57 1.43 1.65 2.30 1.42 1.64 1.46 1.86 1.41 1.42 2.39
Neutron Porosity (percent) 65 77 83 80 88 69 27 86 77 77 69 86 92 28
NEAR Neutron Raw (snu) 5776 5362 4912 5276 4934 5329 12597 4925 5711 5187 6140 5069 5166 12161
FAR Neutron Raw (snu) 264 220 177 191 179 213 933 162 205 212 250 182 173 851
PE (barns/electron) 1.2 1.4 1.2 1.4 1.3 1.5 2.2 1.4 1.7 1.1 1.9 1.1 1.2 2.5
Shallow Laterolog (ohm.m) 24 22 29 64 144 19 7 38 11 43 4 121 108 5
Deep Laterolog (ohm.m) 24 23 24 21 101 16 7 26 8 34 5 115 89 4
Sample Length (m) 0.36 0.54 0.41 0.65 0.69 0.8 0.8 0.48 0.8 0.51 0.54 0.45 0.81 0.57
Depth (m) 515.0 549.2 550.1 553.5 555.3 557.3 575.2 592.0 593.4 619.9 626.2 633.1 633.7 651.0
SEAM T. Std Auburn Bulwer Condamine

230
Appendices

Well 8- Sample 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Gamma ray (API) 59 73 55 22 35 51 78 19 38 36 60 47 70 74
Compensated Density (g/cm3) 1.78 1.69 1.49 1.35 1.38 1.46 1.75 1.33 1.49 1.32 1.71 1.70 2.18 1.80
Mean wireline responses

Vetar Density (g/cm3) 1.76 1.70 1.50 1.35 1.38 1.44 1.78 1.31 1.51 1.31 1.58 1.68 2.16 1.75
NEAR Density (g/cm3) 1.74 1.63 1.43 1.30 1.32 1.36 1.71 1.26 1.47 1.26 1.57 1.64 2.09 1.69
(unfiltered)

FAR Density (g/cm3) 1.77 1.67 1.47 1.33 1.36 1.43 1.72 1.31 1.48 1.30 1.70 1.68 2.15 1.77
Neutron Porosity (percent) 76 72 83 87 89 84 73 83 83 84 69 78 44 68
NEAR Neutron Raw (snu) 5384 5921 5344 5101 5154 5455 6410 5143 5638 5023 6062 6403 8909 6977
FAR Neutron Raw (snu) 197 225 180 169 169 184 238 176 189 170 241 225 490 299
PE (barns/electron) 2.5 1.6 1.4 1.4 1.4 1.4 1.7 1.3 1.4 1.3 2.0 1.5 2.2 1.7
Shallow Laterolog (ohm.m) 19 64 145 606 241 126 59 1468 114 227 38 44 5 8
Deep Laterolog (ohm.m) 16 51 115 467 179 89 45 1160 101 212 36 41 2 8
Sample Length (m) 0.73 0.61 0.75 0.82 0.79 0.82 0.81 0.8 0.78 0.74 0.37 0.73 0.66 0.56
Depth (m) 316 329 330 331 332 332 333 334 335 336 341 342 363 368
SEAM Macalister Upper Macalister Lower

Well 8 - Sample 15 16 17 18 19 20 21 22 23 24 25 26 27 28
Gamma ray (API) 65 61 78 57 53 47 53 47 39 80 51 68 87 91
Compensated Density (g/cm3) 1.68 1.67 1.81 1.52 1.54 1.45 1.58 1.60 1.45 1.97 1.53 1.63 1.79 1.91
Mean wireline responses

Vetar Density (g/cm3) 1.64 1.55 1.79 1.49 1.41 1.37 1.47 1.45 1.31 1.91 1.47 1.51 1.74 1.83
NEAR Density (g/cm3) 1.59 1.50 1.75 1.45 1.39 1.34 1.42 1.42 1.28 1.88 1.42 1.46 1.70 1.76
(unfiltered)

FAR Density (g/cm3) 1.66 1.64 1.79 1.50 1.52 1.43 1.55 1.58 1.43 1.96 1.51 1.60 1.77 1.88
Neutron Porosity (percent) 69 69 65 77 84 74 73 69 73 51 71 64 64 59
NEAR Neutron Raw (snu) 6259 5975 6860 5575 5323 5298 5517 5632 5332 8049 5991 5697 6524 6859
FAR Neutron Raw (snu) 259 260 305 209 208 207 227 248 229 409 254 258 289 318
PE (barns/electron) 1.6 1.6 1.7 1.5 1.5 1.3 1.4 1.4 1.3 1.8 1.5 1.4 1.7 1.9
Shallow Laterolog (ohm.m) 30 22 18 29 17 37 20 33 26 14 35 16 15 8
Deep Laterolog (ohm.m) 29 22 17 26 16 33 20 30 24 13 33 16 13 8
Sample Length (m) 0.69 0.43 0.77 0.8 0.43 0.74 0.56 0.38 0.42 0.52 0.65 0.45 0.62 0.63
Depth (m) 376 411 443 452 453 454 458 475 476 482 495 505 516 527
SEAM Macalister Lower Nangram Wambo Iona

231
Appendices

Well 8 - Sample 29 30 31 32 33 34 35
Gamma ray (API) 55 73 49 62 77 94 64
Compensated Density (g/cm3) 1.61 1.66 1.38 1.64 1.65 1.74 1.64
Mean wireline responses

Vetar Density (g/cm3) 1.57 1.61 1.38 1.61 1.59 1.69 1.50
NEAR Density (g/cm3) 1.54 1.56 1.31 1.56 1.52 1.67 1.44
(unfiltered)

FAR Density (g/cm3) 1.59 1.64 1.36 1.62 1.62 1.73 1.61
Neutron Porosity (percent) 77 73 95 70 73 66 61
NEAR Neutron Raw (snu) 6077 5827 5165 5903 6141 6431 6039
FAR Neutron Raw (snu) 234 237 161 241 248 269 275
PE (barns/electron) 1.5 1.6 1.4 1.6 1.5 1.7 1.5
Shallow Laterolog (ohm.m) 29 16 54 24 17 9 11
Deep Laterolog (ohm.m) 27 15 49 21 16 9 11
Sample Length (m) 0.82 0.71 0.8 0.8 0.64 0.52 0.39
Depth (m) 527 533 551 552 599 620 621
SEAM Iona Argyle Auburn Condamine

232
Appendices

Well 10- Sample 1 2 3 4 5 6 7 8 9 10 11 12 13 14


Gamma ray (API) 51 63 47 45 47 26 28 22 56 24 40 29 45 33
Compensated Density (g/cm3) 1.59 1.69 1.58 1.60 1.56 1.36 1.34 1.37 1.56 1.33 1.42 1.44 1.54 1.52
Mean wireline responses

Vetar Density (g/cm3) 1.52 1.59 1.50 1.56 1.44 1.34 1.35 1.29 1.45 1.29 1.36 1.37 1.49 1.42
NEAR Density (g/cm3) 1.47 1.51 1.42 1.49 1.38 1.26 1.30 1.23 1.37 1.21 1.28 1.25 1.44 1.33
(unfiltered)

FAR Density (g/cm3) 1.57 1.65 1.55 1.57 1.53 1.33 1.33 1.34 1.53 1.30 1.39 1.39 1.52 1.48
Neutron Porosity (percent) 84 77 81 87 78 104 103 96 78 100 91 92 82 83
NEAR Neutron Raw (snu) 4972 5076 4749 5016 4884 4449 4467 4548 4823 4444 4547 4535 4838 4757
FAR Neutron Raw (snu) 174 201 176 179 193 139 133 160 192 138 160 158 172 181
PE (barns/electron) 1.57 1.68 1.54 1.78 1.41 1.34 1.50 1.29 1.39 1.30 1.32 1.21 1.40 1.30
Shallow Laterolog (ohm.m) 27 15 12 27 15 137 110 95 21 80 51 54 20 27
Deep Laterolog (ohm.m) 8 14 12 26 12 138 100 80 20 78 46 51 20 28
Sample Length (m) 0.54 0.49 0.59 0.74 0.32 0.77 0.75 0.8 0.43 0.98 0.73 0.53 0.74 0.52
Depth (m) 85.1 100.0 111.7 145.2 147.4 177.5 178.3 179.6 206.7 208.0 210.5 291.7 292.7 297.9
SEAM Nangram Wambo Iona Argyle

Well 10 - Sample 15 16 17 18 19 20 21 22 23 24 25 26 27
Gamma ray (API) 40 36 26 43 47 38 41 38 40 42 45 53 37
Compensated Density (g/cm3) 1.44 1.63 1.40 1.52 1.53 1.51 1.47 1.53 1.49 1.46 1.52 1.52 1.48
Mean wireline responses

Vetar Density (g/cm3) 1.40 1.58 1.36 1.42 1.45 1.41 1.37 1.43 1.42 1.37 1.50 1.54 1.39
NEAR Density (g/cm3) 1.32 1.49 1.29 1.33 1.38 1.37 1.29 1.36 1.35 1.30 1.42 1.46 1.35
(unfiltered)

FAR Density (g/cm3) 1.41 1.60 1.37 1.48 1.50 1.49 1.44 1.49 1.46 1.43 1.49 1.50 1.46
Neutron Porosity (percent) 88 74 93 47 81 76 80 85 88 91 85 87 76
NEAR Neutron Raw (snu) 4687 5393 4592 4917 5032 4940 4806 5043 4950 4810 4917 5051 4946
FAR Neutron Raw (snu) 166 217 151 187 194 197 185 190 183 169 172 165 200
PE (barns/electron) 1.33 1.56 1.29 1.32 1.28 1.17 1.20 1.26 1.31 1.20 1.42 1.44 1.25
Shallow Laterolog (ohm.m) 43 25 50 23 27 34 35 30 40 31 84 133 28
Deep Laterolog (ohm.m) 40 23 46 20 26 26 31 32 46 28 70 96 24
Sample Length (m) 0.7 0.49 0.79 0.54 0.65 0.52 0.56 0.63 0.55 0.54 0.7 0.74 0.41
Depth (m) 298.8 315.0 315.7 316.6 402.1 410.6 412.0 418.1 439.3 444.3 445.3 446.0 446.5
SEAM Argyle Bulwer Condamine

233
Appendices

Well 11 - Sample 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Gamma ray (API) 54 43 31 28 65 31 58 30 26 13 18 27 22 29
Compensated Density (g/cm3) 2.26 1.62 1.70 1.47 2.25 1.62 1.90 1.57 1.45 1.38 1.44 1.49 1.43 1.47
Mean wireline responses

Vetar Density (g/cm3) 2.27 1.58 1.66 1.44 2.27 1.46 1.88 1.57 1.45 1.38 1.42 1.48 1.44 1.47
NEAR Density (g/cm3) 1.85 1.54 1.60 1.42 2.19 1.45 1.79 1.48 1.35 1.29 1.33 1.41 1.39 1.37
(unfiltered)

FAR Density (g/cm3) 2.11 1.60 1.68 1.46 2.22 1.61 1.86 1.54 1.41 1.35 1.41 1.46 1.41 1.43
Neutron Porosity (percent) 37 81 69 83 42 71 59 82 98 88 92 89 88 87
NEAR Neutron Raw (snu) 6338 4829 5628 4822 8340 4905 6338 5030 4835 4750 4801 5002 4868 4760
FAR Neutron Raw (snu) 347 166 247 168 439 208 280 170 149 153 156 159 158 155
PE (barns/electron) 2.32 1.36 1.46 1.20 2.13 1.24 1.94 1.45 1.44 1.31 1.41 1.45 1.38 1.50
Shallow Laterolog (ohm.m) 5 37 15 28 7 16 13 39 101 239 142 109 167 169
Deep Laterolog (ohm.m) 4 38 14 25 5 16 11 32 77 179 105 81 129 146
Sample Length (m) 0.480 0.490 0.440 0.760 0.380 0.320 0.760 0.820 0.750 0.350 0.640 0.630 0.820
Depth (m) 308.36 311.845 329.52 330.07 332.27 333.67 357.5 358.29 359.1 359.875 360.425 360.92 361.575 362.32
SEAM Kogan Macalister Upper Macalister Lower

Well 11 - Sample 15 16 17 18 19 20 21 22 23 24 25 26 27 28
Gamma ray (API) 7 15 24 28 34 45 64 35 37 65 20 16 37 49
Compensated Density (g/cm3) 1.33 1.36 1.51 1.57 2.27 1.64 2.30 1.61 1.57 2.33 1.38 1.39 1.61 1.65
Mean wireline responses

Vetar Density (g/cm3) 1.32 1.35 1.51 1.54 2.27 1.60 2.31 1.58 1.53 2.39 1.34 1.37 1.55 1.55
NEAR Density (g/cm3) 1.23 1.29 1.45 1.52 2.22 1.59 2.17 1.50 1.49 2.22 1.30 1.30 1.55 1.54
(unfiltered)

FAR Density (g/cm3) 1.29 1.33 1.48 1.56 2.25 1.63 2.25 1.58 1.56 2.27 1.36 1.36 1.61 1.64
Neutron Porosity (percent) 92 91 84 76 29 79 42 78 75 41 87 88 72 73
NEAR Neutron Raw (snu) 4488 4745 4995 5348 11025 5188 8540 5154 5201 8502 4765 4758 5366 5207
FAR Neutron Raw (snu) 142 150 167 195 745 192 458 190 191 450 162 158 202 207
PE (barns/electron) 1.27 1.22 1.31 1.29 2.36 1.53 2.35 1.53 1.52 3.03 1.00 1.30 1.29 1.34
Shallow Laterolog (ohm.m) 370 210 41 20 6 12 4 18 23 6 114 61 23 18
Deep Laterolog (ohm.m) 321 178 34 17 6 10 3 15 19 6 128 55 18 16
Sample Length (m) 0.720 0.550 0.630 0.530 0.660 0.780 0.500 0.330 0.730 0.700 0.380 0.620 0.555
Depth (m) 363.09 363.725 364.31 364.89 390.06 430.03 432.15 433.2 465.4 465.97 492.23 493.07 501.41 502.34
SEAM Macalister Lower Nangram Wambo Iona Argyle

234
Appendices

Well 11 - Sample 29 30 31 32 33 34 35 36 37 38 39 40
Gamma ray (API) 29 27 35 32 55 19 36 36 28 36 28 63
Compensated Density (g/cm3) 1.56 1.41 1.52 1.55 1.83 1.42 1.62 1.56 1.43 1.47 1.46 2.39
Mean wireline responses

Vetar Density (g/cm3) 1.50 1.39 1.45 1.52 1.76 1.37 1.59 1.45 1.39 1.44 1.43 2.42
NEAR Density (g/cm3) 1.46 1.37 1.30 1.46 1.72 1.33 1.52 1.39 1.35 1.38 1.41 2.38
(unfiltered)

FAR Density (g/cm3) 1.54 1.41 1.46 1.52 1.82 1.40 1.59 1.53 1.41 1.44 1.45 2.38
Neutron Porosity (percent) 85 79 78 83 60 83 79 79 91 85 78 35
NEAR Neutron Raw (snu) 5160 4889 4777 5068 6083 4882 5136 5322 4906 4774 4989 10787
FAR Neutron Raw (snu) 189 168 178 179 268 172 183 205 165 163 177 684
PE (barns/electron) 1.31 1.13 1.22 1.33 1.68 1.13 1.37 1.29 1.31 1.41 1.36 2.36
Shallow Laterolog (ohm.m) 61 89 15 47 11 80 15 35 202 33 90 7
Deep Laterolog (ohm.m) 54 83 15 45 10 78 13 29 178 35 92 7
Sample Length (m) 0.400 0.710 0.350 0.800 0.490 0.710 0.630 0.440 0.800 0.800 0.760 0.400
Depth (m) 503.31 510.60 511.16 518.03 521.91 524.10 613.38 649.70 652.22 655.46 656.24 657.69
SEAM Argyle Auburn Bulwer

Well 11 - Sample 41 42 43 44 45 46 47 48
Gamma ray (API) 32 56 34 23 46 28 34 65
Compensated Density (g/cm3) 1.57 2.49 1.55 1.43 1.60 1.56 1.55 2.37
Mean wireline responses

Vetar Density (g/cm3) 1.48 2.49 1.41 1.35 1.55 1.50 1.47 2.38
NEAR Density (g/cm3) 1.37 2.40 1.37 1.32 1.52 1.44 1.44 2.33
(unfiltered)

FAR Density (g/cm3) 1.53 2.45 1.53 1.42 1.58 1.53 1.53 2.35
Neutron Porosity (percent) 73 27 77 76 71 77 75 35
NEAR Neutron Raw (snu) 5298 11870 5222 4958 5196 5357 5303 9624
FAR Neutron Raw (snu) 214 847 213 188 203 205 210 576
PE (barns/electron) 1.21 2.49 1.12 1.18 1.44 1.35 1.17 2.37
Shallow Laterolog (ohm.m) 14 12 16 34 17 30 38 5
Deep Laterolog (ohm.m) 14 11 15 29 14 26 28 4
Sample Length (m) 0.430 0.690 0.350 0.530 0.730 0.690 0.710
Depth (m) 663.72 673.75 675.00 676.61 679.29 689.88 691.20 693.80
SEAM Condamine

235
Appendices

Well 12 - Sample 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Gamma ray (API) 29 31 36 65 34 54 104 40 50 63 48 55 55 41
Compensated Density (g/cm3) 1.43 1.37 1.36 1.55 1.41 1.61 2.34 1.50 1.48 1.55 1.78 1.55 1.59 1.58
Mean wireline responses

Vetar Density (g/cm3) 1.37 1.31 1.32 1.52 1.37 1.49 2.34 1.44 1.43 1.53 1.68 1.52 1.57 1.45
NEAR Density (g/cm3) 1.32 1.31 1.29 1.47 1.30 1.45 2.28 1.41 1.38 1.49 1.76 1.47 1.59 1.42
(unfiltered)

FAR Density (g/cm3) 1.41 1.37 1.35 1.53 1.38 1.59 2.32 1.49 1.46 1.53 1.80 1.54 1.59 1.56
Neutron Porosity (percent) 83 88 88 79 83 68 40 81 88 83 68 78 83 72
NEAR Neutron Raw (snu) 4600 4532 4561 4795 4571 5033 8372 4654 4758 4810 5348 4961 5198 5003
FAR Neutron Raw (snu) 173 155 163 173 157 214 459 167 171 168 230 186 185 210
PE (barns/electron) 1.50 1.23 1.46 1.69 1.46 1.45 2.61 1.38 1.50 1.53 2.44 1.62 1.47 1.39
Shallow Laterolog (ohm.m) 77 87 54 12 50 15 4 12 21 18 20 16 37 27
Deep Laterolog (ohm.m) 65 67 59 12 49 17 4 12 23 17 19 17 37 23
Sample Length (m) 0.45 0.5 0.77 0.77 0.83 0.435 0.775 0.37 0.72 0.73 0.56 0.5 0.66 0.69
Depth (m) 153.24 153.71 159.03 166.90 167.70 170.65 173.15 181.43 185.78 186.70 188.85 190.79 191.38 195.25

Well 12 - Sample 15 16 17 18 19 20 21 22 23 24 25 26 27 28
Gamma ray (API) 39 33 55 101 50 61 60 52 68 46 52 38 38 51
Compensated Density (g/cm3) 1.44 1.42 1.59 2.33 1.60 1.74 1.56 1.62 1.65 1.58 1.54 1.37 1.35 1.52
Mean wireline responses

Vetar Density (g/cm3) 1.38 1.34 1.50 2.36 1.54 1.71 1.53 1.55 1.63 1.54 1.51 1.32 1.31 1.42
NEAR Density (g/cm3) 1.37 1.31 1.47 2.28 1.50 1.66 1.51 1.52 1.60 1.49 1.46 1.28 1.27 1.40
(unfiltered)

FAR Density (g/cm3) 1.43 1.41 1.57 2.31 1.58 1.72 1.54 1.61 1.64 1.56 1.52 1.35 1.33 1.50
Neutron Porosity (percent) 79 85 80 39 76 72 72 73 71 78 77 86 89 79
NEAR Neutron Raw (snu) 4633 4746 4992 8823 5212 5786 5298 5233 5404 5336 5086 4484 4513 4789
FAR Neutron Raw (snu) 169 168 195 485 189 216 207 214 209 211 188 156 156 187
PE (barns/electron) 1.20 1.34 1.49 2.41 1.58 1.83 1.51 1.69 1.87 1.73 1.63 1.49 1.46 1.28
Shallow Laterolog (ohm.m) 32 43 20 4 25 17 27 28 22 16 17 42 50 23
Deep Laterolog (ohm.m) 29 40 19 5 24 15 23 27 22 16 15 43 48 21
Sample Length (m) 0.4 0.62 0.51 0.57 0.66 0.79 0.6 0.58 0.56 0.77 0.64 0.78 0.8 0.62
Depth (m) 193.93 196.00 197.93 228.63 229.58 230.43 231.11 233.66 239.95 263.85 264.60 269.18 330.71 335.13

236
Appendices

Well 12 - Sample 29 30 31 32 33 34 35 36 37 38 39 40 41 42
Gamma ray (API) 19 28 108 68 62 106 41 52 37 71 65 96 49 87
Compensated Density (g/cm3) 1.31 1.32 2.40 2.34 1.64 2.02 1.49 1.47 1.44 1.65 1.61 2.14 1.48 1.89
Mean wireline responses

Vetar Density (g/cm3) 1.28 1.28 2.40 2.34 1.48 1.98 1.42 1.36 1.35 1.55 1.44 2.12 1.39 1.93
NEAR Density (g/cm3) 1.24 1.28 2.32 2.25 1.46 1.94 1.38 1.35 1.31 1.51 1.44 2.04 1.38 1.87
(unfiltered)

FAR Density (g/cm3) 1.30 1.31 2.37 2.30 1.62 2.00 1.47 1.46 1.42 1.63 1.60 2.11 1.47 1.87
Neutron Porosity (percent) 98 88 34 39 72 57 79 83 79 69 61 44 86 64
NEAR Neutron Raw (snu) 4539 4529 9234 9232 5047 6572 4884 4869 4718 5247 5062 7964 4888 6237
FAR Neutron Raw (snu) 148 154 565 519 208 290 188 180 178 226 242 430 178 251
PE (barns/electron) 1.49 1.52 2.50 2.35 1.48 1.90 1.54 1.36 1.31 1.59 1.40 2.26 1.48 2.04
Shallow Laterolog (ohm.m) 126 84 5 4 9 9 26 23 26 18 16 11 15 8
Deep Laterolog (ohm.m) 114 71 5 4 8 8 24 21 24 15 14 10 15 7
Sample Length (m) 0.78 0.8 0.8 0.69 0.32 0.43 0.54 0.41 0.57 0.43 0.32 0.71 0.49 0.51
Depth (m) 336.66 337.45 342.80 351.59 353.58 364.48 365.90 366.54 369.45 371.79 372.45 376.20 390.84 391.34

Well 12 - Sample 43 44 45 46 47 48 49 50 51 52 53 54 55 56
Gamma ray (API) 47 63 50 86 102 51 58 38 54 54 59 46 34 66
Compensated Density (g/cm3) 1.51 1.56 1.49 2.19 2.26 1.53 1.58 1.37 1.58 1.58 1.60 1.35 1.36 2.43
Mean wireline responses

Vetar Density (g/cm3) 1.43 1.57 1.38 2.20 2.24 1.51 1.60 1.33 1.51 1.51 1.51 1.29 1.34 2.43
NEAR Density (g/cm3) 1.38 1.59 1.35 2.08 2.17 1.45 1.55 1.28 1.46 1.49 1.48 1.32 1.31 2.36
(unfiltered)

FAR Density (g/cm3) 1.48 1.57 1.48 2.15 2.23 1.51 1.56 1.35 1.56 1.57 1.58 1.36 1.35 2.41
Neutron Porosity (percent) 84 77 74 46 46 83 79 94 78 77 77 86 86 25
NEAR Neutron Raw (snu) 4902 5305 4854 8570 7601 4887 5280 4638 5318 5085 5160 4799 4774 12436
FAR Neutron Raw (snu) 177 193 190 424 385 172 185 151 204 200 207 170 167 992
PE (barns/electron) 1.55 1.90 1.28 2.05 2.33 1.65 1.86 1.43 1.52 1.54 1.45 1.28 1.48 2.61
Shallow Laterolog (ohm.m) 24 17 16 8 5 32 24 47 19 14 18 48 80 7
Deep Laterolog (ohm.m) 21 14 13 7 4 27 19 34 15 14 18 43 83 8
Sample Length (m) 0.44 0.76 0.41 0.67 0.54 0.77 0.77 0.64 0.63 0.66 0.69 0.79 0.37 0.79
Depth (m) 409.49 410.39 410.98 461.70 467.65 470.58 471.33 472.04 473.20 476.50 493.14 500.38 499.79 506.81

237
Appendices

Well 12 - Sample 57 58 59 60 61 62 63 64 65 66 67 68 69 70
Gamma ray (API) 69 80 68 95 90 44 39 30 72 75 57 38 37 65
Compensated Density (g/cm3) 1.67 1.74 1.56 2.48 1.99 1.44 1.35 1.28 1.65 1.75 1.51 1.43 1.45 1.57
Mean wireline responses

Vetar Density (g/cm3) 1.55 1.61 1.45 2.47 1.95 1.45 1.35 1.24 1.57 1.68 1.45 1.39 1.46 1.56
NEAR Density (g/cm3) 1.53 1.58 1.42 2.44 1.89 1.38 1.29 1.21 1.53 1.65 1.43 1.34 1.44 1.48
(unfiltered)

FAR Density (g/cm3) 1.65 1.72 1.54 2.47 1.97 1.41 1.33 1.27 1.62 1.74 1.50 1.41 1.44 1.54
Neutron Porosity (percent) 62 63 79 37 53 91 91 87 70 69 78 82 90 80
NEAR Neutron Raw (snu) 5762 5483 5076 9268 7275 4808 4753 4577 5409 6125 5168 5040 5184 5420
FAR Neutron Raw (snu) 265 255 200 547 355 159 154 157 229 256 190 193 171 205
PE (barns/electron) 1.57 1.73 1.46 3.72 1.78 1.60 1.54 1.32 1.60 1.60 1.56 1.42 1.49 1.69
Shallow Laterolog (ohm.m) 8 6 14 5 10 86 118 116 15 11 14 51 51 39
Deep Laterolog (ohm.m) 9 7 14 5 8 77 110 108 16 12 14 55 48 30
Sample Length (m) 0.36 0.32 0.49 0.8 0.72 0.73 0.82 0.75 0.51 0.51 0.76 0.62 0.61 0.63
Depth (m) 514.06 517.80 523.38 533.74 538.95 540.50 541.28 542.06 547.60 551.45 554.40 566.33 566.95 568.06

Well 12 - Sample 71 72 73 74 75 76 77 78
Gamma ray (API) 35 65 53 131 48 111 33 44
Compensated Density (g/cm3) 1.34 1.89 1.49 1.84 1.53 1.78 1.39 1.40
Mean wireline responses

Vetar Density (g/cm3) 1.30 1.90 1.47 1.78 1.36 1.66 1.35 1.28
NEAR Density (g/cm3) 1.28 1.80 1.44 1.74 1.30 1.61 1.30 1.27
(unfiltered)

FAR Density (g/cm3) 1.33 1.86 1.47 1.82 1.50 1.75 1.37 1.39
Neutron Porosity (percent) 94 60 77 66 73 60 83 71
NEAR Neutron Raw (snu) 4815 7004 5311 5729 5112 5576 4981 4987
FAR Neutron Raw (snu) 156 329 204 243 214 249 184 197
PE (barns/electron) 1.36 2.24 1.45 1.75 1.05 1.53 1.48 1.15
Shallow Laterolog (ohm.m) 71 15 28 8 14 8 89 32
Deep Laterolog (ohm.m) 54 14 29 7 12 6 105 27
Sample Length (m) 0.65 0.83 0.54 0.36 0.32 0.35 0.6 0.27
Depth (m) 568.71 578.15 578.84 584.84 586.48 587.16 589.41 589.85

238
Appendices

Well 13 - Sample 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Gamma ray (API) 62 78 86 80 78 32 49 88 63 87 78 37 57 47
Compensated Density (g/cm3) 1.7 1.7 1.8 1.7 2.3 1.4 1.5 2.3 1.5 1.6 2.0 1.4 1.7 1.5
Mean wireline responses

Vetar Density (g/cm3) 1.70 1.34 1.81 1.70 2 1.41 1.49 2.35 1.49 1.61 2.02 1.42 1.72 1.50
NEAR Density (g/cm3) 1.75 1.86 1.90 1.82 2 1.57 1.67 2.37 1.56 1.50 2.05 1.52 1.85 1.62
(unfiltered)

FAR Density (g/cm3) 1.75 1.78 1.83 1.77 2 1.47 1.56 2.34 1.53 1.59 2.02 1.47 1.77 1.56
Neutron Porosity (percent) 66 69 81 79 51 88 82 42 79 58 59 87 74 83
NEAR Neutron Raw (snu) 5062 4808 4736 4482 6643 4369 4450 8672 4582 4002 6243 4251 5484 4594
FAR Neutron Raw (snu) 218 181 164 161 316 149 159 232 164 184 255 150 201 170
PE (barns/electron) 1.31 20.30 1.39 1.35 2.01 0.96 0.90 2.17 1.25 1.64 1.60 1.05 1.24 0.73
Shallow Laterolog (ohm.m) 17 18 16 10 4 34 22 6 14 9 19 73 40 29
Deep Laterolog (ohm.m) 17 390 14 10 4 36 24 6 13 7 17 70 36 29
Sample Length (m) 0.51 0.80 0.78 0.50 0.45 0.75 0.49 0.80 0.52 0.35 0.76 0.77 0.66 0.51
Depth (m) 387.7 388.3 389.1 389.8 390.2 410.8 411.5 412.2 416.6 418.4 420.2 427.4 428.1 431.5
SEAM Springbok Std Kogan Macalister Upper Macalister Lower

Well 13 - Sample 15 16 17 18 19 20 21 22 23 24 25 26 27 28
Gamma ray (API) 48 53 38 78 101 47 48 92 32 91 71 64 66 49
Compensated Density (g/cm3) 1.6 1.5 1.4 1.9 2.4 1.5 1.6 2.1 1.4 2.1 1.6 1.5 1.6 1.4
Mean wireline responses

Vetar Density (g/cm3) 1.56 1.48 1.38 1.93 2.36 1.49 1.54 2.07 1.39 2.12 1.67 1.54 1.62 1.43
NEAR Density (g/cm3) 1.72 1.60 1.58 1.95 2.34 1.56 1.65 1.81 1.56 1.82 1.68 1.64 1.61 1.56
(unfiltered)

FAR Density (g/cm3) 1.61 1.53 1.45 1.94 2.36 1.55 1.61 2.01 1.46 2.02 1.66 1.59 1.61 1.47
Neutron Porosity (percent) 81 80 90 60 37 68 82 45 88 49 69 74 73 87
NEAR Neutron Raw (snu) 4638 4468 4324 6046 7959 4628 4779 6780 4265 6366 4951 4646 4461 4051
FAR Neutron Raw (snu) 162 161 141 265 477 198 174 349 148 322 200 181 173 142
PE (barns/electron) 1.08 0.91 0.47 1.56 2.05 1.03 0.94 1.98 0.87 2.00 1.39 0.98 1.52 0.98
Shallow Laterolog (ohm.m) 27 42 44 13 6 8 16 11 60 4 15 21 8 22
Deep Laterolog (ohm.m) 25 53 41 12 6 9 15 10 54 4 13 19 8 23
Sample Length (m) 0.78 0.72 0.78 0.75 0.36 0.40 0.76 0.23 0.83 0.56 0.82 0.58 0.80 0.82
Depth (m) 433.2 437.8 458.1 458.9 460.1 472.7 485.4 500.5 501.0 538.7 550.2 551.3 571.7 572.9
SEAM Macalister Lower Nangram Wambo Iona Argyle

239
Appendices

Well 13 - Sample 29 30 31 32 33 34 35 36 37 38 39 40 41 42
Gamma ray (API) 47 60 36 40 65 86 49 86 80 50 51 42 99 53
Compensated Density (g/cm3) 1.4 1.6 1.4 1.5 1.5 1.8 1.4 1.9 1.8 1.6 1.6 1.4 2.1 1.5
Mean wireline responses

Vetar Density (g/cm3) 1.37 1.60 1.39 1.52 1.53 1.82 1.44 1.94 1.83 1.59 1.56 1.43 2.08 1.48
NEAR Density (g/cm3) 1.54 1.63 1.53 1.56 1.58 1.79 1.60 1.65 1.81 1.70 1.64 1.54 1.99 1.58
(unfiltered)

FAR Density (g/cm3) 1.41 1.62 1.45 1.54 1.53 1.83 1.50 1.82 1.82 1.67 1.61 1.50 2.07 1.55
Neutron Porosity (percent) 95 76 84 79 83 62 88 58 62 66 72 81 51 87
NEAR Neutron Raw (snu) 4203 4373 4104 4553 4121 5174 4437 5589 5503 5352 4874 4573 6201 4587
FAR Neutron Raw (snu) 136 175 142 175 144 232 156 249 232 249 198 179 306 168
PE (barns/electron) 0.69 1.05 0.77 1.04 1.17 1.58 0.97 1.80 1.56 0.96 0.95 0.52 1.96 1.07
Shallow Laterolog (ohm.m) 22 8 22 13 8 8 26 6 7 11 12 13 6 30
Deep Laterolog (ohm.m) 24 8 24 12 8 8 26 6 6 11 11 14 5 25
Sample Length (m) 0.49 0.34 0.75 0.80 0.51 0.73 0.80 0.40 0.78 0.53 0.55 0.41 0.43 0.66
Depth (m) 580.0 582.9 585.3 602.2 603.2 611.1 626.9 627.4 654.6 666.4 668.7 671.5 691.7 692.9
SEAM Argyle Tangalooma Std. Auburn

Well 13 - Sample 43 44 45 46 47 48 49 50 51 52 53 54 55 56
Gamma ray (API) 58 50 48 52 60 38 64 79 59 46 31 57 47 38
Compensated Density (g/cm3) 1.6 1.5 1.7 1.6 2.4 1.4 1.8 1.9 1.7 1.5 1.4 1.7 1.5 1.5
Mean wireline responses

Vetar Density (g/cm3) 1.54 1.52 1.67 1.55 2.36 1.41 1.79 1.87 1.65 1.48 1.35 1.68 1.48 1.43
NEAR Density (g/cm3) 1.68 1.67 1.75 1.69 2.37 1.60 1.94 1.94 1.78 1.68 1.56 1.84 1.64 1.56
(unfiltered)

FAR Density (g/cm3) 1.60 1.60 1.74 1.61 2.36 1.49 1.82 1.92 1.70 1.53 1.43 1.73 1.54 1.51
Neutron Porosity (percent) 83 80 64 76 30 82 76 61 73 89 93 75 81 82
NEAR Neutron Raw (snu) 4724 4808 5574 4964 10553 4925 5806 6025 5508 4674 4504 5524 4568 4854
FAR Neutron Raw (snu) 170 180 267 190 731 188 209 270 219 154 149 201 174 184
PE (barns/electron) 1.10 0.85 1.09 1.13 2.23 0.55 1.30 1.56 1.23 1.14 0.94 1.24 0.98 0.64
Shallow Laterolog (ohm.m) 18 19 12 17 10 33 10 10 54 101 130 26 48 36
Deep Laterolog (ohm.m) 18 21 11 15 10 31 9 10 54 108 138 25 54 32
Sample Length (m) 0.80 0.56 0.34 0.75 0.78 0.82 0.39 0.44 0.81 0.81 0.77 0.47 0.75 0.78
Depth (m) 695.5 701.0 715.3 716.6 727.4 774.9 775.5 778.3 781.4 782.2 783.0 783.6 785.6 819.9
SEAM Auburn Bulwer Condamine

240
Appendices

Well 13 - Sample 57 58 59
Gamma ray (API) 42 49 51
Compensated Density (g/cm3) 1.4 1.5 1.6
Mean wireline responses

Vetar Density (g/cm3) 1.40 1.45 1.56


NEAR Density (g/cm3) 1.56 1.65 1.75
(unfiltered)

FAR Density (g/cm3) 1.48 1.53 1.62


Neutron Porosity (percent) 81 92 79
NEAR Neutron Raw (snu) 4612 4754 5083
FAR Neutron Raw (snu) 180 159 193
PE (barns/electron) 0.75 0.75 0.94
Shallow Laterolog (ohm.m) 34 29 19
Deep Laterolog (ohm.m) 30 27 18
Sample Length (m) 0.65 0.54 0.47
Depth (m) 821.2 842.6 843.1
SEAM Condamine

241
Appendices

Well 14 - Sample 1 2 3 4 5 6 7 8 9 10 11 12 13 14
Gamma ray (API) 25 33 100 32 38 22 30 102 29 88 66 60 70 46
Compensated Density (g/cm3) 1.45 1.52 2.35 1.42 1.41 1.46 1.39 2.44 1.47 2.27 1.65 1.67 1.62 1.51
Mean wireline responses

Vetar Density (g/cm3) 1.43 1.49 2.37 1.44 1.35 1.40 1.36 2.44 1.42 2.27 1.59 1.63 1.56 1.44
NEAR Density (g/cm3) 1.50 1.59 2.33 1.54 1.48 1.47 1.46 2.40 1.51 2.24 1.64 1.71 1.64 1.52
(unfiltered)

FAR Density (g/cm3) 1.47 1.55 2.33 1.46 1.45 1.48 1.43 2.42 1.50 2.26 1.66 1.69 1.64 1.53
Neutron Porosity (percent) 89 76 37 79 88 80 91 29 81 44 73 76 79 76
NEAR Neutron Raw (snu) 4869 5271 9040 4895 4805 5038 4809 10213 4966 8031 5388 5893 5303 4969
FAR Neutron Raw (snu) 158 190 509 168 159 197 158 692 185 410 218 237 199 193
PE (barns/electron) 1.19 1.04 2.32 1.05 0.88 0.95 0.90 2.48 0.95 2.10 1.21 1.34 1.24 1.19
Shallow Laterolog (ohm.m) 40 34 5 380 70 115 169 8 154 3 26 25 15 22
Deep Laterolog (ohm.m) 33 28 5 349 54 128 178 9 126 5 22 29 16 24
Sample Length (m) 0.76 0.79 0.77 0.80 0.33 0.45 0.79 0.81 0.75 0.80 0.66 0.71 0.77 0.63
Depth (m) 583.22 584.09 588.24 605.83 606.40 616.69 617.31 621.91 629.62 644.71 650.02 653.96 679.72 684.64
SEAM Macalister Lower Nangram Wambo Iona Argyle

Well 14 - Sample 15 16 17 18 19 20 21 22 23 24 25 26 27 28
Gamma ray (API) 58 78 98 52 27 19 46 39 66 52 64 74 102 94
Compensated Density (g/cm3) 1.60 1.77 1.77 1.58 1.37 1.32 1.47 1.37 1.64 1.48 1.86 1.72 2.06 1.83
Mean wireline responses

Vetar Density (g/cm3) 1.52 1.74 1.76 1.46 1.33 1.28 1.44 1.35 1.61 1.41 1.74 1.59 2.04 1.79
NEAR Density (g/cm3) 1.53 1.75 1.62 1.54 1.47 1.40 1.57 1.46 1.60 1.53 1.76 1.63 2.01 1.78
(unfiltered)

FAR Density (g/cm3) 1.60 1.77 1.72 1.60 1.42 1.35 1.51 1.41 1.63 1.52 1.86 1.73 2.05 1.83
Neutron Porosity (percent) 79 66 64 77 91 88 92 89 74 74 53 68 50 68
NEAR Neutron Raw (snu) 5293 5989 5557 5250 4849 4669 5059 4701 5271 5034 5937 5424 7196 6001
FAR Neutron Raw (snu) 204 247 224 209 164 160 168 159 199 194 304 236 344 257
PE (barns/electron) 1.36 1.55 1.56 1.24 1.06 1.12 1.16 0.88 1.23 1.10 1.31 1.17 1.91 1.40
Shallow Laterolog (ohm.m) 21 19 12 27 257 976 70 171 17 33 9 11 8 15
Deep Laterolog (ohm.m) 19 15 11 22 208 731 65 196 17 33 11 12 7 14
Sample Length (m) 0.53 0.81 0.80 0.42 0.80 0.80 0.63 0.70 0.55 0.42 0.32 0.33 0.73 0.79
Depth (m) 687.99 689.01 690.59 703.11 705.12 706.09 707.18 710.47 715.38 715.86 728.19 730.38 779.44 780.40
SEAM Argyle Auburn

242
Appendices

Well 14 - Samples 29 30 31 32 33 34 35 36 37 38 39 40 41 42
Gamma ray (API) 73 78 63 65 43 69 80 41 54 38 42 50 31 30
Compensated Density (g/cm3) 1.84 1.71 1.60 2.40 1.50 1.61 1.75 1.58 1.60 1.47 1.55 1.47 1.37 1.39
Mean wireline responses

Vetar Density (g/cm3) 1.78 1.64 1.59 2.39 1.46 1.57 1.74 1.55 1.61 1.46 1.57 1.45 1.37 1.36
NEAR Density (g/cm3) 1.80 1.72 1.65 2.34 1.57 1.62 1.66 1.59 1.63 1.53 1.62 1.51 1.45 1.48
(unfiltered)

FAR Density (g/cm3) 1.84 1.73 1.62 2.38 1.54 1.62 1.72 1.59 1.60 1.49 1.57 1.48 1.39 1.43
Neutron Porosity (percent) 66 69 80 24 76 77 70 77 82 82 80 78 94 83
NEAR Neutron Raw (snu) 6088 5419 5031 12466 5010 5178 4993 5648 5238 4973 5007 4747 4777 4994
FAR Neutron Raw (snu) 263 224 178 963 185 202 186 227 179 168 173 168 148 173
PE (barns/electron) 1.44 1.20 1.37 2.02 1.06 1.25 1.68 1.31 1.52 1.28 1.79 1.31 1.23 1.03
Shallow Laterolog (ohm.m) 14 15 28 12 40 21 3 28 19 52 95 199 1158 953
Deep Laterolog (ohm.m) 13 11 32 16 34 20 3 22 13 35 62 139 1042 998
Sample Length (m) 0.60 0.47 0.82 0.78 0.80 0.56 0.60 0.81 0.64 0.78 0.76 0.76 0.77 0.81
Depth (m) 784.52 786.15 792.76 796.79 802.28 812.75 813.50 833.78 834.50 835.21 835.98 836.74 837.51 838.30
SEAM Auburn Bulwer Condamine

Well 14 - Samples 43 44 45
Gamma ray (API) 87 37 54
Compensated Density (g/cm3) 2.03 1.40 1.50
Mean wireline responses

Vetar Density (g/cm3) 2.06 1.40 1.49


NEAR Density (g/cm3) 2.06 1.49 1.59
(unfiltered)

FAR Density (g/cm3) 2.04 1.43 1.54


Neutron Porosity (percent) 56 92 82
NEAR Neutron Raw (snu) 7602 4804 5076
FAR Neutron Raw (snu) 339 155 171
PE (barns/electron) 1.88 1.04 1.08
Shallow Laterolog (ohm.m) 15 946 118
Deep Laterolog (ohm.m) 12 716 83
Sample Length (m) 0.82 0.80 0.80
Depth (m) 845.01 845.82 846.62
SEAM Condamine

243
Appendices

APPENDIX E Permeability and coal bed characteristics

244
Appendices

APPENDIX F organic and inorganic content density estimation

The following equation was used to estimate density of organic content (o) and density of inorganic content (ash+moisture) given the

percentage of weigh of ash (&'RS ) and moisture (&T)KRLUH, ) from proximate analysis. The two following tables show the values used for a
and b respectively.
1#  ! ? &  &
'RS T)KRLUH,   F

Coal Seams Well 1 2 3 4 5 6 7 8 10 11 13 14 All Wells

Kogan -0.0051 -0.003 -0.0044 -0.0048 -0.0052 -0.0042 -0.0043

Macalister Upper -0.0031 -0.0041 -0.0047 -0.0043 -0.0045 -0.0046 -0.0049 -0.0041 -0.0048 -0.0043
Macalister Lower -0.0026 -0.0049 -0.0035 -0.0048 -0.0043 -0.0045 -0.0048 -0.0042 -0.0043

Nangram -0.0049 -0.004 -0.0046 -0.004 -0.004 -0.005 -0.0043


-0.005 -0.0044 -0.0051 -0.0048 -0.0049 -0.0048 -0.005 -0.0042 -0.0051 -0.0046 -0.0047
Coal Seams

Wambo
Iona -0.0049 -0.0047 -0.0045 -0.0046 -0.004 -0.0048 -0.0044 -0.0048 -0.0047 -0.0043 -0.0037 -0.0045

Argyle -0.0047 -0.0048 -0.0044 -0.0045 -0.0048 -0.0044 -0.0047 -0.0044 -0.0045 -0.0045 -0.0045 -0.0045

Tangalooma Ss. -0.0054 -0.0047 -0.0045 -0.0045


Auburn -0.0042 -0.0051 -0.0046 -0.005 -0.0044 -0.0044 -0.0047 -0.0046

Burwer -0.004 -0.0041 -0.0027 -0.005 -0.0048 -0.0043 -0.0042 -0.0046 -0.0047 -0.0034 -0.0042

Condamine -0.0044 -0.0048 -0.0045 -0.0049 -0.0046 -0.0043 -0.0042 -0.0048 -0.0045 -0.0046 -0.0043 -0.0046 -0.0044

Upper Juandah -0.0047 -0.0033


Measures
Coal

Lower Juandah -0.0047 -0.0046

Taroom -0.0046 -0.0045

All Samples -0.0047 -0.0044 -0.0045 -0.0047 -0.0046 -0.0045 -0.0046 -0.0046 -0.0045 -0.0044 -0.0044 -0.0043

245
Appendices

Coal Seams Well 1 2 3 4 5 6 7 8 10 11 13 14 All Wells


Kogan 0.8846 0.7836 0.8356 0.8383 0.8668 0.8338 0.8253

Macalister Upper 0.7548 0.8088 0.8159 0.8294 0.837 0.8403 0.8352 0.8277 0.8621 0.8219

Macalister Lower 0.7711 0.8326 0.8107 0.856 0.8249 0.8343 0.8336 0.8253 0.8255
Nangram 0.8418 0.8269 0.8345 0.822 0.8034 0.837 0.8267
0.8457 0.839 0.8743 0.8551 0.8619 0.8481 0.8466 0.8514 0.8646 0.8352 0.8462
Coal Seams

Wambo
Iona 0.8454 0.8568 0.8394 0.831 0.8192 0.8523 0.8262 0.8445 0.8541 0.8325 0.7943 0.8402
Argyle 0.8392 0.8563 0.8398 0.8375 0.8463 0.8341 0.8551 0.8311 0.8385 0.8372 0.8368 0.8379

Tangalooma Ss. 0.8633 0.8366 0.8389 0.8368

Auburn 0.8323 0.8672 0.8306 0.8447 0.8351 0.8251 0.8529 0.8373

Burwer 0.8218 0.8249 0.7648 0.8361 0.8443 0.8328 0.8167 0.8415 0.8432 0.8046 0.8236

Condamine 0.8253 0.8462 0.8452 0.8236 0.8309 0.8265 0.8075 0.8441 0.8209 0.8372 0.8164 0.8287 0.8255

Upper Juandah 0.8589 0.7904


Measures
Coal

Lower Juandah 0.8507 0.8462

Taroom 0.8383 0.8451

All Samples 0.8371 0.8394 0.8409 0.8267 0.8364 0.8347 0.8464 0.8331 0.8321 0.833 0.8293 0.8278

246
Appendices

APPENDIX G Error estimation


The following table refers to Table 5.2 where is shown the results of the estimation of core organic matter density in each coal seam unit in 12

wells. Here is show the sample count for each seam and in case of well 2 and 3 measure and mean absolute error of the estimate.

Well 1 2 3 4 5 6 7 8 10 11 13 14
n MAE n MAE n MAE n MAE n MAE n MAE n MAE n MAE n MAE n MAE n MAE n MAE

Kogan 2 0.004 2 0.048 2 0.004 6 0.034 2 0.008 3 0.017


Macalister
3 0.028 4 0.005 4 0.003 7 0.038 7 0.015 8 0.011 12 0.038 4 0.007 6 0.009
Upper
Macalister
2 0.002 5 0.013 2 0.005 13 0.133 4 0.012 11 0.013 5 0.005 3 0.010
Lower
Nangram 3 0.003 2 0.002 5 0.007 3 0.012 3 0.041 2 0.000
Coal Seams

Wambo 4 0.008 9 0.011 4 0.006 2 0.007 5 0.020 5 0.009 2 0.001 2 0.005 4 0.026 4 0.005

Iona 2 0.004 3 0.001 6 0.012 2 0.005 6 0.026 8 0.017 4 0.023 6 0.006 3 0.008 3 0.034 3 0.008
Argyle 7 0.010 6 0.008 7 0.020 4 0.003 9 0.032 9 0.015 5 0.016 7 0.007 8 0.008 10 0.024 14 0.010

Tang. Ss. 2 0.003 2 0.004 4 0.017

Auburn 2 0.000 4 0.012 7 0.018 5 0.007 2 0.003 4 0.005 6 0.050


Burwer 2 0.001 2 0.001 3 0.010 2 0.003 11 0.007 7 0.033 4 0.003 5 0.015 3 0.005 3 0.029

Condamine 5 0.004 12 0.010 8 0.015 7 0.012 6 0.008 6 0.005 3 0.003 2 0.002 5 0.010 7 0.023 12 0.015 11 0.015
Upper
5 0.023 6 0.024
Measures

Juandah
Coal

Lower
18 0.011 17 0.024
Juandah
Taroom 15 0.011 15 0.027

Well 22 0.012 38 0.018 38 0.027 33 0.020 56 0.036 41 0.028 51 0.060 35 0.028 27 0.014 45 0.028 57 0.033 46 0.031

247
Appendices

The following table refers to Table 5.2 where is shown the results of the estimation of fixed carbon to volatile matter ratio for each coal seam

unit in 12 wells. Here is shown the sample count for each seam and standard deviation.

Well 1 2 3 4 5 6 7 8 10 11 13 14
n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D.
Kogan 2 0.046 2 1.019 6 0.367 2 0.089 3 0.425
Macalister
3 0.237 4 0.050 4 0.938 7 0.386 7 0.244 8 0.303 12 0.152 4 0.430 6 0.392
Upper
Macalister
2 0.102 5 0.919 2 0.057 13 0.327 4 0.176 11 0.039 5 0.054 3 0.570
Lower
Nangram 3 0.802 2 0.044 5 0.188 3 0.102 3 0.379 2 0.111
Coal Seams

Wambo 4 0.129 9 0.100 4 0.333 2 0.151 5 0.135 5 0.081 2 0.138 2 0.475 4 0.408 4 0.445
Iona 2 0.076 3 0.115 6 0.097 2 0.777 6 0.113 8 0.169 4 0.306 6 0.051 3 0.457 3 0.292 3 0.339

Argyle 7 0.104 6 0.088 7 0.233 4 0.851 9 0.290 9 0.117 5 0.214 7 0.037 8 0.059 10 0.079 14 0.095

Tang. Ss. 2 0.814 2 0.139 4 0.436


Auburn 2 0.157 4 0.169 7 0.018 5 0.063 2 0.066 4 0.168 6 0.254

Burwer 2 0.125 2 0.059 3 0.041 2 0.842 11 0.092 7 0.297 4 0.061 5 0.337 3 0.469 3 0.136

Condamine 5 0.253 12 0.078 8 0.101 7 0.908 6 0.090 6 0.172 3 0.243 2 0.056 5 0.095 7 0.467 12 0.112 11 0.132

All coal seams 22 0.147 38 0.118 38 0.191 33 0.892 55 0.216 41 0.226 51 0.168 35 0.168 27 0.099 45 0.316 57 0.237 46 0.237

248
Appendices

The following table refers to Table 5.2 where is shown the results of the estimation of gas saturation for each coal seam unit in 12 wells. Here is

shown the sample count for each seam and standard deviation.

Well 2 3 4 5 6 7 8 10 11 13 14
n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D. n S.D.
Kogan 2 12% 2 17% 2 5% 6 19% 2 21% 3 41%
Macalister
3 10% 4 19% 4 55% 5 18% 7 12% 8 80% 12 11% 4 64% 6 15%
Upper
Macalister
2 23% 5 19% 2 20% 2 2% 13 74% 4 12% 11 11% 5 17%
Lower
Nangram 3 16% 1 2 17% 1 5 11% 3 5% 1 3 30% 2 24%
Coal Seams

Wambo 9 28% 4 3% 1 1 5 255% 5 25% 2 5% 2 230% 4 6% 4 21%


Iona 3 36% 6 6% 2 40% 5 16% 1 8 199% 4 8% 6 21% 3 46% 3 10% 3 9%

Argyle 6 6% 7 19% 4 30% 7 42% 9 18% 5 22% 2 7% 7 40% 8 13% 10 13% 14 23%

Tang. Ss. 2 2% 1 1 4 17%


Auburn 1 4 7% 7 17% 5 16% 1 2 12% 4 9% 6 10%

Burwer 2 62% 3 11% 2 24% 11 13% 1 33% 7 65% 4 18% 5 116% 1 3 6%

Condamine 12 34% 8 11% 7 27% 6 11% 7 12% 3 20% 2 8% 5 19% 8 34% 12 16% 10 14%

249
Appendices

APPENDIX H density and gas content

equations

The following equation was used to estimate density of ash (ash) given the

percentage of weigh of ash (&'RS ) from proximate analysis. The following table show

the values used for a and b.

1#  ! ? &   F
'RS

Where a and b are:

Well 2 3
Coal seam a b a b
Kogan
Macalister Upper -0.0051 0.8294 -0.0039 0.7762
Macalister Lower -0.0028 0.7559
Nangram
Coal Seams

Wambo -0.0042 0.7936 -0.0048 0.8293


Iona -0.0044 0.8049 -0.0043 0.8036
Argyle -0.0045 0.8104 -0.0043 0.8072
Tangalooma Ss.
Auburn -0.0045 -0.0048 0.8265
Burwer -0.0045 0.8054 -0.0031 0.7629
Condamine -0.0049 0.8155 -0.0046 0.8219
Upper Juandah -0.0048 0.8145 -0.0034 0.7679
Measure
Coal

Lower Juandah -0.0044 0.8019 -0.0044 0.8100


s

Taroom -0.0047 0.8118 -0.0045 0.8189

250
Appendices

Estimation of density as a function of ash, organic and moisture used in Rogerss

methodology.

&  ! ?    F

Where a and b are

Ash Organic Moisture


a b a b a b
Well 2 89.761 -105.34 -92.691 201.54 2.935 3.68
Well 3 71.097 -74.475 -68.238 163.70 -2.802 10.7

Equations used to estimate gas content in using Calvert et al. (2011) and created

methodologies (W,WP, V and VP).

Calvert et al. (2011)

-  ! ? D1# E  F

Where a and b are:

Upper Lower
Taroom
Juandah Juandah
a b a b a b
Well 2 5.035 -2.237 15.175 -6.905 28.601 -15.713
Well 3 8.757 -3.075 14.875 -5.513 11.754 -3.782

251
Appendices

Method W

-  ! ? D1# E  F

Where a and b are:

Well 2 3
a b a b
Kogan 5.034 -2.237 8.755 -3.075
Macalister Upper 5.034 -2.237 8.755 -3.075
Macalister Lower 5.034 -2.237 8.755 -3.075
Juandah Coal
Nangram 5.034 -2.237 8.755 -3.075
Measures
Wambo 18.087 -8.484 14.875 -5.513
Iona 7.786 -2.479 14.875 -5.513
Argyle 7.786 -2.479 14.875 -5.513
Tangalooma Sandstone 63.253 -39.904 11.754 -3.782
Auburn 63.253 -39.904 11.754 -3.782
Taroom Coal
Burwer 63.253 -39.904 11.754 -3.782
Measures
Condamine 17.613 -8.131 11.754 -3.782

Method WP

Juandah coal measures:

-  ! ? ND1# E ? MO  F

Taroom coal measure & Tangalooma Sandstone coal seam:

-  ! ? P1# ? MQ  F

Where a and b are:

Well 2 3
a b a b
Kogan 2.989 -2.374 2.333 -2.766
Macalister Upper 2.989 --2.374 2.333 -2.766
Macalister Lower 2.989 -2.374 2.333 -2.766
Juandah Coal
Nangram 2.989 -2.374 2.333 -2.766
Measures
Wambo 6.638 -7.563 3.029 -4.573
Iona 1.665 -1.015 3.029 -4.573
Argyle 1.665 -1.015 3.029 -4.573
Tangalooma Sandstone 63.253 485.13 -72.683 92.926
Auburn 485.13 -72.683 92.926 -6.522
Taroom Coal
Bulwer 485.13 -72.683 92.926 -6.522
Measures
Condamine 81.184 -6.987 92.926 -6.522

252
Appendices

Method V

-  ! ? V1# G WF
)HI' JK- -)JL,JL

Where a and b are:

Well 2 3
a b a b
Kogan 0.021 0.154 0.349 1.258
Macalister Upper 0.021 0.154 0.349 1.258
Macalister Lower 0.021 0.154 0.349 1.258
Juandah Coal
Nangram 0.021 0.154 0.349 1.258
Measures
Wambo 0.009 1.977 0.079 0.023
Iona 0.022 1.836 0.079 0.023
Argyle 0.022 1.836 0.079 0.023
Tangalooma Sandstone 0.001 3.735 0.052 1.876
Auburn 0.001 3.735 0.052 1.876
Taroom Coal
Bulwer 0.001 3.735 0.052 1.876
Measures
Condamine 0.188 -9.966 0.052 1.876

Method VP

Juandah coal measures:

-  ! ? XV1# G W ? MY  F
)HI' JK- -)JL,JL

Taroom coal measure & Tangalooma Sandstone coal seam:

-  ! ? X1# G ? MY  F
)HI' JK- -)JL,JL

Where a and b are:

Well 2 3
a b a b
Kogan 0.012 0.146 0.011 1.083
Macalister Upper 0.012 0.146 0.011 1.083
Macalister Lower 0.012 0.146 0.011 1.083
Juandah Coal
Nangram 0.012 0.146 0.011 1.083
Measures
Wambo 0.012 0.146 0.018 0.03
Iona 0.013 0.089 0.018 0.03
Argyle 0.013 0.089 0.018 0.03
Tangalooma Sandstone 0.016 0.089 0.018 0.03
Auburn 0.016 0.694 0.380 0.931
Taroom Coal
Bulwer 0.016 0.694 0.380 0.931
Measures
Condamine 0.908 -9.013 0.380 0.931

253
Appendices

Proximate analysis and as received gas content in v/v in wells 2 and 3.

*
GZRS  &'
'

*
G)HI' JK-  &)
)

G[)KRLUH,  1  GZRS  G)HI' JK- -)JL,JL

]!403 !66 ? -)


-\  #]!403 8/073

WELL 2
Ash 0rganic Moisture Gcv Ash 0rganic Moisture Gcv
Sample Sample
(v/v) (v/v) (v/v) (v/v) (v/v) (v/v) (v/v) (v/v)
1 38.88 47.13 13.99 0.72 25 14.47 78.35 7.18 3.33
2 17.00 66.34 16.65 1.29 26 4.33 91.76 3.91 3.66
3 40.74 41.76 17.50 0.82 27 12.08 83.94 3.98 6.38
4 28.22 59.21 12.57 1.03 28 7.25 87.54 5.21 8.43
5 39.67 49.72 10.61 1.44 29 8.87 86.31 4.82 7.02
6 15.08 78.25 6.67 2.36 30 19.48 74.99 5.53 4.85
7 39.03 55.27 5.70 2.58 31 14.50 80.72 4.77 4.79
8 13.95 80.90 5.15 2.99 32 8.86 87.11 4.03 5.75
9 14.66 80.16 5.18 1.98 33 4.45 92.93 2.62 4.09
10 7.11 86.94 5.95 3.24 34 9.27 87.78 2.95 5.23
11 9.66 85.36 4.98 2.64 35 20.42 73.39 6.18 3.83
12 21.56 72.24 6.20 2.70 36 12.70 82.55 4.76 5.11
13 12.07 81.46 6.47 2.77 37 12.35 83.16 4.48 7.05
14 27.40 63.59 9.01 2.56 38 13.32 81.75 4.93 3.98
15 17.69 75.95 6.36 3.05

16 7.94 86.70 5.35 3.70

17 1.76 92.72 5.52 3.10

18 15.47 78.94 5.59 3.27

19 8.49 87.03 4.48 4.43

20 7.74 87.96 4.31 4.05

21 33.04 60.78 6.17 3.08

22 14.53 80.45 5.03 4.36

23 26.34 70.35 3.31 3.50

24 17.57 77.51 4.92 3.96

254
Appendices

WELL 3
Ash 0rganic Moisture Gcv
Sample
(v/v) (v/v) (v/v) (v/v)
1 12.83 82.94 4.23 4.15

2 11.07 84.93 4.00 4.26

3 30.11 64.56 5.33 3.39


4 8.81 85.97 5.21 2.50

5 19.10 74.12 6.78 2.81

6 33.39 57.50 9.11 3.36

7 -- -- -- 0.41

8 27.89 66.89 5.22 4.21

9 81.73 12.08 6.19 0.80


10 30.46 65.10 4.44 4.57

11 49.56 44.72 5.72 2.87

12 32.51 63.25 4.23 4.95

13 19.02 77.47 3.51 5.55

14 39.74 53.65 6.61 4.08


15 12.87 82.49 4.64 6.87

16 20.75 73.22 6.03 6.09

17 4.75 90.96 4.29 7.59


18 9.47 87.02 3.51 6.81

19 40.88 51.99 7.13 4.80

20 70.50 23.53 5.97 1.59


21 13.11 82.01 4.89 5.69

22 29.57 63.09 7.33 5.63

23 20.95 75.15 3.90 7.12

24 54.65 37.76 7.59 4.19

25 23.49 73.87 2.65 6.56


26 42.71 51.79 5.50 4.02

27 25.47 69.84 4.69 6.68

28 10.33 85.34 4.33 7.18


29 34.48 56.95 8.57 4.89

30 37.88 51.16 10.96 4.55


31 24.95 67.96 7.09 4.85
32 17.68 77.57 4.76 6.20

33 13.46 82.20 4.35 7.01

34 11.12 83.19 5.69 7.08


35 14.80 77.35 7.85 5.84

36 27.22 66.51 6.27 5.13

37 12.13 84.98 2.89 4.65

38 9.02 87.90 3.08 5.39

39 17.15 78.59 4.26 6.56

255
Appendices

APPENDIX I tools and logs acronyms

Tool Name
AS Acoustic Scanner
ATV Acoustic Televiewer
CMI Compact Micro Imager
MAI Compact Array Induction
MCG Compact Gamma Ray
MDL Compact Dual Laterolog
MDN Compact Dual Neutron
MFE High-resolution shallow focused electric
MMR Micro Laterolog
MPD Compact density
MSS Compact Sonic Sonde

Log Name
CLDC Caliper log
DDLL Deep dual laterolog
DEN Compensated density log
DEPT Measured depth
DSLL Shallow dual laterolog
DT35 Compressional travel time log
DT35wf Compressional travel time log (filtered)
FEFE Shallow focused electric log
FEFEwf Shallow focused electric log (filtered)
GRGC Gamma ray log
GRGCwf Gamma ray log (filtered)
HDEN High-resolution density log
HDENwf High-resolution density log (filtered)
NPRS Neutron porosity
NPRSwf Neutron porosity (filtered)
PDPE PE log
PDPEwf PE log (filtered)
RILD Deep induction log
RILM Medium induction log
SPCG Spontaneous Potential log
TGC Total core desorbed gas

256
References

REFERENCES

AAPG, 2011. Unconventional Energy Resources: 2011 Review. Natural Resources


Research 20, 279328. doi:10.1007/s11053-011-9157-x
Ahlbrandt, T.S., 2002. Future Petroleum Energy Resources of the World. International
Geology Review 44, 10921104. doi:10.2747/0020-6814.44.12.1092
Ahmed, A.-J., Johnston, S., Boyer, C., Lambert, S.W., Bustos, O.A., Pashin, J.C.,
Wray, A., 2009. Coalbed methane: Clean energy for the world. Oilfield Review
413.
Ahmed, U., Johnston, D., Colson, L., 1991. An Advanced and Integrated Approach to
Coal Formation Evaluation. Presented at the SPE Annual Technical Conference
and Exhibition, 6-9 October, Dallas, Texas, Society of Petroleum Engineers, p.
16. doi:10.2118/22736-MS
Ammosov, I.I., Eremin, I.V., 1963. Fracturing in Coal. Israel Program for Scientific
Translations 112.
Anderson, J., Simpson, M., Basinski, P., Beaton, A., Bulat, C., Ray, S., Reinheimer, D.,
Schlachter, G., Colson, L., Olse, T., Zachariah, J., Khan, R., Low, N., Ryan, B.,
Schoderbek, D., 2003. Producing Natural Gas from Coal. Oilfield Review 24.
Arastoopour, H., Chen, S.-T., 1991. Sensitivity Analysis of key Reservoir Parameters in
Gas Reservoirs. Presented at the Sensitivity Analysis of key Reservoir
Parameters in Gas Reservoirs, Society of Petroleum Engineers, p. 333.
doi:10.2118/21515-MS
Australian Standard, AS 2856.1-2000. Coal petrography Part 1: Preparation of coal
samples for incident light microscopy, 2000. . Standards Association of
Australia 6.
Australian Standard, AS 2856.2-1998. Coal petrography Part 2: Maceral analysis.,
1998. . Standards Association of Australia 32.
Australian Standard, AS 2856.3-2000. Coal petrography Part 3: Methods for
microscopical determination of the reflectance of coal macerals, 2000. .
Standards Association of Australia 16.
Australian Standard, AS 2856-1986. Coal-maceral analysis, 1986. . Standards
Association of Australia 22.
Australian Standard, AS 2916-2007. Symbols for graphical representation of coal seams
and associated strata, 2007. . Standards Association of Australia 13.
Australian Standard, AS1038.3-2000. Methods for the Analysis and Testing of Coal and
Coke. Part 3: Proximate Analysis of Higher Rank, 2000. . Standards Association
of Australia 12.
Bachu, S., Michael, K., 2003. Possible controls of hydrogeological and stress regimes
on the producibility of coalbed methane in Upper CretaceousTertiary strata of
the Alberta basin, Canada. AAPG Bulletin 87, 17291754.
doi:10.1306/06030302015
Barton, C., Hickman, S., Morin, R., Zoback, M., Finkbeiner, T., Sass, J., Benoit, D.,
1997. Fracture Permeability and its Relationship to In-Situ Stress in the Dixie
Valley, Nevada, Geothermal Reservoir. USGS Staff -- Published Research 155.
Beamish, B.B., ODonnell, G., 1992. Microbalance applications to sorption testing of
coal. Presented at the Symp. Coalbed Methane Research and Development,
Dept. Earth Sci., James Cook Univ, Townsville, pp. 3141.

257
References

Bodden, W.R., Ehrlich, R., 1998. Permeability of coals and characteristics of desorption
tests: Implications for coalbed methane production. International Journal of Coal
Geology 35, 333347. doi:10.1016/S0166-5162(97)00039-6
Boyer, C.M., 2004. Understanding the Impact of Technology on Full Cycle CBM
Financial Viability. Presented at the Prodeedings of The Successful
Commercialisation of Global Coalbed and Coalmine Methane Projects, CWC
Associates, Ltd., London, UK.
Boyer, C.M., Popvich, G.S., Schraufnagel, R.A., 1986. The Rock Creek Methane From
Multiple Coal Seams Completion Project. Presented at the SPE Unconventional
Gas Technology Symposium, 18-21 May,, Society of Petroleum Engineers,
Louisville, Kentucky,US, p. 8. doi:10.2118/15259-MS
Bradshaw, M., Hall, L., Copeland, A., Hitchins, N., 2012. Australian Gas Resource
Assessment 2012 (Geoscience Australia and Bureau of Resources and Energy
Economics No. 74032). Department of Resources, Energy and Tourism.
Bradshaw, M., Yeung, M., 1990. The Jurassic Palaeogeography of Australia. Bureau of
Mineral Resources, Geology and Geophysics 60.
Brunauer, S., Deming, L.S., Deming, W.E., Teller, E., 1940. On a Theory of the van der
Waals Adsorption of Gases. J. Am. Chem. Soc. 62, 17231732.
doi:10.1021/ja01864a025
Bustin, A.M.M., Bustin, R.M., 2008. Coal reservoir saturation: Impact of temperature
and pressure. AAPG Bulletin 92, 7786. doi:10.1306/08270706133
Bustin, R.., Clarkson, C.., 1998. Geological controls on coalbed methane reservoir
capacity and gas content. International Journal of Coal Geology 38, 326.
doi:10.1016/S0166-5162(98)00030-5
Calvert, S., Percy, I., Pritchard, T., Morgan, N., Graham, J., Al-Ojeh, M., Maddren, J.,
Crosdale, P., 2011. Coal petrophysical properties for realistic coal gas reservoir
modelling. SPWLA 52nd Annual Symp., Colorado Springs, Colorado, May 14-
19. 16.
Cao, X., Mastalerz, M., Chappell, M.A., Miller, L.F., Li, Y., Mao, J., 2011. Chemical
structures of coal lithotypes before and after CO2 adsorption as investigated by
advanced solid-state 13C nuclear magnetic resonance spectroscopy.
International Journal of Coal Geology 88, 6774.
doi:10.1016/j.coal.2011.08.003
Carroll, R.E., Pashin, J.C., 2003. Relationship of sorption capacity to coal quality: CO2
sequestration potential of coalbed methane reservoirs in the Black Warrior
Basin, in: Proceedings,paper. Presented at the International Coalbed Methane
Symposium, Tuscaloosa, Alabama, USA, p. 11.
Cervik, J., 1967. Behavior of Coal-Gas Reservoirs. Presented at the SPE Eastern
Regional Meeting, 2-3 November, Society of Petroleum Engineers, Pittsburgh,
Pennsylvania, US, p. 10. doi:10.2118/1973-MS
Chalmers, G.R.L., Marc Bustin, R., 2007. On the effects of petrographic composition
on coalbed methane sorption. International Journal of Coal Geology 69, 288
304. doi:10.1016/j.coal.2006.06.002
Choi, C.Y., Dyrkacz, G.R., Stock, L.M., 1987. Density separation of alkylated coal
macerals. Energy Fuels 1, 280286. doi:10.1021/ef00003a010
Clarkson, C., Rahmanian, M., Kantzas, A., Morad, K., 2011. Relative Permeability of
CBM Reservoirs: Controls on Curve Shape. International Journal of Coal
Geology Volume 88, 204217. doi:10.2118/137404-MS
Clarkson, C.R., Bustin, R.M., 1996. Variation in micropore capacity and size
distribution with composition in bituminous coal of the Western Canadian

258
References

Sedimentary Basin: Implications for coalbed methane potential. Fuel 75, 1483
1498. doi:10.1016/0016-2361(96)00142-1
Clarkson, C.R., Bustin, R.M., 1997. Variation in permeability with lithotype and
maceral composition of Cretaceous coals of the Canadian Cordillera.
International Journal of Coal Geology 33, 135151. doi:10.1016/S0166-
5162(96)00023-7
Clarkson, C.R., Bustin, R.M., 1999. The effect of pore structure and gas pressure upon
the transport properties of coal: a laboratory and modeling study. 1. Isotherms
and pore volume distributions. Fuel 78, 13331344. doi:10.1016/S0016-
2361(99)00055-1
Clayton, J.., 1998. Geochemistry of coalbed gas A review. International Journal of
Coal Geology 35, 159173. doi:10.1016/S0166-5162(97)00017-7
Close, J.C., 1993. Natural Fractures in Coal, in: Hydrocarbons from Coal. American
Association of Petroleum Geologists, Tulsa, Oklahoma, US, pp. 119132.
Close, J.C., Mavor, M.J., 1991. Influence of Coal Composition and Rank on Fracture
Development in Fruitland Coal Gas Reservoirs of San Juan Basin. SG 38:
Hydrocarbons from Coal, AAPG Special Volumes 119 132.
Cody, G.D., Larsen, J.W., Siskin, M., 1988. Anisotropic solvent swelling of coals.
Energy Fuels 2, 340344. doi:10.1021/ef00009a020
Cohen, A.D., Spackman, W., 1972. Methods in Peat Petrology and their Application to
Reconstruction of Paleoenvironments. Geological Society of America Bulletin
83, 129142. doi:10.1130/0016-7606(1972)83[129:MIPPAT]2.0.CO;2
Colson, J.L., 1991. Evaluating Gas Content of Black Warrior Basin Coalbeds From
Wireline Log Data. Society of Petroleum Engineers 3, 15 22.
doi:10.2118/21489-PA
Creedy, D.P., 1979. A study of variations in gas content of coal seams in relation to
petrographic and stratigraphic variations. University of Wales.
Creedy, D.P., Garner, K., Holloway, S., Wardell, A., Ren, T.X., 2001. A review of the
worldwide status of coalbed methane extraction and utilisation. Crown
Copyright 1925.
Crosdale, P.J., 1995. Lithotype sequences in the Early Miocene Maryville Coal
Measures, New Zealand. International Journal of Coal Geology 28, 3750.
doi:10.1016/0166-5162(95)00003-V
Crosdale, P.J., 2004. Coal facies studies in Australia. International Journal of Coal
Geology 58, 125130. doi:10.1016/j.coal.2003.10.004
Crosdale, P.J., Beamish, B., 1993. Maceral effects on methane sorption. Presented at
the New Developments in Coal Geology: A Symposium., Coal Geol. Group,
Geol. Soc. of Aust, Brisbane, pp. 9598.
Crosdale, P.J., Beamish, B., Valix, M., 1998. Coalbed methane sorption related to coal
composition. International Journal of Coal Geology 35, 147158.
doi:10.1016/S0166-5162(97)00015-3
Crosdale, P.J., Moore, T.A., Mares, T.E., 2008. Influence of moisture content and
temperature on methane adsorption isotherm analysis for coals from a low-rank,
biogenically-sourced gas reservoir. International Journal of Coal Geology 76,
166174. doi:10.1016/j.coal.2008.04.004
Cui, X., Bustin, R.M., Dipple, G., 2004. Selective transport of CO2, CH4, and N2 in
coals: insights from modeling of experimental gas adsorption data. Fuel 83,
293303. doi:10.1016/j.fuel.2003.09.001
Curl, S.J., 1978. Methane production in coal mines (No. ICTIS/TR 04). London,
International Energy Agency of Coal Research.

259
References

Dake, L.P., 1983. Fundamentals of Reservoir Engineering. Elsevier.


Daniels, E.J., Altaner, S.P., 1990. Clay Mineral Authigenesis in Coal and Shale from
the Anthracite Region, Pennsylvania. American Mineralogist; (USA) 75:7-8,
825839.
Daniels, E.J., Marshak, S., Altaner, S.P., 1996. Use of clay-mineral alteration patterns
to define syntectonic permeability of joints (cleat) in Pennsylvania anthracite
coal. Tectonophysics 263, 123136. doi:10.1016/S0040-1951(96)00019-4
Das, B.M., Nikols, D.J., Das, Z.U., Hucka, V.J., 1991. Factors Affecting Rate and Total
Volume of Methane Desorption from Coalbeds. Rocky Mountain Association of
Geologists 69.
Dawson, G.K.W., Esterle, J.S., 2010. Controls on coal cleat spacing. International
Journal of Coal Geology 82, 213218. doi:10.1016/j.coal.2009.10.004
Dhir, R., Dern Jr., R.R., Mavor, M.J., 1991. Economic and Reserve Evaluation of
Coalbed Methane Reservoirs. Journal of Petroleum Technology 43, 1424 1518.
doi:10.2118/22024-PA
Diamond, W.P., Schatzel, S.J., 1998. Measuring the gas content of coal: A review.
International Journal of Coal Geology 35, 311331. doi:10.1016/S0166-
5162(97)00040-2
Diessel, C.F.K., 1992. Coal-bearing depositional systems. Springer-Verlag.
Dron, R.W., 1925. Notes on cleat in the Scottish coalfield. Trans. Inst. Min. Eng. 70,
115117.
Dyrkacz, G.R., Bloomquist, C.A.A., 1992. Use of continuous flow centrifugation
techniques for coal maceral separation. 1. Fundamentals. Energy Fuels 6, 357
374. doi:10.1021/ef00034a005
Dyrkacz, G.R., Horwitz, E.P., 1982. Separation of coal macerals. Fuel 61, 312.
doi:10.1016/0016-2361(82)90285-X
Earlougher, R.C., 1977. Advances in Well Test Analysis. Henry L. Doherty Memorial
Fund of AIME.
Eddy, G., Rightmire, C., Byrer, C., 1982. Relationship of Methane Content of Coal
Rank and Depth: Theoretical vs. Observed. Presented at the SPE
Unconventional Gas Recovery Symposium, 16-18 May, Pittsburgh,
Pennsylvania, Society of Petroleum Engineers, p. 6. doi:10.2118/10800-MS
Ehrenberg, S.N., Nadeau, P.H., 2005. Sandstone vs. carbonate petroleum reservoirs: A
global perspective on porosity-depth and porosity-permeability relationships.
AAPG Bulletin 89, 435445. doi:10.1306/11230404071
Enever, J., Hennig, A., 1997. The Relationship between Permeability and Effective
Stress for Australian Coals and Its Implication with Respect to Coalbed
Methane Exploration and Reservoir Modelling. Presented at the International
Coalbed Methane Symposium Proceedings, The University of Alabama,
Tuscaloosa, USA, pp. 1322.
Ettinger, I., Eremin, B., Zimakov, B., Yanovskaya, M., 1966. Natural factors
influencing coal sorption properties I: Petrography and the sorption properties
of coals. Fuel 267275.
Exon, N.F. (Neville F., 1976. Geology of the Surat Basin in Queensland, Bulletin
(Australia. Bureau of Mineral Resources, Geology and Geophysics); 166.
Australian Govt. Pub. Service.
Faiz, M., Aziz, N.I., Hutton, A.C., Jones, B.G., 1992. Porosity and gas sorption capacity
of some eastern Australian coals in relation to coal rank and composition. Proc.
Symp. Coalbed Methane Res. Dev. Australia 920.

260
References

Faiz, M., Hendry, P., 2006. Significance of microbial activity in Australian coal bed
methane reservoirs a review. Bulletin of Canadian Petroleum Geology 54,
261272. doi:10.2113/gscpgbull.54.3.261
Faiz, M., Saghafi, A., Sherwood, N., Wang, I., 2007. The influence of petrological
properties and burial history on coal seam methane reservoir characterisation,
Sydney Basin, Australia. International Journal of Coal Geology 70, 193208.
doi:10.1016/j.coal.2006.02.012
Fielding, C.R., 1996. Mesozoic sedimentary basins and resources in eastern Australia; a
review of current understanding. In: Mesozoic 96. Proceedings of the Mesozoic
geology of the Eastern Australia Plate Conference. Geological Society of
Australia.
Gamson, P.D., Beamish, B.B., 1991. Characterisation of coal microstructure using
scanning electron microscopy. Presented at the AusIMM Queensland Coal
Symposium, Brisbane, Australia, pp. 921.
Gamson, P.D., Beamish, B.B., Johnson, D.P., 1993. Coal microstructure and
micropermeability and their effects on natural gas recovery. Fuel 72, 8799.
doi:10.1016/0016-2361(93)90381-B
Gamson, P.D., Beamish, B.B., Johnson, D.P., 1996. Coal microstructure and secondary
mineralization: their effect on methane recovery. Geological Society, London,
Special Publications 109, 165179. doi:10.1144/GSL.SP.1996.109.01.12
Gan, H., Nandi, S.P., Walker Jr, P.L., 1972. Nature of the porosity in American coals.
Fuel 51, 272277. doi:10.1016/0016-2361(72)90003-8
Gash, B., Volz, R., Potter, G., Corgan, J., 1992. The effects of cleat orientation and
confining pressure on cleat porosity, permeability, and relative permeability in
coal. Presented at the SCA Conference, p. 14.
Gaurav, K., Akbar Ali, A.H., Saada, T., Kumar, S., 2012. Performance Analysis in Coal
Seam Gas. Presented at the SPETT 2012 Energy Conference and Exhibition, 11-
13 June, Port-of-Spain, Trinidad, Society of Petroleum Engineers, p. 15.
doi:10.2118/157696-MS
Gray, I., 1980. The Mechanism of and Energy Release Associated with Outbursts.
Presented at the Symposium on the Occurrence, Prediction, and Control of
Outbursts in Coal Mines, Australasian Inst. of Mining and Metallurgy, Southern
Queensland Branch,, Brisbane, pp. 11125.
Gray, I., 1987a. Reservoir Engineering in Coal Seams: Part 1-The Physical Process of
Gas Storage and Movement in Coal Seams. SPE Reservoir Engineering 2, 28
34.
Gray, I., 1987b. Reservoir Engineering in Coal Seams: Part 2 - Observations of Gas
Movement in Coal Seams. SPE Reservoir Engineering 2, 3540.
Green, T., West, T., 1985. Coal swelling in n-amines and n-alcohols. American
Chemical Society 488492.
Gregg, S.J., 1951. The surface chemistry of solids. Chapman & Hall.
Gregg, S.J., Sing, K.S.W., 1967. Adsorption, surface area, and porosity. Academic
Press, London; New York.
Gu, F., Chalaturnyk, R., 2010. Permeability and porosity models considering anisotropy
and discontinuity of coalbeds and application in coupled simulation. Journal of
Petroleum Science and Engineering 74, 113131.
doi:10.1016/j.petrol.2010.09.002
Guo, R., Mannhardt, K., Kantzas, A., 2007. Laboratory Investigation on the
Permeability of Coal During Primary and Enhanced Coalbed Methane

261
References

Production. Journal of Canadian Petroleum Technology 47, 6.


doi:10.2118/2007-042
Grdal, G., Yaln, M.N., 2001. Pore volume and surface area of the Carboniferous
coals from the Zonguldak basin (NW Turkey) and their variations with rank and
maceral composition. International Journal of Coal Geology 48, 133144.
doi:10.1016/S0166-5162(01)00051-9
Hamilton, S.K., Esterle, J.S., Golding, S.D., 2012. Geological interpretation of gas
content trends, Walloon Subgroup, eastern Surat Basin, Queensland, Australia.
International Journal of Coal Geology 101, 2135.
doi:10.1016/j.coal.2012.07.001
Hargraves, A.J., 1963. Instantaneous Outbursts of Coal and Gas. University of Sydney,
Sydney.
Harpalani, S., Chen, G., 1997. Influence of gas production induced volumetric strain on
permeability of coal. Geotech Geol Eng 15, 303325. doi:10.1007/BF00880711
Harpalani, S., Schraufnagel, A., 1990a. Measurement of parameters impacting methane
recovery from coal seams. International Journal of Mining and Geological
Engineering 8, 369384. doi:10.1007/BF00920648
Harpalani, S., Schraufnagel, A., 1990b. Shrinkage of coal matrix with release of gas and
its impact on permeability of coal. Fuel 69, 551556. doi:10.1016/0016-
2361(90)90137-F
Harris, L.A., Yust, C.R., 1981. The Ultrafine Structure of Coal Determined by Electron
Microscopy, in: Coal Structure, Advances in Chemistry. American Chemical
Society, pp. 321336.
Harris, L.A., Yust, C.S., 1976. Transmission electron microscope observations of
porosity in coal. Fuel 55, 233236. doi:10.1016/0016-2361(76)90094-6
Hatch, J.R., Gluskoter, H.J., Lindahl, P.C., 1976. Sphalerite in coals from the Illinois
Basin. Economic Geology 71, 613624. doi:10.2113/gsecongeo.71.3.613
Hawkins, J.M., Schraufnagel, R.A., Olszewski, A.J., 1992a. Estimating Coalbed Gas
Content and Sorption Isotherm Using Well Log Data. Presented at the SPE
Annual Technical Conference and Exhibition, 4-7 October, Washington, D.C.,
Society of Petroleum Engineers, p. 10. doi:10.2118/24905-MS
Hawkins, J.M., Schraufnagel, R.A., Olszewski, A.J., 1992b. Estimating Coalbed Gas
Content and Sorption Isotherm Using Well Log Data. Society of Petroleum
Engineers. doi:10.2118/24905-MS
Hemela, M.L., Chapman, G., Reichman, J., 1982. Mathematical Model for Simulation
of Seam Gas Drainage. Presented at the Symposium on Seam Gas Drainage
With Particular Reference to the Working Seam, Australasian Ins. of Mining
and Metallurgy, Southern Illawarra Branch, Wollongong, pp. 5359.
Hewitt, J.L., 1984. Geologic Overview, Coal, and Coalbed Methane Resources of the
Warrior Basin - Alabama and Mississippi. AAPG Special Volumes 138, 73
104.
Hilchie, D.W., 1982. Applied openhole log interpretation (for geologists and engineers).
Douglas W. Hilchie Inc.
Hildenbrand, A., Krooss, B.M., Busch, A., Gaschnitz, R., 2006. Evolution of methane
sorption capacity of coal seams as a function of burial history a case study
from the Campine Basin, NE Belgium. International Journal of Coal Geology
66, 179203. doi:10.1016/j.coal.2005.07.006
Hoffmann, K.L., Totterdell, J.M., Dixon, O., Simpson, G.A., Brakel, A.T., Wells, A.T.,
McKellar, J.L., 2009. Sequence stratigraphy of Jurassic strata in the lower Surat

262
References

Basin succession, Queensland. Australian Journal of Earth Sciences 56, 461


476.
Holdgate, G.R., Kershaw, A.P., Sluiter, I.R.K., 1995. Sequence stratigraphic analysis
and the origins of Tertiary brown coal lithotypes, Latrobe Valley, Gippsland
Basin, Australia. International Journal of Coal Geology 28, 249275.
doi:10.1016/0166-5162(95)00020-8
Hower, J.C., Esterle, J.S., Wild, G.D., Pollock, J.D., 1990. Perspectives on coal
lithotype analysis. Journal of Coal Quality 9 4852.
Hower, J.C., Wagner, N.J., 2012. Notes on the methods of the combined
maceral/microlithotype determination in coal. International Journal of Coal
Geology 95, 4753. doi:10.1016/j.coal.2012.02.011
Hoyer, D.L., 1991. Evaluation Of Coalbed Fracture Porosity From Dual Laterolog. The
Log Analyst 32, 15.
Hyndman, R.J., Koehler, A.B., 2006. Another look at measures of forecast accuracy.
International Journal of Forecasting 22, 679688.
doi:10.1016/j.ijforecast.2006.03.001
International Energy Agency, 2013. World energy outlook. 2013. International Energy
Agency, Paris OECD.
Jenkins, C.D., II, C.M.B., 2008. Coalbed- and Shale-Gas Reservoirs. Journal of
Petroleum Technology 60, 9299.
Johnston, D.J., Gales, R.H., Ahmed, U., 1991. A New Logging Method for Enhanced
Coal Grading. Presented at the Low Permeability Reservoirs Symposium, 15-17
April, Society of Petroleum Engineers, Denver, Colorado, US, p. 8.
doi:10.2118/21810-MS
Jones, A.H., Bell, G.J., Schraufnagel, R.A., 1988. A Review of the Physical and
Mechanical Properties of Coal with Implications for Coal-Bed Methane Well
Completion and Production. Rocky Mountain Association of Geologists 169
183.
Jones, E.J.P., Voytek, M.A., Warwick, P.D., Corum, M.D., Cohn, A., Bunnell, J.E.,
Clark, A.C., Orem, W.H., 2008. Bioassay for estimating the biogenic methane-
generating potential of coal samples. International Journal of Coal Geology 76,
138150. doi:10.1016/j.coal.2008.05.011
Jones, G.D., Patrick, R.B., 1981. Stratigraphy and coal exploration geology of the
northeast Surat Basin. International Journal of Coal Geology 153163.
Joubert, J.I., Grein, C.T., Bienstock, D., 1974. Effect of moisture on the methane
capacity of American coals. Fuel 53, 186191. doi:10.1016/0016-
2361(74)90009-X
Kabir, A., McCalmont, S., Street, T., Johnson, R., 2011. Reservoir Characterisation of
Surat Basin Coal Seams using Drill Stem Tests. Presented at the SPE Asia
Pacific Oil and Gas Conference and Exhibition, 20-22 September, Society of
Petroleum Engineers, Jakarta, Indonesia. doi:10.2118/147828-MS
Karacan, C.O., Mitchell, G.D., 2003. Behavior and effect of different coal
microlithotypes during gas transport for carbon dioxide sequestration into coal
seams. International Journal of Coal Geology 53, 201217. doi:10.1016/S0166-
5162(03)00030-2
Karacan, C.O., Okandan, E., 2001. Adsorption and gas transport in coal microstructure:
investigation and evaluation by quantitative X-ray CT imaging. Fuel 80, 509
520. doi:10.1016/S0016-2361(00)00112-5
Kempton, N.H., Peeters, M., 1977. Wireline Logging For Coal Exploration In Australia.
The Log Analyst 18, 6.

263
References

Kendall, P.F., Briggs, H., 1933. The Formation of Rock Joints and the Cleat of Coal.
Proc. R. Soc. Edinburgh.
Khavari-Khorasani, G., Michelsen, J.K., 1999. Coal Bed Gas Content and Gas
Undersaturation, in: Mastalerz, M., Glikson, M., Golding, S.D. (Eds.), Coalbed
Methane: Scientific, Environmental and Economic Evaluation. Springer
Netherlands, pp. 207231.
Kim, A.G., 1978. Experimental studies on the origin and accumulation of coalbed gas.
Dept. of the Interior, Bureau of Mines, US.
Kim, G., 1977. Estimating methane content of bituminous coal beds form adsorption
data. Dept. of the Interior, Bureau of Mines.
King, G., 1985. Numerical simulation of the simultaneous flow of methane and water
through dual porosity coal seams during the degasification process (PhD
Thesis). Pennsylvania State University, P.A., USA.
King, G., Ertekin, T., Schwerer, F., 1986. Numerical Simulation of the Transient
Behavior of Coal-Seam Degasification Wells. SPE Formation Evaluation 1, 165
183. doi:10.2118/12258-PA
Kissell, F.N., McCulloch, C.M., Elder, C.H., 1973. The Direct Method of Determining
Methane Content of Coalbeds for Ventilation Design. Bureau of Mines, US
NTIS No. PB-221628, RI 7767, 117.
Koenig, R.A., Stubbs, P.B., 1986. Interference Testing of a Coalbed Methane
Reservoir. Presented at the SPE Unconventional Gas Technology Symposium,
18-21 May, Society of Petroleum Engineers, Louisville, Kentucky, US, p. 12.
doi:10.2118/15225-MS
Kolesar, J.E., Ertekin, T., Obut, S.T., 1990a. The Unsteady-State Nature of Sorption
and Diffusion Phenomena in the Micropore Structure of Coal: Part 1 Theory
and Mathematical Formulation. SPE Formation Evaluation 5, 8188.
Kolesar, J.E., Ertekin, T., Obut, S.T., 1990b. The Unsteady-State Nature of Sorption
and Diffusion Phenomena in the Micropore Structure of Coal: Part 2 - Solution.
SPE Formation Evaluation 5, 8997.
Krevelen, D.W. van, 1981. Coal: Typology - chemistry - physics - constitution. Elsevier
Science Publishers, Amsterdam.
Krooss, B.., van Bergen, F., Gensterblum, Y., Siemons, N., Pagnier, H.J.., David, P.,
2002. High-pressure methane and carbon dioxide adsorption on dry and
moisture-equilibrated Pennsylvanian coals. International Journal of Coal
Geology 51, 6992. doi:10.1016/S0166-5162(02)00078-2
Kulander, B.R., Dean, S.L., 1993. Coal-Cleat Domains and Domain Boundaries in the
Allegheny Plateau of West Virginia. AAPG Bulletin 77, 13741388.
Kulander, B.R., Dean, S.L., Jr, B.J.W., 1990. Fractured Core Analysis: Interpretation,
Logging, and Use of Natural and Induced Fractures in Core. American
Association of Petroleum Geologists.
Lagendijk, E., Ryan, D., 2010. From CSG to LNG: Modeling and Understanding Key
Subsurface Uncertainties for the Development of a Surat Basin Opportunity, in
Queensland, Australia. Presented at the Canadian Unconventional Resources
and International Petroleum Conference, 19-21 October, Society of Petroleum
Engineers, Calgary, Alberta, Canada, p. 13. doi:10.2118/137651-MS
Lama, R.D., Bartosiewicz, H., 1982. Determination of Gas Content of Coal Seams.
Presented at the Symposium on Seam Gas Drainage with Particular Reference to
the Working Seam, Australasian Inst. of Mining and Metallurgy, Southern
Illawarra Branch, Wollongong, pp. 3649.

264
References

Lamberson, M.N., Bustin, R.M., 1993. Coalbed Methane Characteristics of Gates


Formation Coals, Northeastern British Columbia: Effect of Maceral
Composition. AAPG Bulletin 77, 20622076.
Lamberson, M.N., Bustin, R.M., Kalkreuth, W., 1991. Lithotype (maceral) composition
and variation as correlated with paleo-wetland environments, Gates Formation,
northeastern British Columbia, Canada. International Journal of Coal Geology
18, 87124. doi:10.1016/0166-5162(91)90045-K
Langenberg, C.W., Kalkreuth, W., Levine, J.R., Strobl, R., Demchunk, T., Hoffman,
G., Jerzykiewicz, T., 1990. Coal geology and its application to coal-bed
methane reservoirs: lecture notes for short course, Edmonton, August 20-24,
1990. Alberta Research Council.
Langmuir, I., 1918. THE ADSORPTION OF GASES ON PLANE SURFACES OF
GLASS, MICA AND PLATINUM. J. Am. Chem. Soc. 40, 13611403.
doi:10.1021/ja02242a004
Laubach, S.., Marrett, R.., Olson, J.., Scott, A.., 1998. Characteristics and origins of
coal cleat: A review. International Journal of Coal Geology 35, 175207.
doi:10.1016/S0166-5162(97)00012-8
Laubach, S.E., Tremain, C.M., 1991. Regional Coal Fracture Patterns and Coalbed
Methane Development. Presented at the The 32nd U.S. Symposium on Rock
Mechanics (USRMS), July 10 - 12, American Rock Mechanics Association,
Norman, Oklahoma,US, p. 10.
Law, D.H.S., Van Der Meer, L.G.H., Gunder, W.D., 2002. Numerical Simulator
Comparison Study for Enhanced Coalbed Methane Recovery Processes, Part I:
Pure Carbon Dioxide Injection, in: SPE Paper 75669. Presented at the SPE gas
Technology Symposium, Society of Petroleum Engineers, Calgary, Alberta,
Canada, p. 14.
Laxminarayana, C., Crosdale, P.J., 1999. Role of coal type and rank on methane
sorption characteristics of Bowen Basin, Australia coals. International Journal of
Coal Geology 40, 309325. doi:10.1016/S0166-5162(99)00005-1
Laxminarayana, C., Crosdale, P.J., 2002. Controls on methane sorption capacity of
Indian coals. AAPG Bulletin 86, 201212.
Leather, D.T.B., Bahadori, A., Nwaoha, C., Wood, D.A., 2013. A review of Australias
natural gas resources and their exploitation. Journal of Natural Gas Science and
Engineering 10, 6888. doi:10.1016/j.jngse.2012.09.003
Levine, J.R., 1992a. Influences of coal composition on coal seam reservoir quality: a
review, in: Symp. Coalbed Methane Res. Dev., Vol. 1. Townsville, p. 27.
Levine, J.R., 1992b. Oversimplifications can lead to faulty coalbed gas reservoir
analysis. The Oil and Gas Journal 90, 63+.
Levine, J.R., 1993. Coalification: The Evolution of Coal as Source Rock and Reservoir
Rock for Oil and Gas: Chapter 3. AAPG Special Volumes 180, 3977.
Levine, J.R., 1996. Model study of the influence of matrix shrinkage on absolute
permeability of coal bed reservoirs. Geological Society, London, Special
Publications 109, 197212. doi:10.1144/GSL.SP.1996.109.01.14
Levine, J.R., Clarkson, C.., Levy, J.H., 1995. , in: Proceedings of the 29th Newcastle
Symposium. Presented at the Advances in the Study of the Sydney Basin,
University of Newcastle, pp. 228.
Levine, J.R., Johnson, P., Beamish, B., 1993. . Proceedings of the 1993 International
Coalbed Methane Symposium, Tuscaloosa, pp. 18795.

265
References

Levy, J.H., Day, S.J., Killingley, J.S., 1997. Methane capacities of Bowen Basin coals
related to coal properties. Fuel 76, 813819. doi:10.1016/S0016-
2361(97)00078-1
Li, J., Liu, D., Yao, Y., Cai, Y., Qiu, Y., 2011. Evaluation of the reservoir permeability
of anthracite coals by geophysical logging data. International Journal of Coal
Geology 87, 121127. doi:10.1016/j.coal.2011.06.001
MacLennan, J.D., Schafer, P.S., Pratt, T.J., 1995. A Guide to Determining Coalbed Gas
Content. Gas Research Institute.
Mahajan, O.P., Walker, J., 1978. Porosity of Coals and Coal Products (No. FE-2030-
TR7). Pennsylvania State Univ., University Park (USA). Coal Research Section.
Malone, P.G., Briscoe, F.H., Camp, B.S., Boyer II, C.M., 1987. Discovery and
explanation of low gas contents encountered in coalbeds at the GRI/USSC Big
Indian Creek Site, Warrior Basin, Alabama. Presented at the 1987 Coalbed
Methane Symp., Univ. Alabama, Tuscaloosa, US, pp. 6372.
Mares, T.E., Moore, T.A., 2008. The influence of macroscopic texture on biogenically-
derived coalbed methane, Huntly coalfield, New Zealand. International Journal
of Coal Geology 76, 175185. doi:10.1016/j.coal.2008.05.013
Mares, T.E., Moore, T.A., Moore, C.R., 2009a. Uncertainty of gas saturation estimates
in a subbituminous coal seam. International Journal of Coal Geology 77, 320
327. doi:10.1016/j.coal.2008.07.002
Mares, T.E., Radliski, A.P., Moore, T.A., Cookson, D., Thiyagarajan, P., Ilavsky, J.,
Klepp, J., 2009b. Assessing the potential for CO2 adsorption in a subbituminous
coal, Huntly Coalfield, New Zealand, using small angle scattering techniques.
International Journal of Coal Geology 77, 5468.
doi:10.1016/j.coal.2008.07.007
Martin, M.A., Wakefield, M., MacPhail, M.K., Pearce, T., Edwards, H.E., 2013.
Sedimentology and stratigraphy of an intra-cratonic basin coal seam gas play:
Walloon Subgroup of the Surat Basin, eastern Australia. Petroleum Geoscience
19, 2138.
Mastalerz, M., Drobniak, A., Strpo, D., Solano Acosta, W., Rupp, J., 2008. Variations
in pore characteristics in high volatile bituminous coals: Implications for coal
bed gas content. International Journal of Coal Geology 76, 205216.
doi:10.1016/j.coal.2008.07.006
Mastalerz, M., Gluskoter, H., Rupp, J., 2004. Carbon dioxide and methane sorption in
high volatile bituminous coals from Indiana, USA. International Journal of Coal
Geology 60, 4355. doi:10.1016/j.coal.2004.04.001
Mathews, J.P., Hatcher, P.G., Scaroni, A.W., 1997. Particle size dependence of coal
volatile matter: is there a non-maceral-related effect? Fuel 76, 359362.
doi:10.1016/S0016-2361(96)00220-7
Mavor, M.J., 1994. Coal Gas Openhole Well Performance. Presented at the University
of Tulsa Centennial Petroleum Engineering Symposium, 29-31 August, Society
of Petroleum Engineers, Tulsa, Oklahoma,US, p. 14. doi:10.2118/27993-MS
Mavor, M.J., Close, J., McBane, R., 1990a. Formation Evaluation Of Exploration
Coalbed Methane Wells. Presented at the Annual Technical Meeting, June 10 -
13, Petroleum Society of Canada, Calgary, Alberta, Canada, p. 20.
doi:10.2118/90-101
Mavor, M.J., Close, J.C., McBane, R.A., 1994. Formation Evaluation of Exploration
Coalbed-Methane Wells. SPE Formation Evaluation 9, 285 294.
doi:10.2118/21589-PA

266
References

Mavor, M.J., Gunter, W., 2006. Secondary Porosity and Permeability of Coal vs. Gas
Composition and Pressure. SPE Reservoir Evaluation & Engineering 9, 114
125. doi:10.2118/90255-PA
Mavor, M.J., Nelson, C.R., 1997. Coalbed Reservoir Gas-in-place Analysis, 1st ed. Gas
Research Institute.
Mavor, M.J., Owen, L.B., Pratt, T.J., 1990b. Measurement and Evaluation of Coal
Sorption Isotherm Data. Presented at the SPE Annual Technical Conference and
Exhibition, 23-26 September, Society of Petroleum Engineers, New Orleans,
Louisiana, US, p. 14.
Mavor, M.J., Pratt, T., Nelson, C., Casey, T., 1996. Improved Gas-In-Place
Determination for Coal Gas Reservoirs. Presented at the SPE Gas Technology
Symposium, 28 April-1 May, Society of Petroleum Engineers, Calgary, Alberta,
Canada, p. 12. doi:10.2118/35623-MS
Mavor, M.J., Pratt, T.J., 1996. Improved Methodology for Determining Total Gas
Content. Volume 2. Comparative Evaluation of the Accuracy of Gas-in-Place
Estimates and Review of Lost Gas Models (No. Topical Report GRI-94/0429).
Gas Research Institute, Chicago, Illinois.
McCulloch, C.M., Levine, J.R., Kissell, F.N., Deul, M., 1975. Measuring the methane
content of bituminous coalbeds. Bureau of Mines, US.
Montgomery, S.L., 1999. Powder River basin, Wyoming; an expanding coalbed
methane (CBM) play. AAPG Bulletin 83, 12071222.
Moore, T.A., 2012. Coalbed methane: A review. International Journal of Coal Geology
101, 3681. doi:10.1016/j.coal.2012.05.011
Mullen, M.J., 1988. Log Evaluation in wells drilled for Coalbed Methane. Rocky
Mountain Association of Geologists 113124.
Mullen, M.J., 1989. Coalbed Methane Resource Evaluation From Wireline Logs in the
Northeastern San Juan Basin: A Case Study. Presented at the Low Permeability
Reservoirs Symposium, 6-8 March, Society of Petroleum Engineers, Denver,
Colorado, US. doi:10.2118/18946-MS
Nandi, S.P., Walker Jr, P.L., 1970. Activated diffusion of methane in coal. Fuel 49,
309323. doi:10.1016/0016-2361(70)90023-2
Neavel, R.C., Smith, S.E., Hippo, E.J., Miller, R.N., 1986. Interrelationships between
coal compositional parameters. Fuel 65, 312320. doi:10.1016/0016-
2361(86)90289-9
Nelson, C., 1999. Effects of Coalbed Reservoir Property Analysis Methods on Gas-In-
Place Estimates. Presented at the SPE Eastern Regional Conference and
Exhibition, 21-22 October, Society of Petroleum Engineers, Charleston, West
Virginia, US, p. 10. doi:10.2118/57443-MS
Nelson, C., 2000a. Effects of Geologic Variables on Cleat Porosity Trends in Coalbed
Gas Reservoirs, in: SPE/CERI Gas Technology Symposium, 3-5 April.
Presented at the Calgary, Alberta, Canada, Copyright 2000, Society of
Petroleum Engineers Inc., Calgary, Alberta, Canada, p. 5.
Nelson, C., 2000b. Effects of Geologic Variables on Cleat Porosity Trends in Coalbed
Gas Reservoirs. Society of Petroleum Engineers. doi:10.2118/59787-MS
Nelson, C., Hill, D., Pratt, T., 2000. Properties of Paleocene Fort Union Formation
Canyon Seam Coal at the Triton Federal Coalbed Methane Well, Campbell
County, Wyoming. Presented at the SPE/CERI Gas Technology Symposium, 3-
5 April, Society of Petroleum Engineers, Calgary, Alberta, Canada, p. 11.
doi:10.2118/59786-MS

267
References

Palmer, I., 2010. Coalbed methane completions: A world view. International Journal of
Coal Geology 82, 184195. doi:10.1016/j.coal.2009.12.010
Palmer, I., Mansoori, J., 1996. How Permeability Depends on Stress and Pore Pressure
in Coalbeds: A New Model. Presented at the SPE Annual Technical Conference
and Exhibition, 6-9 October, Society of Petroleum Engineers, Denver, Colorado,
US, p. 8. doi:10.2118/36737-MS
Papendick, S.L., Downs, K.R., Vo, K.D., Hamilton, S.K., Dawson, G.K.W., Golding,
S.D., Gilcrease, P.C., 2011. Biogenic methane potential for Surat Basin,
Queensland coal seams. International Journal of Coal Geology 88, 123134.
doi:10.1016/j.coal.2011.09.005
Parkash, S., Du Plessis, M.P., Cameron, A.R., 1983. Application of Coal Petrography in
the Liquefaction of Subbituminous Coals and Lignites. Alberta Research
Council.
Pashin, J.C., 1998. Stratigraphy and structure of coalbed methane reservoirs in the
United States: An overview. International Journal of Coal Geology 35, 209240.
doi:10.1016/S0166-5162(97)00013-X
Pashin, J.C., 2010. Variable gas saturation in coalbed methane reservoirs of the Black
Warrior Basin: Implications for exploration and production. International
Journal of Coal Geology 82, 135146. doi:10.1016/j.coal.2009.10.017
Pashin, J.C., Carroll, R.E., Richard H. Groshong, J., Raymond, D.E., McIntyre, M.,
Payton, J.W., 2003. Geologic Screening Criteria for sequestration of CO2 in
Coal: Quantifying Potential of the Black Warrior Coalbed Methane Fairway,
Alabama. U.S. Department of Energy National Energy Technology Laboratory.
Paterson, L., Meaney, K., Smyth, M., 1992. Measurements of relative permeability,
absolute permeability and fracture geometry in coal, in: CP 210. Presented at the
Coabed Methane Symposium, James Cook University of North Queensland,
Townsville, Australia, p. 7.
Patton S.B., 1996. Simulator for degasification, methane emission prediction and mine
ventilation. SME 37, 5. doi:10.1016/0140-6701(96)87454-6
Petersen, H.I., Bojesen-Koefoed, J.A., Nytoft, H.P., Surlyk, F., Therkelsen, J.,
Vosgerau, H., 1998. Relative sea-level changes recorded by paralic liptinite-
enriched coal facies cycles, Middle Jurassic Muslingebjerg Formation,
Hochstetter Forland, Northeast Greenland. International Journal of Coal
Geology 36, 130. doi:10.1016/S0166-5162(97)00032-3
Pratt, T., Mavor, M., DeBruyn, R., 1999. Coal Gas Resource and Production Potential
of Subbituminous Coal in the Powder River Basin. Presented at the SPE Rocky
Mountain Regional Meeting, 15-18 May, Society of Petroleum Engineers,
Gillette, Wyoming, US, p. 10. doi:10.2118/55599-MS
Prinz, D., Littke, R., 2005. Development of the micro- and ultramicroporous structure
of coals with rank as deduced from the accessibility to water. Fuel 84, 1645
1652. doi:10.1016/j.fuel.2005.01.010
Prinz, D., Pyckhout-Hintzen, W., Littke, R., 2004. Development of the meso- and
macroporous structure of coals with rank as analysed with small angle neutron
scattering and adsorption experiments. Fuel 83, 547556.
doi:10.1016/j.fuel.2003.09.006
Purl, R., Evanoff, J.C., Brugler, M.L., 1991. Measurement of Coal Cleat Porosity and
Relative Permeability Characteristics, in: SPE Gas Technology Symposium.
Presented at the SPE Gas Technology Symposium, 22-24 January, Society of
Petroleum Engineers Inc., Houston, Texas, US, p. 12.

268
References

QGC Pty Limited, 2012a. Surat Basin Geological Model (No. Appendix D).
Queensland, Australia.
QGC Pty Limited, 2012b. Vertical Connectivity and Geological Review (No. Appendix
O). Queensland, Australia.
QSQ, 2013. Queenslands coal seam gas overview (Industry update). Geological
Survey of Queensland.
Rao, P.D., Wolff, E.N., 1980. Petrographic, mineralogical, and chemical
characterization of certain Alaskan coals and washability products (No. Final
Report). University of Alaska Fairbanks. Mineral Industry Research Laboratory,
United States. Dept. of Energy, Alaska.
Reiss, L.H., 1980. The Reservoir Engineering Aspects of Fractured Formations.
Editions TECHNIP.
Ried, G.W., Towler, B.F., Harris, H.G., 1992. Simulation and Economics of Coalbed
Methane Production in the Powder River Basin. Presented at the PE Rocky
Mountain Regional Meeting, 18-21 May, Society of Petroleum Engineers,
Casper, Wyoming, US, p. 8. doi:10.2118/24360-MS
Rieke III, H., Rightmire, C., Fertl, W., 1981. Evaluation of Gas-Bearing Coal Seams.
Journal of Petroleum Technology 33, 195 204. doi:10.2118/8359-PA
Rightmire, C.T., 1984. Coalbed Methane Resources of the United States. AAPG
Special Volumes 138, 113.
Riley, J.T., 2007. Routine Coal and Coke Analysis: Collection, Interpretation, and Use
of Analytical Data. ASTM International.
Rodrigues, C.., Lemos de Sousa, M.., 2002. The measurement of coal porosity with
different gases. International Journal of Coal Geology 48, 245251.
doi:10.1016/S0166-5162(01)00061-1
Rogers, R.E., 2007. Coalbed methane: principles and practice. PTR Prentice Hall.
Ryan, B., 2003. Cleat Development in Some British Columbia Coals. New Ventures
Branch, Ministry of Energy, Mines and Petroleum Resources, Government of
British Columbia. 20.
Samworth, J.R., 1992. The Dual-spaced Density Log-characteristics, Calibration, And
Compensation. The Log Analyst. Society of Petrophysicists and Well-Log
Analysts 33, 8.
Samworth, J.R., Cherrie, M.A., 1976. A Focussed resistivity tool for slimline coal
logging systems. Society of Professional Well Log Analysts, European
Formation Evaluation Symposium, 4th, London, England 16.
Schopf, J.M., 1960. Field description and sampling of coal beds (No. 1111-B). United
States Geological Survey.
Schwarzer, R.R., Byrer, C., 1983. Variation in the quantity of methane adsorbed by
selected coals as a function of coal petrology and coal chemistry. U.S.
Department of Energy.
Scott, A.R., 2002. Hydrogeologic factors affecting gas content distribution in coal beds.
International Journal of Coal Geology 50, 363387. doi:10.1016/S0166-
5162(02)00135-0
Scott, A.R., Kaiser, W.R., Ayers, W.B.J., 1994. Thermogenic and Secondary Biogenic
Gases, San Juan Basin, Colorado and New Mexico - Implications for Coalbed
Gas Producibility. AAPG Bulletin 78, 11861209.
Scott, A.R., Zhou, N., Levine, J.R., 1995. A Modified Approach to Estimating Coal and
Coal Gas Resources: Example from the Sand Wash Basin, Colorado. AAPG
Bulletin 79, 13201336.

269
References

Scott, S., Anderson, B., Crosdale, P., Dingwall, J., 2007. Coal petrology and coal seam
gas contents of the Walloon Subgroup Surat Basin, Queensland, Australia.
International Journal of Coal Geology 70, 209222.
doi:10.1016/j.coal.2006.04.010
Seidle, J.P., 2011. Fundamentals of Coalbed Methane Reservoir Engineering, 1st ed.
PennWell Books.
Seidle, J.P., Jeansonne, M.W., Erickson, D.J., 1992. Application of Matchstick
Geometry To Stress Dependent Permeability in Coals. Presented at the SPE
Rocky Mountain Regional Meeting,18-21 May, Society of Petroleum Engineers,
Casper, Wyoming, US, p. 12. doi:10.2118/24361-MS
Sibbit, A.M., Faivre, O., 1985. The Dual Laterolog Response in Fractured Rocks.
Presented at the Society of Petrophysicists and Well-Log Analysts 26th Annual
Logging Symposium, 17-20 June, Society of Petrophysicists & Well Log
Analysts, Dallas, Texas, US, p. 34.
Smyth, M., Cook, A.C., 1976. Sequence in Australian coal seams. Mathematical
Geology 8, 529547. doi:10.1007/BF01042992
Sparks, D., McLendon, T., Saulsberry, J., Lambert, S., 1995. The Effects of Stress on
Coalbed Reservoir Performance, Black Warrior Basin, U.S.A. Presented at the
SPE Annual Technical Conference and Exhibition, 22-25 October, Society of
Petroleum Engineers, Dallas, Texas, US. doi:10.2118/30734-MS
Spears, D.A., Caswell, S.A., 1986. Mineral matter in coals: cleat minerals and their
origin in some coals from the english midlands. International Journal of Coal
Geology 6, 107125. doi:10.1016/0166-5162(86)90015-7
St. George, J.., Barakat, M.., 2001. The change in effective stress associated with
shrinkage from gas desorption in coal. International Journal of Coal Geology 45,
105113. doi:10.1016/S0166-5162(00)00026-4
Stach, E., Mackowsky, M.T.H., Taylor, D.G., Chandra, D.R., 1982. Stachs Textbook
of coal petrology. Borntraeger.
Stankiewicz, B.A., Kruge, M.A., Crelling, J.C., Salmon, G.L., 1994. Density Gradient
Centrifugation: Application to the Separation of Macerals of Type I, II, and III
Sedimentary Organic Matter. Energy Fuels 8, 15131521.
doi:10.1021/ef00048a042
Stopes, M.C., 1919. On the four visible ingredients in banded bituminous coals. Proc.
Royal Soc. 90B, 470487.
Stracher, G.B., Prakash, A., Sokol, E.V., 2010. Coal and Peat Fires: A Global
Perspective: Volume 1: Coal - Geology and Combustion. Elsevier.
Surez-Ruiz, I., Crelling, J.C., 2008. Applied Coal Petrology: The Role of Petrology in
Coal Utilization. Academic Press.
Sung, W., Ertekin, T., 1987. An Analysis of Field Development Strategies for Methane
Production From Coal Seams. Presented at the SPE Annual Technical
Conference and Exhibition, 27-30 September, Society of Petroleum Engineers,
Dallas, Texas, US, p. 11. doi:10.2118/16858-MS
Taulbee, D., Poe, S.H., Robl, T., Keogh, B., 1989. Density gradient centrifugation
separation and characterization of maceral groups from a mixed maceral
bituminous coal. Energy Fuels 3, 662670. doi:10.1021/ef00018a002
Taylor, G.H., Liu, S.Y., 1987. Biodegradation in coals and other organic-rich rocks.
Fuel 66, 12691273. doi:10.1016/0016-2361(87)90066-4
Taylor, G.H., Teichmller, M., Davis, A., Diessel, C.F.K., Littke, R., Robert, P., 1998.
Organic Petrology. Gebrder Borntraeger, Berlin.

270
References

Teichmller, M., Teichmller, R., 1968. Geological Aspects of Coal Metamorphism, D.


Murchison and T.S. Westoll. ed. Elsevier Publishing Co., New York.
Thimons, E.D., Kissell, F.N., 1973. Diffusion of methane through coal. Fuel 52, 274
280. doi:10.1016/0016-2361(73)90057-4
Thomas, L., 2002. Coal Geology. John Wiley & Sons.
Ting, F.T.C., 1977. Origin and Spacing of Cleats in Coal Beds. J. Pressure Vessel
Technol. 99, 624626. doi:10.1115/1.3454584
Tremain, C.M., Laubach, S.E., Iii, N.H.W., 1991. Coal Fracture (Cleat) Patterns in
Upper Cretaceous Fruitland Formation, San Juan Basin, Colorado and New
Mexico - Implications for Coalbed Methane Exploration and Development.
AAPG Bulletin 4964.
Unsworth, J.F., Fowler, C.S., Jones, L.F., 1989. Moisture in coal: 2. Maceral effects on
pore structure. Fuel 68, 1826. doi:10.1016/0016-2361(89)90005-7
Walker Jr., P.L., Verma, S.K., Rivera-Utrilla, J., Davis, A., 1988. Densities, porosities
and surface areas of coal macerals as measured by their interaction with gases,
vapours and liquids. Fuel 67, 16151623. doi:10.1016/0016-2361(88)90204-9
Wang, X., Ward, C., 2009. Experimental investigation of permeability changes with
pressure depletion in relation to coal quality. Presented at the International
Coalbed Methane Symposium. University of Alabama, Tuscaloosa, Alabama, p.
31.
Ward, C.R., 1984. Coal geology and coal technology. Blackwell Scientific Publications.
Weatherford - CMI, 2013. Compact Microimager (CMI) Tool [WWW Document].
URL http://www.weatherford.com/dn/WFT055879 (accessed 8.6.13).
Weatherford - MAI, 2012. Compact Array Induction (MAI) Tool [WWW Document].
URL
http://www.weatherford.com/weatherford/groups/web/documents/weatherfordco
rp/wft033170.pdf (accessed 8.6.13).
Weatherford - MCG, 2012. Compact Gamma Ray (MCG) Tool [WWW Document].
URL
http://www.weatherford.com/weatherford/groups/web/documents/weatherfordco
rp/WFT033097.pdf (accessed 7.12.13).
Weatherford - MDL, 2012. Compact Dual Laterolog (MDL) Tool [WWW Document].
URL http://www.weatherford.com/dn/WFT129361 (accessed 8.6.13).
Weatherford - MDN, 2012. Compact Dual Neutron (MDN) Tool [WWW Document].
URL http://www.weatherford.com/dn/WFT033180 (accessed 7.22.13).
Weatherford - MPD, 2010. Compact Photodensity (MPD) Tool [WWW Document].
URL
http://www.weatherford.com/weatherford/groups/web/documents/weatherfordco
rp/www018943.pdf (accessed 7.22.13).
Weatherford - MSS, 2012. Compact Sonic Sonde (MSS) Tool [WWW Document].
URL http://www.weatherford.com/dn/WFT033176 (accessed 7.22.13).
Weida, S., Lambert, S., Boyer, C., 2005. Challenging the Traditional Coalbed Methane
Exploration and Evaluation Model. Presented at the SPE Eastern Regional
Meeting, 14-16 September, Society of Petroleum Engineers, Morgantown, West
Virginia, US, p. 8. doi:10.2118/98069-MS
Weijermars, R., 2012. Jumps in Proved Unconventional Gas Reserves Present
Challenges to Reserves Auditing. SPE Economics & Management 4, pp. 131
146.
Whiticar, M.J., 1994. Correlation of Natural Gases with Their Sources: Chapter 16: Part
IV. Identification and Characterization. AAPG Special Volumes 77, 261283.

271
References

Willmott, C.J., Matsuura, K., 2005. Advantages of the mean absolute error (MAE) over
the root mean square error (RMSE) in assessing average model performance.
Clim Res 30, 7982. doi:10.3354/cr030079
Wold, M., Davidson, S., Wu, B., Choi, S., Koenig, R., 1995. Cavity Completion For
Coalbed Methane Stimulation - An Integrated Investigation And Trial In The
Bowen Basin, Queensland. Presented at the SPE Annual Technical Conference
and Exhibition, 22-25 October, Society of Petroleum Engineers, Dallas, Texas,
US, p. 15. doi:10.2118/30733-MS
Wold, M.B., Jeffrey, R.G., 1999. A Comparison of Coal Seam Directional Permeability
as Measured in Laboratory Core Tests and in Well Interference Tests, in: SPE
Rocky Mountain Regional Meeting, 15-18 May. Copyright 1999, Society of
Petroleum Engineers Inc., Gillette, Wyoming, US, p. 9.
Wyman, R.E., 1984. Gas Resources in Elmworth Coal Seams. AAPG Special Volumes
13, 173187.
Yalin, E., Durucan, ., 1991. Methane capacities of Zonguldak coals and the factors
affecting methane adsorption. Mining Science and Technology 13, 215222.
doi:10.1016/0167-9031(91)91346-J
Yang, R.T., Saunders, J.T., 1985. Adsorption of gases on coals and heattreated coals at
elevated temperature and pressure: 1. Adsorption from hydrogen and methane as
single gases. Fuel 64, 616620. doi:10.1016/0016-2361(85)90043-2
Yang, Y., Peeters, M., Kirk, C.W.V., Cloud, T.A., 2006. Gas Productivity Related to
Cleat Volumes Derived from Focused Resistivity Tools in Coalbed Methane
(CBM) Fields. PetroPhysics 47, 250257.
Yao, Y., Liu, D., 2009. Microscopic characteristics of microfractures in coals: an
investigation into permeability of coal. Procedia Earth and Planetary Science 1,
903910. doi:10.1016/j.proeps.2009.09.140
Yee, D., Seidle, J.P., Hanson, W.B., 1993. Gas Sorption on Coal and Measurement of
Gas Content, in: Law, B.E., Rice, D.D. (Eds.), Hydrocarbons from Coal.
American Association of Petroleum Geologists, Tulsa, Oklahoma. pp. 203218.
Young, G.B.C., McElhiney, J.E., Paul, G.W., McBane, R.A., 1991. An Analysis of
Fruitland Coalbed Methane Production, Cedar Hill Field, Northern San Juan
Basin. Presented at the SPE Annual Technical Conference and Exhibition, 6-9
October, Society of Petroleum Engineers, Dallas, Texas, US, p. 14.
doi:10.2118/22913-MS
Zhang, E., Hill, R.J., Katz, B.J., Tang, Y., 2008. Modeling of gas generation from the
Cameo coal zone in the Piceance Basin, Colorado. AAPG Bulletin 92, 1077
1106.
Zuber, M.D., Olszewski, A.J., 1992. The Impact of Errors in Measurements of Coalbed
Methane Reservoir Properties on Well Production Forecasts. Presented at the
SPE Annual Technical Conference and Exhibition, 4-7 October, Society of
Petroleum Engineers, Washington, D.C., US, p. 15. doi:10.2118/24908-MS.

272

You might also like