You are on page 1of 626

University of Wollongong

Research Online
University of Wollongong Thesis Collection

University of Wollongong Thesis Collections

2012

The relationships between petrology, porosity and


permeability in the southern Sydney Basin
succession, NSW
Fahad Mubarak Al gahtani

Recommended Citation
Al gahtani, Fahad Mubarak, The relationships between petrology, porosity and permeability in the southern Sydney Basin succession,
NSW, Doctor of Philosophy thesis, School of Earth and Environmental Sciences, University of Wollongong, 2012.
http://ro.uow.edu.au/theses/3614

Research Online is the open access institutional repository for the


University of Wollongong. For further information contact the UOW
Library: research-pubs@uow.edu.au

The relationships between petrology, porosity and permeability in the


southern Sydney Basin succession, NSW

A thesis submitted in fulfilment of the requirements for the award of the degree
of

Doctor of Philosophy

from

The University of Wollongong


New South Wales, Australia

By

Fahad Mubarak Al gahtani

School of Earth and Environmental Sciences


2012

The contents of this thesis are the results of original research by the author and
material contained herein has not been submitted to any other university or
similar institution for a higher degree

Fahad Mubarak Al gahtani

Abstract
Petrography of the Illawarra Coal Measures, Narrabeen Group and Hawkesbury
Sandstone was described by thin section, scanning electron microscope and Xray diffraction techniques. Sandstone composition of the Illawarra Coal
Measures and Narrabeen Group include mostly lithic grains with minor quartz,
feldspar, mica and heavy minerals. Quartz is dominant in the Hawkesbury
Sandstone which contains very minor feldspar, lithic grains, mica and heavy
minerals. The Illawarra Coal Measures consists of litharenite and rarely
sublitharenite, whereas the Narrabeen Group is litharenite and sublitharenite
with rare quartzarenite and the Hawkesbury Sandstone is quartzarenite to
sublitharenite. Quartz includes monocrystalline and polycrystalline grains while
the feldspar includes both K-feldspar and plagioclase. Volcanic, sedimentary
and chert rock fragments are present, particularly in the Illawarra Coal
Measures and Narrabeen Group.
Thin section porosity occurs in all units, particularly the coarse-grained deposits.
The Hawkesbury Sandstone has more visible porosity than the Narrabeen
Group or Illawarra Coal Measures. Primary porosity is more common than
secondary porosity in the Hawkesbury Sandstone whereas secondary porosity
is greater in the Narrabeen Group and Illawarra Coal Measures.
Thin sections and scanning electron microscopy were used to describe
diagenetic alterations and their influence on porosity in Illawarra Coal
Measures, Narrabeen Group and Hawkesbury Sandstone. These diagenetic
alterations include compaction, quartz overgrowths, authigenic clay minerals,
carbonate cement, authigenic feldspar, authigenic pyrite and iron oxide cement.
Compaction occurred in all units during throughout diagenesis. Mechanical
compaction reduced thin section porosity in the Illawarra Coal Measures and
Narrabeen Group whereas the influence of chemical compaction on thin section
porosity was greater in the Hawkesbury Sandstone. Both early and late
diagenetic carbonate cement is importantant in the Illawarra Coal Measures and
Narrabeen Group. Pore-filling carbonate cement reduced porosity whereas
dissolution of carbonate resulted in secondary porosity. Authigenic clay
i

minerals are the widespread in all units filling pores and occurring as graincoatings on detrital and authigenic grains. Where they coat quartz grains they
preserve porosity by preventing growth of quartz overgrowths. Quartz
overgrowths are only common in the Hawkesbury Sandstone where they have a
strong influence on porosity. Dissolution of unstable feldspar and lithic grains
provides secondary porosity in the Illawarra Coal Measures and Narrabeen
Group. This process is absent in the Hawkesbury Sandstone.
In the Illawarra Coal Measures, shale, siltstone and fine-grained sandstone are
common in the Wilton Formation, Bargo Claystone, Darkes Forest Sandstone,
Allans Creek Formation, Unnamed Member Three and Unnamed Member Two.
These units have low porosity and permeability and form lithological seals and
confining layers. Medium- and coarse-grained sandstone is common in the the
Kembla Sandstone, Lawrence Sandstone Member and Loddon Sandstone and
contains low porosity. Thus, gas or water may be present in these formations.
The Tongarra Coal, Allans Creek Coal Member, Wongawilli Coal, Hargrave
Coal Member, Cape Horn Coal Member, Balgownie Coal Member and Bulli
Coal are the main sources for gas in the Illawarra Coal Measures.
In the Narrabeen Group, shale and siltstone are abundant in the Wombarra
Claystone, Stanwell Park Claystone, Bald Hill Claystone and Newport
Formation and form lithological seals and confining layers. Minor sandstone
beds occur in the Wombarra Claystone and contain more porosity and
permeability. Medium- and coarse-grained sandstone is common in the Coalcliff
Sandstone, Scarborough Sandstone and Bulgo Sandstone and shows low to
moderate thin section porosity. Thus, these sandstone unis probably contain
gas or water.
The Hawkesbury Sandstone is characterized by primary and secondary
porosity, thus it has good groundwater storage and flow potential. Medium- and
coarse-grained sandstone beds are common in the Hawkesbury Sandstone but
it shows vertical variations in porosity. A few impermeable shale and siltstone
units occur in the Hawkesbury Sandstone forming local confining layers.

ii

Acknowledgements
This thesis was supervised by Associate Professor Brian Jones who helped me
to achieve this project through his discussions, opinions and reviews. Also, he
helped me to contact BHPBilliton Illawarra Coal to obtain the samples for this
study. I would like to thank him on all his efforts during the study period.
I would like to thank Dr. Adrian Hutton who was my co-supervisor during the
study period and helped me with many questions and suggestions.
I would like to thank BHPBilliton Illawarra Coal who provided my samples from
study area three years ago. Also, they supported me with well logs, maps and
additional information.
I would like to thank Mr Jose Abrantes for his help with thin sections, scanning
electron microscope photos and X-ray diffraction analyses.
I would like to thank the library staff in the University of Wollongong who helped
me in searching for information and borrowing scientific journals from other
universities.
I would like to thank my father and my mother for their patience with my
absence during the study period.
I would like to thank my wife who helped me during the past four years and
supported me to achieve my thesis.

iii

iv

Table of Contents
Abstract...i
Acknowledgment ....iii
List of figures.xvii
List of tablesxxi
List of appendices.xxi
Chapter One: Introduction
1-1Introduction.....................................................................................................1
1-2 Location of study area...................................................................................1
1-3 Stratigraphy of study area.............................................................................5
1-3-1 Illawarra Coal Measures.........................................................................7
1-3-1-1 Cumberland Subgroup.....................................................................7
1-3-1-2 Sydney Subgroup.............................................................................8
1-3-2 Narrabeen Group11
1-3-2-1 Source.............................................................................................11
1-3-2-2 Stratigraphy.....................................................................................12
1-3-2-2-1 Clifton Subgroup....................................................................12
1-3-2-2-2 Gosford Subgroup.................................................................13
1-3-3 Hawkesbury Sandstone........................................................................14
1-3-3-1 Thickness.......................................................................................14
1-3-3-2 Source............................................................................................14
1-4 Aims and Objectives....................................................................................15
Chapter Two: Literature review
2-1 Sandstone texture and composition............................................................17
2-2 Diagenesis...................................................................................................18
2-2-1 Compaction...........................................................................................18
2-2-2 Quartz cementation...............................................................................19
2-2-3 Carbonate cementation.........................................................................19
2-2-4 Authigenic Clay minerals.......................................................................21
2-2-5 Feldspars overgrowths..........................................................................22
2-3 Diagenesis and reservoir quality.................................................................23
v

2-3-1 Compaction...........................................................................................23
2-3-2 Quartz cementation...............................................................................24
2-3-3 Carbonate cementation.........................................................................24
2-3-4 Authigenic clay minerals........................................................................25
2-3-5 Dissolution.............................................................................................25
2-4 Grain size and reservoir quality...................................................................26
2-5 Detrital grains and reservoir quality.............................................................27
2-6 Depositional environment and reservoir quality...........................................27
2-7 Provenance..................................................................................................28
2-8 Conclusions.................................................................................................29
Chapter Three: Methodology
3-1 Introduction..................................................................................................31
3-2 Field Work....................................................................................................31
3-3 Laboratory work...........................................................................................32
3-3-1 Thin section...........................................................................................32
3-3-2 X-ray diffraction analysis.......................................................................34
3-3-3 Scanning electron microscope (SEM)...................................................34
3-4 Well logs......................................................................................................34
3-5 Porosity and Permeability measurements...................................................35
Chapter Four: Environment of deposition in the southern Sydney Basin
4-1 Introduction..................................................................................................39
4-2 Previous work..............................................................................................40
4-2-1 Illawarra Coal Measures........................................................................40
4-2-1-1 Cumberland Sub Group................................................................. 40
4-2-1-1-1 Pheasants Nest Formation....................................................40
4-2-1-1-2 Erins Vale Formation............................................................ 40
4-2-1-2 Sydney Sub-group......................................................................... 41
4-2-1-2-1 Wilton Formation................................................................... 41
4-2-1-2-2 Tongarra Coal....................................................................... 42
4-2-1-2-3 Bargo Claystone................................................................... 42
4-2-1-2-4 Darkes Forest Sandstone..................................................... 43
4-2-1-2-5 Allans Creek Formation..........................................................43
vi

4-2-1-2-6 American Creek Coal Member............................................. 44


4-2-1-2-7 Kembla Sandstone............................................................... 44
4-2-1-2-8 Wongawilli Coal.................................................................... 44
4-2-1-2-9 Lower Eckersley Formation.................................................. 45
4-2-1-2-10 Upper Eckersley Formation................................................ 46
4-2-1-2-11 Loddon Sandstone............................................................. 46
4-2-1-2-12 Bulli Coal............................................................................. 46
4-2-2 Narrabeen Group.................................................................................. 47
4-2-2-1 Clifton Sub-Group........................................................................... 47
4-2-2-1-1 Coalcliff Sandstone............................................................... 47
4-2-2-1-2 Wombarra Claystone............................................................ 48
4-2-2-1-3 Scarborough Sandstone....................................................... 48
4-2-2-1-4 Stanwell Park Claystone....................................................... 49
4-2-2-1-5 Bulgo Sandstone................................................................... 50
4-2-2-1-6 Bald Hill Claystone................................................................ 51
4-2-2-2 Gosford Sub Group........................................................................ 52
4-2-2-2-1 Garie Formation.................................................................... 52
4-2-2-2-2 Newport Formation............................................................... 52
4-2-3 Hawkesbury Sandstone....................................................................... 53
4-2-3-1 Lithofacies...................................................................................... 53
4-2-3-1-1 Massive sandstone facies..................................................... 53
4-2-3-1-2 Stratified sandstone facies.................................................... 54
4-2-3-1-3 Mudrock facies...................................................................... 55
4-2-3-2 Sedimentary structures.................................................................. 56
4-2-3-3 Depositional environment.............................................................. 56
4-3 Facies analysis........................................................................................... 57
4-3-1 Illawarra Coal Measures....................................................................... 57
4-3-1-1 Sydney Sub-group.......................................................................... 57
4-3-1-1-1 Wilton Formation................................................................... 57
4-3-1-1-2 Tongarra Coal....................................................................... 58
4-3-1-1-3 Bargo Claystone.................................................................... 58
4-3-1-1-4 Darkes Forest Sandstone..................................................... 60
vii

4-3-1-1-5 Allans Creek Formation........................................................ .61


4-3-1-1-6 American Creek Coal Member.............................................. 62
4-3-1-1-7 Kembla Sandstone................................................................ 63
4-3-1-1-8 Wongawilli Coal..................................................................... 64
4-3-1-1-9 Unnamed Member Three...................................................... 64
4-3-1-1-10 Unnamed Member Two....................................................... 65
4-3-1-1-11 Cape Horn Coal Member.................................................... 66
4-3-1-1-12 Laurence Sandstone........................................................... 67
4-3-1-1-13 Balgownie Coal Member..................................................... 67
4-3-1-1-14 Loddon Sandstone.............................................................. 68
4-3-1-1-15 Bulli Coal............................................................................. 69
4-3-2 Narrabeen Group................................................................................. 69
4-3-2-1 Clifton Sub-Group........................................................................... 70
4-3-2-1-1 Coalcliff Sandstone............................................................... 70
4-3-2-1-2 Wombarra Claystone............................................................ 72
4-3-2-1-3 Scarborough Sandstone....................................................... 73
4-3-2-1-4 Stanwell Park Claystone....................................................... 76
4-3-2-1-5 Bulgo Sandstone................................................................... 77
4-3-2-1-6 Bald Hill Claystone............................................................... 82
4-3-2-2 Gosford Sub-Group........................................................................ 83
4-3-2-2-1 Garie Formation.................................................................... 83
4-3-2-2-2 Newport Formation................................................................ 83
4-3-3 Hawkesbury Sandstone........................................................................ 84
4-4 The comparison between previous work and the present study................. 90
Chapter Five: Petrology of Southern Sydney Basin
5-1 Introduction....97
5-2 Description of Components.98
5-2-1 Quartz..98
5-2-1-1 Monocrystalline quartz...98
5-2-1-2 Polycrystalline quartz.99
5-2-2 Feldspar.101
5-2-2-1 K-feldspar...101
viii

5-2-2-2 Plagioclase102
5-2-3 Rock fragments102
5-2-3-1 Volcanic rock fragments..103
5-2-3-2 Sedimentary rock fragments104
5-2-4 Chert...104
5-2-5 Mica105
5-2-6 Heavy minerals.105
5-2-7 Clay minerals106
5-2-8 Matrix..107
5-3 Lateral and vertical variations of composition108
Illawarra Coal Measures108
5-3-1 Introduction108
5-3-2 Thin section...108
5-3-2-1 Detrital composition and texture.108
5-3-2-1-1 Quartz120
5-3-2-1-2 Feldspar122
5-3-2-1-3 Lithic grains..123
5-3-2-1-3-1 Volcanic Rock Fragment plus chert and carbonate
fragments......126
5-3-2-1-3-2 Sedimentary Rock Fragments126
5-3-2-1-3-3 Chert...126
5-3-2-1-4 Muscovite.128
5-3-2-1-5 Heavy minerals128
5-3-2-1-6 Chlorite.. .129
5-3-2-1-7 Matrix.......129
5-3-3 X-ray diffraction (XRD)... 130
5-3-3-1 Quartz..130
5-3-3-2 Feldspar..134
5-3-3-3 Clay minerals.134
5-3-3-4 Muscovite/illite135
5-3-3-5 Heavy minerals..136
5-3-3-6 Carbonate minerals...136
ix

Narrabeen Group...137
5-3-1 Introduction137
5-3-2 Thin section...138
5-3-2-1 Detrital Composition and Texture..138
5-3-2-1-1 Quartz..138
5-3-2-1-2 Feldspar...149
5-3-2-1-3 Lithic grains.149
5-3-2-1-3-1 Volcanic Rock Fragment plus chert and carbonate
fragments..151
5-3-2-1-3-2 Sedimentary Rock Fragments..151
5-3-2-1-3-3 Chert.152
5-3-2-1-4 Muscovite153
5-3-2-1-5 Heavy minerals..154
5-3-2-1-6 Chlorite155
5-3-2-1-7 Matrix155
5-3-3 X-ray diffraction (XRD)... 156
5-3-3-1 Quartz..156
5-3-3-2 Feldspar..162
5-3-3-3 Clay minerals.162
5-3-3-4 Muscovite/illite163
5-3-3-5 Heavy minerals..163
5-3-3-6 Carbonate Minerals..163
Hawkesbury Sandstone164
5-3-1 Introduction164
5-3-2 Thin Section..165
5-3-2-1 Detrital Composition and Texture...165
5-3-2-1-1 Quartz..166
5-3-2-1-2 Feldspar..167
5-3-2-1-3 Rock Fragments.167
5-3-2-1-4 Chert168
5-3-2-1-5 Muscovite168
5-3-2-1-6 Heavy minerals..168
x

5-3-2-1-7 Chlorite169
5-3-2-1-8 Matrix...169
5-3-3 X-ray diffraction analysis (XRD) 170
5-3-3-1 Quartz..170
5-3-3-2 Feldspar..171
5-3-3-3 Clay minerals.171
5-3-3-4 Muscovite/illite172
5-3-3-5 Heavy minerals..172
5-3-3-6 Carbonate minerals..172
Chapter Six: Diagenesis in the southern Sydney Basin
6-1 Introduction............................................................................................... 175
6-2 Illawarra Coal Measures........................................................................... 175
6-2-1 Scanning Electron Microscope (SEM)................................................ 175
6-2-2 Diagenesis.......................................................................................... 176
6-2-2-1 Diagenetic Minerals...................................................................... 176
6-2-2-1-1 Quartz cementation............................................................. 176
6-2-2-1-2 Authigenic clay minerals..................................................... 179
6-2-2-1-3 Carbonate cement............................................................... 180
6-2-2-1-4 Authigenic feldspar.............................................................. 182
6-2-2-1-5 Authigenic pyrite.................................................................. 182
6-2-2-1-6 Iron oxide cement................................................................ 182
6-2-3 Diagenetic Sequence......................................................................... 182
6-2-3-1 Compaction.................................................................................. 183
6-2-3-2 Cementation................................................................................. 183
6-2-3-2-1 Authigenic clay minerals..................................................... 183
6-2-3-2-2 Carbonate cementation....................................................... 185
6-2-3-2-3 Dissolution/alterations of unstable detrital grains................ 186
6-2-3-2-4 Authigenic quartz................................................................ 187
6-2-3-2-5 Late carbonate cementation................................................ 187
6-2-3-2-6 Late mixed layer illite/smectite and chlorite......................... 188
6-2-3-2-7 Ferruginous Cement........................................................... 188
6-2-3-2-8 Recrystallization.................................................................. 189
xi

6-3 Narrabeen Group...................................................................................... 190


6-3-1 Scanning Electron Microscope (SEM)................................................ 190
6-3-2 Diagenesis.......................................................................................... 190
6-3-2-1 Diagenetic Minerals...................................................................... 191
6-3-2-1-1 Quartz cement..................................................................... 191
6-3-2-1-2 Authigenic clay minerals..................................................... 192
6-3-2-1-3 Carbonate cement.............................................................. 193
6-3-2-1-4 Authigenic feldspars............................................................ 195
6-3-2-1-5 Authigenic pyrite.................................................................. 195
6-3-2-1-6 Iron oxide cement................................................................ 195
6-3-3 Diagenetic Sequence......................................................................... 195
6-3-3-1 Compaction.................................................................................. 196
6-3-3-2 Cementation................................................................................. 196
6-3-3-2-1 Authigenic clay minerals..................................................... 196
6-3-3-2-2 Carbonate cement.............................................................. 198
6-3-3-2-3 Dissolution/alterations of unstable detrital grains................ 199
6-3-3-2-4 Authigenic quartz................................................................ 199
6-3-3-2-5 Late carbonate cements...................................................... 200
6-3-3-2-6 Late authigenic clays............................................................200
6-3-3-2-7 Ferruginous Cement........................................................... 201
6 -4 Hawkesbury Sandstone........................................................................... 202
6-4-1 Scanning Electron Microscope (SEM)............................................... 202
6-4-2 Diagenesis........................................................................................ 203
6-4-2-1 Diagenetic Minerals..................................................................... 203
6-4-2-1-1 Quartz cementation............................................................ 203
6-4-2-1-2 Authigenic clay minerals.................................................... 205
6-4-2-1-3 Carbonate cement............................................................. 207
6-4-2-1-4 Authigenic Feldspars.......................................................... 208
6-4-2-1-5 Iron oxide cement............................................................... 209
6-4-3 Diagenetic Sequence........................................................................ 209
6-4-3-1 Compaction................................................................................. 209
6-4-3-2 Cementation............................................................................... 210
xii

6-4-3-2-1 Authigenic clay minerals.................................................... 210


6-4-3-2-2 Carbonate cement.............................................................. 211
6-4-3-2-3 Dissolution/alteration of detrital grains............................... 211
6-4-3-2-4 Authigenic quartz............................................................... 212
6-4-3-2-5 Late authigenic kaolinite, illite and chlorite......................... 213
6-4-3-2-6 Late siderite and ankerite cement...................................... 214
6-4-3-2-7 Ferruginous Cement.......................................................... 214
Chapter Seven: Porosity and permeability
7-1 Introduction............................................................................................... 217
7-2 Porosity and permeability data................................................................. 218
7-2-1 Illawarra Coal Measures.................................................................... 218
7-2-1-1 Thin section porosity................................................................ 218
7-2-1-2 Well log density porosity........................................................... 218
7-2-1-3 Helium porosity and permeability to air.................................... 223
7-2-1-4 Comparison between thin section porosity and density
porosity.................................................................................... 223
7-2-1-5 Comparison between thin section porosity and helium
porosity.................................................................................... 224
7-2-1-6 Comparison between density porosity and helium
porosity.................................................................................... 225
7-2-2 Narrabeen Group................................................................................ 225
7-2-2-1 Thin section porosity................................................................ 225
7-2-2-2 Well log density porosity........................................................... 231
7-2-2-3 Helium porosity and permeability to air.................................... 231
7-2-2-4 Comparison between thin section porosity and density
porosity.................................................................................... 231
7-2-2-5 Comparison between thin section porosity and helium
porosity.................................................................................... 232
7-2-2-6 The comparison between density porosity and helium
porosity.................................................................................... 233
7-2-3 Hawkesbury Sandstone...................................................................... 233
7-2-3-1 Thin section porosity................................................................ 233
xiii

7-2-3-2 Well log density porosity........................................................... 234


7-2-3-3 Helium porosity and permeability to air.................................... 237
7-2-3-4 Comparison between thin section porosity and density
porosity.................................................................................... 237
7-2-3-5 Comparison between thin section porosity and helium
porosity.................................................................................... 237
7-2-3-6 Comparison between density porosity and helium
porosity.................................................................................... 238
7-3 The interpretation of the difference between thin section porosity,
helium porosity and density porosity........................................................ 238
7-4 Correlation of helium porosity and air permeability with detrital and
authigenic grains...................................................................................... 239
7-5 Equivalent helium porosity........................................................................ 246
7-6 Distribution of density porosity in wells......................................................249
7-6-1 Illawarra Coal Measures..................................................................... 249
7-6-2 Narrabeen Group............................................................................... 249
7-6-3 Hawkesbury Sandstone...................................................................... 249
7-7 Pore types................................................................................................ 249
7-7-1 Illawarra Coal Measures..................................................................... 249
7-7-1-1 Primary porosity........................................................................... 249
7-7-1-2 Secondary porosity....................................................................... 250
7-7-2 Narrabeen Group................................................................................ 251
7-7-2-1 Primary porosity........................................................................... 251
7-7-2-2 Secondary porosity....................................................................... 251
7-7-3 Hawkesbury Sandstone...................................................................... 252
7-7-3-1 Primary porosity........................................................................... 252
7-7-3-2 Secondary porosity...................................................................... 252
7-8 Distribution of thin section porosity with facies......................................... 253
7-8-1 Illawarra Coal Measures..................................................................... 253
7-8-2 Narrabeen Group................................................................................ 255
7-8-3 Hawkesbury Sandstone...................................................................... 256
xiv

Chapter Eight: Discussion


8-1 Introduction.............................................................................................. 259
8-2 Linking depositional environment, provenance and composition
variation................................................................................................... 260
8-2-1 Illawarra Coal Measures and Narrabeen Group................................ 260
8-2-2 Hawkesbury Sandstone..................................................................... 277
8-3 Linking depositional environment, diagenetic and composition
variation................................................................................................... 281
8-3-1 Illawarra Coal Measures and Narrabeen Group................................. 281
8-3-2 Hawkesbury Sandstone...................................................................... 289
8-4 The influence of diagenetic alteration and composition on reservoir
Quality...................................................................................................... 292
8-4-1 Illawarra Coal Measures..................................................................... 292
8-4-1-1 Compaction................................................................................. 292
8-4-1-2 Carbonate cementation............................................................... 294
8-4-1-3 Authigenic clay minerals.............................................................. 296
8-4-1-4 Dissolution................................................................................... 299
8-4-1-5 Quartz overgrowths..................................................................... 301
8-4-2 Narrabeen Group............................................................................... 302
8-4-2-1 Compaction.................................................................................. 302
8-4-2-2 Carbonate cementation................................................................ 304
8-4-2-3 Authigenic clay minerals...............................................................305
8-4-2-4 Dissolution.................................................................................... 307
8-4-2-5 Quartz overgrowths...................................................................... 309
8-4-3 Hawkesbury Sandstone...................................................................... 310
8-4-3-1 Compaction.................................................................................. 310
8-4-3-2 Quartz overgrowths...................................................................... 312
8-4-3-3 Authigenic clay minerals............................................................... 314
8-4-3-4 Carbonate cementation................................................................ 316
8-4-3-5 Dissolution.....................................................................................316
8-5 The influence of compaction and cementation on porosity and
permeability.............................................................................................. 316
xv

8-5-1 Illawarra Coal Measures..................................................................... 317


8-5-2 Narrabeen Group................................................................................ 318
8-5-3 Hawkesbury Sandstone...................................................................... 318
Chapter Nine:
9-1 Reservoir Potential of the Southern Sydney Basin................................... 321
9-1-1 Illawarra Coal Measures..................................................................... 321
9-1-1-1 Wilton Formation.......................................................................... 321
9-1-1-2 Bargo Claystone........................................................................... 321
9-1-1-3 Darkes Forest Sandstone.............................................................322
9-1-1-4 Allans Creek Formation................................................................ 322
9-1-1-5 Kembla Sandstone....................................................................... 323
9-1-1-6 Eckersley Formation..................................................................... 325
9-1-1-6-1 Unnamed Member Three................................................... 325
9-1-1-6-2 Unnamed Member Two..................................................... 325
9-1-1-6-3 Lawrence Sandstone Member........................................... 326
9-1-1-7 Loddon Sandstone....................................................................... 327
9-1-1-8 Bulli Coal...................................................................................... 329
9-1-1-9 Summary...................................................................................... 329
9-1-1-10 Hydrocarbons in the Illawarra Coal Measures........................... 329
9-1-2 Narrabeen Group................................................................................ 331
9-1-2-1 Coalcliff Sandstone...................................................................... 331
9-1-2-2 Wombarra Claystone.................................................................... 333
9-1-2-3 Scarborough Sandstone............................................................... 335
9-1-2-4 Stanwell Park Claystone.............................................................. 337
9-1-2-5 Bulgo Sandstone.......................................................................... 337
9-1-2-6 Bald Hill Claystone....................................................................... 339
9-1-2-7 Newport Formation....................................................................... 339
9-1-3 Hawkesbury Sandstone..................................................................... 340
9-2 Discussion................................................................................................ 342
Chapter Ten: Conclusions........................................................................... 352
References......................................................................................................359

xvi

List of Figures
Chapter One
Figure 1-12
Figure 1-23
Figure 1-34
Chapter Three
Figure 3-1..37
Figure 3-2..38
Chapter Four
Figure 4-1..58
Figure 4-2..60
Figure 4-3..61
Figure 4-4..62
Figure 4-5..64
Figure 4-6..65
Figure 4-7..66
Figure 4-8..67
Figure 4-9..69
Figure 4-1071
Figure 4-1173
Figure 4-1275
Figure 4-1377
Figure 4-1479
Figure 4-1582
Figure 4-1684
Figure 4-1787
Chapter Five
Figure 5-1109
Figure 5-2110
Figure 5-3 ...112
Figure 5-4113
xvii

Figure 5-5114
Figure 5-6115
Figure 5-7116
Figure 5-8117
Figure 5-9118
Figure 5-10.119
Figure 5-11.121
Figure 5-12.122
Figure 5-13.124
Figure 5-14.125
Figure 5-15.127
Figure 5-16.129
Figure 5-17.130
Figure 5-18.131
Figure 5-19.132
Figure 5-20.133
Figure 5-21.139
Figure 5-22.140
Figure 5-23.141
Figure 5-24.142
Figure 5-25.143
Figure 5-26.144
Figure 5-27.146
Figure 5-28.147
Figure 5-29.148
Figure 5-30.150
Figure 5-31.151
Figure 5-32.153
Figure 5-33.154
Figure 5-34.155
Figure 5-35.156
Figure 5-36.157
xviii

Figure 5-37.158
Figure 5-38.159
Figure 5-39.160
Figure 5-40.161
Figure 5-41.165
Figure 5-42.166
Figure 5-43.168
Figure 5-44.169
Figure 5-45.170
Figure 5-46.171
Chapter Six
Figure 6-1177
Figure 6-2178
Figure 6-3189
Figure 6-4192
Figure 6-5202
Figure 6-6204
Figure 6-7205
Figure 6-8207
Figure 6-9215
Chapter Seven
Figure 7-1219
Figure 7-2220
Figure 7-3221
Figure 7-4224
Figure 7-5226
Figure 7-6227
Figure 7-7228
Figure 7-8229
Figure 7-9230
Figure 7-10..232
Figure 7-11..234
xix

Figure 7-12..235
Figure 7-13..236
Figure 7-14..237
Figure 7-15..240
Figure 7-16..241
Figure 7-17..242
Figure 7-18..243
Figure 7-19..244
Figure 7-20..244
Figure 7-21..245
Figure 7-22..247
Figure 7-23..248
Figure 7-24..253
Figure 7-25..254
Figure 7-26..255
Figure 7-27..256
Figure 7-28..258
Chapter Eight
Figure 8-1317
Figure 8-2318
Figure 8-3319
Chapter Nine
Figure 9-1323
Figure 9-2324
Figure 9-3 ...324
Figure 9-4327
Figure 9-5328
Figure 9-6333
Figure 9-7334
Figure 9-8335
Figure 9-9336
Figure 9-10..337
xx

Figure 9-11..338
Figure 9-12..341

List of Tables
Chapter One
Table 1-16
Table 1-211
Chapter Three
Table 3-133
Chapter Five
Table 5-1.111
Chapter Seven
Table 7-1.222
Table 7-2.222
Table 7-3.223
Table 7-4.250
Chapter Eight
Table 8-1.319
Chapter Nine
Table 9-1.350

List of Appendices (Volume 2)


Appendix one.1
Appendix Two5
Appendix Three24
Appendix Four..25
Appendix Five..71
Appendix Six..116
Appendix Seven167
Appendix Eight..179

xxi

Chapter one
1-1 Introduction
The Sydney Basin forms the southern division of the Sydney-Gunnedah-Bowen
Basin (Fig.1-1). This study includes the Permian-Triassic Hawkesbury
Sandstone, Narrabeen Group and Illawarra Coal Measures, which form part of
the southern Sydney Basin, New South Wales. It addresses the relationships
between petrology, porosity and permeability in the southern Sydney Basin
succession, NSW.
This thesis comprises ten chapters. Chapter one includes an introduction which
supplies significant information about the importance of research in the study
area. The introduction contains the location of the study area, previous work
and aims. Chapter two includes a general literature review of similar studies to
this research project. It concentrates on petrology, diagenesis, sedimentology
and facies analysis. Chapter three indicates the methods that were used to
accomplish this research. Chapter four includes a review of previous work on
sandstone environments of deposition and depositional systems. Chapter five
shows the results of the study. This chapter describes the main components,
textures, porosities and permeabilities of conglomerate, sandstones, siltstone,
mudstone and claystone. Chapter six interprets the diagenetic alteration and
cement types in the study area. Chapter seven discusses the porosity,
permeability and hydrological implications of the studied succession. Chapter
eight includes linking depositional environment, provenance and composition
variation. Also, it includes linking depositional environment, diagenetic and
composition variation. Also, it includes the influence of diagenetic alteration and
composition on reservoir quality. Chapter nine shows reservoir potential of the
Southern

Sydney

recommendations.

Basin
This

Chapter

chapter

ten

includes

develops
a

summary

conclusions
of

the

results,

achievements of the research in addition to the important recommendations.

and

Figure 1-1: Structural-element map of the Bowen-Gunnedah-Sydney Basin


system and the Surat Basin. Structural elements after Exon (1976), Beckett
et al. (1983), Balfe et al., (1988), Yoo (1988) and Brakel (1989).
Superimposed on the structural elements is the position of the Meandarra
Gravity Ridge (from Krassay et al., 2009).
2

Study area

Figure 1-2: Location of the Sydney Basin and the Coalfields within it
(from Grevenitz et al., 2003).

EC Appin West DDH 156

Figure 1-3: Location of wells.

1-2 Location of study area


The study area consists of part of the southern Sydney Basin in the Illawarra
district of New South Wales, Australia (Fig.1-2). It includes the Hawkesbury
Sandstone, Narrabeen Group and Illawarra Coal Measures with Figure 1-3
showing the location of wells studied in this thesis.

The Late Permian Illawarra Coal Measures is located over the Shoalhaven
Group and is predominantly overlain by the freshwater Triassic Narrabeen
Group. An area of about 8600 km is covered by the Southern Coalfield, which
is centred around the city of Wollongong (Bamberry, 1992). In the southern and
western divisions of Sydney Basin, the Illawarra Coal Measures are extensive.
Up to 50-100 km south, west and north of Wollongong, the coal measures are
mined where they are nearer to the surface than in the Sydney area (Haworth,
2003).
In the Sydney Basin, the Narrabeen Group is the lowermost of three Triassic
rock units and is located between the Late Permian Illawarra Coal Measures
and the Hawkesbury Sandstone (Ward, 1971a).
The Middle Triassic Hawkesbury Sandstone is exposed widely in the Sydney
Basin, lying above the Narrabeen Group and beneath the Wianamatta Group. In
the Garie Bundeena area, the thickness of the Hawkesbury Sandstone is 230
m (Rust and Jones, 1987) where the floodplain facies of the Mittagong
Formation conformably overlies the Hawkesbury Sandstone and is overlain by
the argillaceous Wianamatta Group (Herbert, 1980a).

1-3 Stratigraphy of study area


Permian and Triassic marine and non-marine sedimentary rocks are observed
in the southern Sydney Basin. Permian rocks include the Shoalhaven Group
and the Illawarra Coal Measures (Table 1-1). Triassic rocks consist of the
Narrabeen Group, the Hawkesbury Sandstone and the Wianamatta Group
(Table 1-1).
5

Table 1-1: Stratigraphy of Southern Coalfield (Modified after Bunny, 1972;


Bowman, 1974, 1980; Carr, 1983; Bamberry, 1992).
6

1-3-1 Illawarra Coal Measures


Many studies have discussed the nomenclature of the Illawarra Coal Measures
(e.g. Bowman, 1970; Bunny, 1972; Bowman, 1974; Carr, 1983; Bamberry,
1992). The coal seams in the southern Sydney Basin were included in the
Illawarra Coal Measures by Harper (1915) while Hanlon (1956) introduced
proposed nomenclature for the six major seams. In addition to this, the
nomenclature of the non-coal formations in the Illawarra Coal Measures was
developed by Bunny (1966). The Illawarra Coal Measures are classified into two
subgroups, which include the Cumberland Subgroup and the overlying Sydney
Subgroup (Table 1-1; Bowman, 1970; Bowman, 1980; Jones, 1983). The
Sydney Subgroup is separated from the underlying Cumberland Subgroup by a
noticeable erosion surface (Bowman, 1980).
1-3-1-1 Cumberland Subgroup
The Cumberland Subgroup contains the Pheasants Nest Formation and the
Erins Vale Formation, and originated from a volcanic source to the east and
from the more quartzose basement rocks located southwest of the Sydney
Basin (Jones, 1983). The Pheasants Nest Formation is present at the base of
the Cumberland Subgroup and overlies the upper Shoalhaven Group. The
thickness of the Pheasants Nest Formation is 75 m (Bowman, 1980) and it
includes lithic sandstone, thin coal seams and plentiful conglomerate in the
Robertson area (Hutton and Bamberry, 1999). It contains four latite members:
Five Islands, Calderwood, Minnamurra and Berkeley (Carr, 1983). Harper
(1915) named the Five Islands Flow, which was altered to the Five Islands
Latite Member by Carr (1983). The Calderwood Latite Member was named by
Carr (1983) for outcrops around Calderwood. The Minnamurra flow was
identified by Jaquet et al. (1905) but modified to the Minnamurra Latite by Joplin
et al. (1952). The Berkeley Flow was recognised by Harper (1915) then
renamed as the Berkeley Latite by McElory (1952) and Joplin et al. (1952).
These members are sourced from the Gerringong Volcanics (Bamberry et al.,
1989).

The Erins Vale Formation represents the uppermost formation of the


Cumberland Subgroup and is 37 m thick (Bowman, 1980). It consists of
conglomerate interbedded with coarse-grained sandstone while bioturbated
siltstone (Kulnura Marine Tongue) is recorded in the upper part of the Erins
Vale Formation (Hutton and Bamberry, 1999).

1-3-1-2 Sydney Subgroup


The Sydney Subgroup is the upper sub-group of the Illawarra Coal Measures
and contains lithic sandstone, siltstone, mudrock, tuff and coals. The Wilton
Formation is present at the base of the Sydney Subgroup and consists of
coarse sandstone, siltstone and coal. It has a thickness that varies between 1530 m (Bowman, 1980). Two members recognised in the Wilton Formation are
the Woonona Coal Member and the Wanganderry Sandstone Member.
Carbonaceous claystone, coal and tuffaceous claystone are present in the
Woonona Coal Member (Roche, 1997). It has a maximum thickness of 3 m
(Bowman, 1974). Also, Roche (1997) noted that the Wanganderry Sandstone
Member included medium- to very coarse-grained quartzose sandstone,
claystone, carbonaceous siltstone and pebbly conglomerate. It has a maximum
thickness of about 22 m at Yerranderie (Hutton and Bamberry, 1999).
The Tongarra Coal overlies the Wilton Formation and consists of carbonaceous
claystone and interbedded coal, sandstone and tuff. It has a thickness that
ranges from 2 to 3 m (Bowman, 1974). It underlies the Bargo Claystone, which
is overlain by the Dark Forest Sandstone (Bowman, 1974). The Bargo
Claystone and the Dark Forest Sandstone were defined as members of the
Appin Formation but Bunny (1972), Bowman (1974; 1980) and Bamberry (1992)
showed that they have extensive development, thus they are recognised as
formations. Bunny (1972) showed that the Bargo Claystone contains dark grey
to black claystone, while light grey, fine-grained, lithic to quartz-lithic sandstone
are recorded in the Dark Forest Sandstone. They have an average thickness of
about 15 and 10 m, respectively (Bowman, 1974). The Austinmer Sandstone
Member was identified at the base of the Bargo Claystone by Bowman (1970).

The Huntley Claystone Member occurs just above the base of the Bargo
Claystone (Hutton and Bamberry, 1999).
The Allans Creek Formation overlies the Darkes Forest Sandstone and contains
four coal intervals separated by sandstone, claystone and siltstone (Hutton and
Bamberry, 1999). It is 8 m thick (Bowman, 1974). The Kembla Sandstone
overlies the Allans Creek Formation and has a maximum thickness of about
23.8 m in the Southern Coalfield. It includes medium- to very coarse-grained
lithic sandstone and interbeds of siltstone and claystone (Roche, 1997). The
Wongawilli Coal overlies the Kembla Sandstone and includes the Farmborough
Claystone Member which is widespread in the Southern Coalfield (Hutton and
Bamberry, 1999). The Wongawilli Coal contains bands and splits and has a
thickness of between 4 and 11 m (Arditto, 2003).
Coals, shales, and interbedded sandstones of the Eckersley Formation overlie
the Wongawilli Coal (Bowman, 1970). The Eckersley Formation is composed of
the Novice Sandstone Member, Woronora Coal Member, Hargrave Coal
Member, Cape Horn Coal Member and Balgownie Coal Member (Bowman,
1974). Bamberry (1992) indicated new changes in the stratigraphy when he
introduced the Burragorang Claystone, above the Cape Horn Coal, and the
Loddon Sandstone Members at the top of the Eckersley Formation. Roche
(1997) noted that the Lawrence Sandstone Member exists between the Cape
Horn Coal and Balgownie Coal Member, and contains lithic sandstone,
claystone, siltstone and conglomerate. It has a maximum thickness of about 16
m in the Southern Coalfield. It is overlain by the Balgownie Coal Member which
consists of coal and carbonaceous claystone (Roche, 1997). The Loddon
Sandstone has subsequently been elevated to formation status (Herbert, 1995).
It occurs between the Balgownie Coal Member and the Bulli Coal and contains
lithic sandstone, claystone, siltstone and conglomerate (Hutton and Bamberry,
1999). The Bulli Coal is present at top of the Illawarra Coal Measures and has a
thickness that varies from 0 to 4 m (Bowman, 1974). It includes coal and
carbonaceous claystone (Roche, 1997).

Hutton (1990) used geophysical logs to study the coal measure sequences in
the southern Sydney Basin and showed that the Southern and Western
Coalfields could be correlated. For example, the Wongawilli seam contains
similar characteristics to the Middle River seam. Hutton et al. (1992), in a
comprehensive study of the Southern Coalfield, indicated the importance of
amendments to the previous stratigraphic nomenclature. They noted that the
Marrangaroo Conglomerate is present in the Southern Coalfield.
The stratigraphic correlations between the Illawarra Coal Measures and
Newcastle Coal Measures are shown by a number of authors (Briggs, 1993; Hill
et al., 1994; Herbert, 1995; Table 1-2). The studies of Briggs (1993) and Hill et
al. (1994) were based on bio-stratigraphic studies, and coal seam and facies
analysis respectively. The study of Herbert (1995) depended on sequence
stratigraphic interpretations. Hill et al. (1994) correlated the units in the
Eckersley Formation with the units in the Western and Newcastle Coalfields.
Thus, the Burragorang Claystone Member correlates with the Awaba Tuff; the
Balgownie Coal Member is similar to the Wallarah Great Northern Coal; and the
Bulli Coal correlates with the Vales Point Coal Member (Table 1-2). Herbert
(1995) introduced a comprehensive sequence stratigraphic analysis of the
Sydney Basin. He identified the sequence boundary and subsequent sequence
set at the top of the Dempsey Formation and overlying Waratah Sandstone of
the Newcastle Coal Measures. He showed the Huntley Claystone Member in
the Bargo Claystone correlates with Nobbys Tuff in the Newcastle Coal
Measures. Additionally, the Mt Hutton Formation and Australasian Coal in the
Newcastle Coal Measures correlate with the Kembla Sandstone and American
Creek Coal in the Illawarra Coal Measures. The Eckersly Formation in the
Illawarra Coal Measures correlates with the Pilot Coals and Reids Mistake
Formation in the Newcastle Coal Measures (Herbert, 1995). Hutton and
Bamberry (1999) provided the latest stratigraphic division and classification of
the Illawarra Coal Measures across the southern coalfield. They indicated that
the Wilton Formation includes the Wanganderry Sandstone Member. The Bargo
Claystone includes the Huntley Claystone Member. The Wongawilli Coal
includes the Farmborough Claystone Member. The Eckersley Formation
includes the Burragorang Claystone and the Loddon Sandstone.
10

Southern Coalfield

Newcastle Coalfield

Narrabeen Group

Narrabeen Group

Illawarra Coal Measures

Newcastle Coal Measures

Bulli Coal

Vales Point Coal Member

Loddon Sandstone

Moon Island Beach Formation

Balgownie Coal Member

Wallarah Great Northern Coal

Burragorang Claystone Member

Awaba Tuff

Eckersly Formation

Boolaroo Formation

Wongawilli Coal

Warners Bay Tuff

Farmborough Claystone Member

Mt Hutton Tuff

Kembla Sandstone
American Creek Coal

Australasian Coal

Allans Creek Formation

Adamstown Formation
Stockrington Tuff

Darkes Forest Sandstone

Lambton Formation

Huntley Claystone Member

Nobbys Tuff

Bargo Claystone

Waratah Sandstone

Tongarra Coal

Dempsey Formation

Wilton Formation

Four Mile Creek Subgroup

Woonona Coal Member


Wallis Creek Subgroup

Erins Vale Formation


Pheasants Nest Formation

Table 1-2: Stratigraphic correlation between the Illawarra Coal


Measures and Newcastle Coal Measures (from Grevenitz et al., 2003).
.

1-3-2 Narrabeen Group


1-3-2-1 Source
Herbert (1997) determined that three sources provided detritus for the
Narrabeen Group, including: The New England Orogen to the northeast (quartz

11

lithic detritus), the Offshore Uplift to the east (volcanolithic detritus), and the
cratonic Lachlan Fold Belt to the west (mature quartz detritus).

1-3-2-2 Stratigraphy
Hanlon et al. (1953) introduced a study of the subdivisions of the Narrabeen
Group and correlations between the South Coast and Narrabeen Wyong
Districts. They classified the Narrabeen Group in the Sydney Basin into the
Clifton Subgroup and the Gosford Subgroup (Table 1-1). The Narrabeen Group
occurs between the Hawkesbury Sandstone and the Illawarra Coal Measures
and is the lowermost part of the three Triassic units in the Sydney Basin (Ward,
1972). Bowman (1974) in a study of the stratigraphy of the Wollongong area
recognised the Clifton Subgroup and Gosford Formation and provided a
description and thickness of each unit. Goldbery and Holland (1973) indicated
that the maximum thickness of the Triassic Narrabeen Group was 520 m.
However, Mayne et al. (1974) noted that the Latest Permian to Middle Triassic
Narrabeen Group has a thickness which varies between 90 and 700 m.
Cameron et al. (1999) noted that the Narrabeen Group has a thickness of 30 m
at Barneys Reef, 80 m at Munghorn Gap and 163 m in the DM Curryall DDH1 at
Turrill.
1-3-2-2-1 Clifton Subgroup
The Clifton Subgroup includes the Coal Cliff Sandstone, Wombarra Claystone,
Scarborough Sandstone, Stanwell Park Claystone, Bulgo Sandstone and Bald
Hill Claystone (Table 1-1; Bowman, 1974).
Dickson (1967; 1969) suggested that the Coal Cliff Sandstone should be
introduced as a member of the Wombarra Claystone. The Coal Cliff Sandstone
is the lowermost unit of the Clifton Subgroup on the Illawarra coast and contains
light grey, quartz lithic fine-skewed sandstone (Bowman, 1974). It has a
maximum thickness varying between 6-20 m. The Coal Cliff Sandstone is
overlain by the Wombarra Claystone which underlies the Scarborough
Sandstone (Bowman, 1974). The Wombarra Claystone includes green
mudstone and siltstone with interbeds of sandstone and conglomerate (Moffitt,
12

2000). It has a thickness that ranges from 0 to 30 m (Bowman, 1974). The


Otford Sandstone Member is a conglomeratic interbed in the Wombarra
Claystone (Dehghani and Jones, 1994a). The Scarborough Sandstone is
located between the Wombarra Claystone and the Stanwell Park Claystone,
and is 24 m thick. It is composed of conglomerate and sandstone (Bowman,
1974). The Stanwell Park Claystone overlies the Scarborough Sandstone and
underlies the Bulgo Sandstone. It has a thickness that varies between 0-37 m,
and includes claystone and sandstone intervals. The Stanwell Park Claystone is
overlain by the Bulgo Sandstone, which has a thickness that ranges from 90 m
to 130 m. Conglomerate and sandstone are observed in the Bulgo Sandstone
(Bowman, 1974). The Bald Hill Claystone overlies the Bulgo Sandstone and
includes red and limited cream claystone and mudstone (Moffitt, 2000). The
thickness of the Bald Hill Claystone varies between 1 and 20 m (Bowman,
1974).
Hamilton and Galloway (1989) showed that the Narrabeen Group includes five
operational units: the Wombarra operational unit, the Scarborough Sandstone
operational unit, the lower Bulgo operational unit, the upper Bulgo operational
unit and the Bald Hill operational unit. These nomenclatures depended on
neutron and gamma ray logs. Also, Bai (1991) used this classifications in his
study. Bai (1991) showed that the thickness of the Wombarra operational unit
varies between 11 and 192 m, the Scarborough Sandstone operational unit
varies between 20 and 152 m, the lower Bulgo operational unit varies between
10 and 149 m, the upper Bulgo operational unit varies between 16 and 185 m,
and the Bald Hill operational unit varies between 12 and 76 m.
1-3-2-2-2 Gosford Subgroup
The Gosford Subgroup consists of the Garie Formation and Newport Formation
(Table 1-1; Bowman, 1974). Loughnan (1969) recognized the Garie Member at
the top of the Bald Hill Claystone and it was later given formation status by
Bunny and Herbert (1971). The Garie Formation is the lower unit of the Gosford
Subgroup and includes cream claystone to grey claystone, tuff with accretionary
lapilli (Heritage, 2005). The interbedded dark grey, mudstone, siltstone and fine
sandstone of the Newport Formation overlie the Garie Formation and are the
13

upper unit of the Gosford Subgroup (Moffitt, 2000; Heritage, 2005). The Garie
Formation and the Newport Formation have thicknesses that range from 0 to 3
m and from 0 to 18 m, respectively (Bowman, 1974).

1-3-3 Hawkesbury Sandstone


1-3-3-1 Thickness
The Hawkesbury Sandstone overlies the Narrabeen Group and underlies the
Wianamatta Group (Table 1-1).
In the western area of outcrop, the Hawkesbury Sandstone is between 30-60 m
in thick, whereas in the central area of outcrop, it ranges from 210 to 290 m in
thick (Raggatt, 1938; McElroy, 1962; Conolly, 1964). The Hawkesbury
Sandstone covers about 20,000 km of the Sydney Basin and has its maximum
thickness of 290 m at the Hawkesbury River (Herbert, 1983). Branagan (2000)
reduced the unit cover area to about 12,500 km and the thickness to vary
between 30 and 240 m. Lee (2000) noted that the thickness of Hawkesbury
Sandstone varies from about 160 m in the Mittagong region, to 250 m in the
Sydney district.

1-3-3-2 Source
Standard (1969) classified the source of the Hawkesbury Sandstone into two
areas: a southwestern source area, and a western source area based on
interpretation of heavy and clay minerals. The southwestern area is the major
source (Standard, 1969) since the sandstone appears have been derived from
the southwest direction based on palaeocurrent studies (Standard, 1969;
Veevers, 1984). The cratonic Lachlan Fold Belt is the source of the Hawkesbury
Sandstone (Standard, 1969; Veevers, 1984). Volcanic pebbles derived from the
New England source are present in some Hawkesbury Sandstone beds in the
Somersby area (Mills et al., 1989). Branagan (2000) suggested the possibility of
an intertonguing of the latest Narrabeen sediment, and the earliest Hawkesbury
sands, derived from the north and the south, respectively. Alternatively, he

14

suggested that deposited Narrabeen sediment was eroded by Hawkesbury


rivers and integrated into the latter beds.

1-4 Aims and Objectives


The purpose of this project is to study the petrology of the Hawkesbury
Sandstone, Narrabeen Group and Illawarra Coal Measures in the southern
Sydney Basin, New South Wales, Australia, and to determine its relationship to
porosity and permeability.
1. A combination of stratigraphy, sedimentology and petrology are used to
define

the

processes

of

deposition,

the

palaeogeography

and

paleoclimatology of the basin.


2. Additionally, the study aims to determine the origin of the main
sandstone types in the basin.
3. Identification of the diagenetic history of the sandstone in the basin is
another major aim of this study.
4. This study will construct a depositional environmental model for the
sandstones in the southern Sydney Basin.
5. The use of petrography, well log analysis and correlation are important to
understand sandstone reservoir characterization, which is a focus of the
study.
6. The definition of hydrological systems in the study succession, and their
relevance for water storage and mining is a new focus of study in this
project.

15

16

Chapter Two
Literature review
2-1 Sandstone texture and composition
Sandstone textures are variable in their shape, size and sorting. Texture varies
between very fine- to very coarse-grained, with grain sizes between 0.063 and
2.0 mm (Stow, 2005). For example, Buyukutku and Bagci (2005) divided the
Kuzgun Formation Sandstone in Adana Basin, southern Turkey, into three
lithological units depending on texture and mineralogy. In addition, Tamrakar et
al. (2007) studied the relationship between texture and composition, along with
mechanical and physical features in Siwalik sandstones of the central Nepal
Sub-Himalayas. This study indicated that the environment can be identified by
texture and composition through sedimentation processes, weathering and
diagenesis. Thus, the mechanical and physical characteristics of the rocks are
related to their texture and composition.
Sandstone composition comprises quartz, feldspar, rock fragments, clay
minerals, micas and heavy minerals (El-Ghali et al., 2009a, b). Quartz is
generally the most common mineral in sandstones, with the proportion reaching
up to 75%, while feldspar can also be an important mineral present in some
sandstones (El-Ghali et al., 2006a, b; Wanas, 2008; Sonel et al., 2009; Wolela,
2009). Additionally, most sandstone units have kaolinite, illite, mixed-layer
illite/smectite and chlorite as clay minerals whether detrital or authigenic
(Mansurbeg et al., 2008; Zhang et al., 2008). Heavy minerals recorded in
sandstones include epidote, garnet, tourmaline, rutile, apatite, zircon and other
minerals depending on specific source area (Mansurbeg et al., 2008; El-Ghali et
al., 2009a; Islam, 2009). Micas include biotite and muscovite in many
sandstones (El-Ghali et al., 2006b, 2009a). Thin section petrology, SEM and
XRD are techniques commonly used to study the composition and texture in
sandstones (Bernet and Gaupp, 2005; Gier et al., 2008; Lee and Lim, 2008;
Zhang et al., 2008; Islam, 2009).

17

2-2 Diagenesis
Several factors control the sandstones diagenetic development including
framework grains, tectonic setting, pore-water chemistry and the burial-thermal
history (Morad et al., 2000; Stonecipher, 2000). Also, the sediment source,
burial environment and depositional system have a role in the controls on
diagenesis (e.g. Burley et al., 1985; Morad et al., 2000).
Many studies have discussed the diagenetic alterations in sandstones, including
compaction, quartz overgrowths, carbonate cement, authigenic clay minerals
and authigenic feldspar (e.g. Salem et al., 2000; Elias et al., 2004; Schmid et
al., 2004). Several authors have studied the development of spatial and
temporal aspects of diagenetic minerals based on detail studies of the
diagenetic minerals in terms of depositional facies and sequence stratigraphy
(e.g., Taylor et al., 1995; Morad et al., 2000; Ketzer et al., 2003a,b; El-Ghali et
al., 2006 a,b,c; Karim et al., 2010). Also, McKinley et al. (2011) studied the
relationship between diagenetic processes and porosity and permeability using
geographically weighted regression (GWR).

2-2-1 Compaction
Sandstones show two types of compaction - mechanical and chemical
compaction (Salem et al., 2000; Abouessa and Morad, 2009; El-Ghali et al.,
2009a). Types of grain contacts support the presence of compaction in
sandstones (Sur et al., 2002; Zhang et al. 2008; El-Ghali et al., 2009a; Higgs et
al., 2010; Umar et al., 2011). Bent mica flakes and pseudoplastic deformation of
lithic grains are indicators for the occurrence of mechanical compaction in
sandstones (El-Ghali et al., 2009a). Also, Salem et al. (2000) showed that grain
fracturing and rearrangement also mark the presence of mechanical
compaction in sandstones. Pressure dissolution is evidence for the presence of
chemical compaction in sandstones as shown by Lima and De Ros (2002),
Zhang et al. (2007) and Zhang et al. (2008).

18

2-2-2 Quartz cementation


Worden et al. (2000) noted that quartz cement may occur as pore throatblocking clots. In most sandstone, quartz cement exists commonly as euhedral,
syntaxial overgrowths around quartz grains (Ketzer et al., 2002, 2003a; El-Ghali
et al., 2006a, b, 2009a, b) and also it fills pore spaces in many types of
sandstone (Sur et al., 2002; Lee and Lim 2008). It also occurs rarely as
prismatic crystals (Reed et al., 2005). Dust lines may be observed in quartz
overgrowths and differentiate them from detrital quartz grains in sandstones
(Reed et al., 2005; Parry et al., 2009). Quartz cement may be distinctive as
discrete crystals where the size is more than 100 m (Luo et al., 2009).
Mork and Moen (2007) used electron backscatter diffraction analysis (EBSD)
with petrography to describe compaction microstructures in quartz grains and
quartz cement in deeply buried sandstone. This reflected the possibility of using
EBSD analysis to determine the importance of ductile deformation in sediment
compaction. Makowitz and Milliken (2003) revealed that micro-fracturing in
quartz occurred throughout the process of quartz cementation.
Many studies have shown the occurrence of quartz cementation in sandstones
(e.g. Purvis, 1995; Baker et al., 2000; Forgotson et al., 2000; Harris and Bustin,
2000; Hartmann et al., 2000; Marfil et al., 2000; Girard et al., 2001; Rossi et al.,
2002; Hiatt et al., 2003; 2007; Salem et al., 2005; Al-Harbi and Khan, 2008;
Bertier et al., 2008; Abouessa and Morad, 2009; Van den Bril and Swennen,
2009).

2-2-3 Carbonate cementation


Carbonate cementation is prevalent in many types of sandstone and has
various forms of the occurrence (El-Ghali et al., 2006a, b, 2009a, b). The main
types of carbonate cementation include siderite, calcite, ankerite and dolomite
(Kim et al., 2007; Zhang et al., 2008; Hammer et al., 2010) and may be formed
during early and late stages of diagenesis (El-Ghali et al., 2006b).

19

Calcite cement is distinguished in many forms in sandstones, such as


microcrystalline, equigranular mosaic and prismatic (Ketzer et al., 2002).
Detrital grains may be surrounded by aggregates generated by microcrystalline
calcite (Ketzer et al., 2003a). Zhang et al. (2007) discovered many forms of
calcite including microcrystalline calcite filling pores, isopachous calcite rims,
blocky and prismatic calcite crystals, fine pseudospar mosaic calcite, coarse
non-ferroan mosaic calcite and coarse ferroan mosaic to poikilotopic calcite.
Also, calcite cement can be described in three forms as micrite, microsparite
and sparite (Hall et al., 2004). During diagenesis calcite may fill pore spaces
and replace detrital grains (Hall et al., 2004; Gecer-Buyukutku and Sahinturk,
2005; Salem et al., 2005; Kim et al., 2007; El-Ghali et al., 2009a).
Many forms of siderite cement are recorded as scattered rhombs and
commonly occur with pyrite in sandstones (Lima and De Ros, 2002). Siderite is
present as flattened rhombs in sandstones (Luo et al., 2009). It is also
characterised by many forms such as coarse-crystalline and microcrystalline
siderite described by El-Ghali et al. (2006c) who introduced a study of the origin
of siderite cement in sandstones. In many types of sandstone, siderite cement
fills pore spaces and replaces detrital grains (Kim et al., 2007). Siderite can be
classified as having a pure composition (Fe90) and siderite that is more
magnesium-rich (~Fe60-80) according to Hammer et al. (2010).
Ankerite cement is an important mineral and has many forms in thin section and
scanning electron microscope. Coarse-crystalline ankerite fills pores in
sandstones (Karim et al., 2010). Also, ankerite cement can occur as distinctive
rhombohedral crystals and many detrital grains are coated by ankerite cement
in sandstones (Kim et al., 2007). It varies in habit from anhedral to subhedral
(Lima and De Ros, 2002).
Schmid et al. (2003) introduced a study of the distribution of dolomite cement in
sandstones including non-ferroan dolomite and ferroan dolomite. The latter is
negligible compared with the former. The presence of ferroan dolomite in
sandstones and in conglomerates was described in a study by Zamanzadeh et
al. (2009) where it is stratabound in the former but occurs as distinct layers in
20

the later. Dolomite was recorded as small scattered rhombic crystals (El-Ghali
et al., 2009a). It has also been observed as syntaxial overgrowths indicated by
bright orange luminescence (Mansurbeg et al., 2009). Salem et al. (2005)
recognised the occurrence of minor rhombohedral dolomite crystals in
sandstone samples. Margins of detrital and overgrowth grains may be
influenced by dolomite cement which is present as grain replacement (Elias et
al., 2004).

2-2-4 Authigenic Clay minerals


In sandstones, the main authigenic clay minerals present are kaolinite, chlorite,
illite and mixed layer illite/smectite (Zhang et al., 2007; 2010). Many studies
have shown authigenic clay minerals in sandstones and described their forms
and sizes. Pay et al. (2000) studied the distribution of clay minerals in
sandstones and noted their occurrence as pore-filling and pore-lining clays.
Kaolinite is distinguished from other authigenic clay minerals by its occurrence
as booklets and vermicular aggregates (Rossi et al., 2001; Ketzer et al., 2003a;
El-Ghali et al., 2006b) whether in micas or in the matrix (El-Ghali et al., 2006b).
Zhang et al. (2007) showed four forms of kaolinite including small columns,
individual plates, multiple and large columnar aggregates. Kaolinite occurs in
intergranular pores and also many detrital grains are coated by kaolinite (Ketzer
et al., 2003a). Other studies showed the presence of kaolinite as a replacement
mineral (Milliken, 1998, 2001, 2003). Temperature affects kaolinite and thus
high temperatures are associated with low kaolinite content (Dutton and Loucks,
2009).
Illite is distinguished from other authigenic clay minerals by its fibrous habit (ElGhali et al., 2006b; Karim et al., 2010). Zhang et al. (2007) described two other
forms of illite in addition to fibrous that are hair-like and membranous. The
smectitic precursor mineral contributes to the formation of clay minerals such as
illite at low temperatures (Storvoll et al., 2002). Cagatay et al. (1996) showed
that illite could occur in the form overgrowths of projecting laths. Also, illite has

21

been observed in the form coatings on grains and also as scattered patches
(e.g. Abouessa and Morad, 2009).
El-Ghali et al. (2006b) described chlorite as scattered pseudohexagonal
platelets in detrital grains. Detrital grains such quartz in sandstones may be
coated by grain-coating chlorite (Tamar-Agha, 2009). The chemical composition
of pore waters can be affected by chlorite in sediments (Berger et al., 2009).
Also, pH and Mg and Fe ions are associated with chlorite formation from fluids
arising during diagenesis (Berger et al., 2009).
Mixed-layer illite/smectite clays are described in many studies as clays filling
pores and coating grains in sandstone (Zhang et al., 2008; 2010). Perri et al.
(2008) described clay mineral assemblages in the Mesozoic Longobucco Group
showing that mudstone mineralogy is characterised by mixed-layer illitesmectite.
Many studies have discussed authigenic clay minerals during diagenesis (e.g.
Aagaard et al., 2000; Grigsby, 2001; Lanson et al., 2002; Billault et al., 2003;
Lemon and Cubitt, 2003; El-Ghali et al., 2006a, b; Zhang et al., 2009; Gould et
al., 2010).

2-2-5 Feldspars overgrowths


Many studies have described the occurrence of feldspar cement as distinct
overgrowths on feldspar grains in sandstones (Ketzer et al., 2002, 2003a; ElGhali et al., 2009a; Wolela, 2009). Also, feldspar cement has been recorded in
the form of prismatic crystals (Ketzer et al., 2003a). Authigenic feldspar has
another role to fill intergranular pores (Zhang et al., 2008). Maraschin et al.
(2004) introduced a study of near-surface K-feldspar precipitation in
sandstones. They showed that feldspar overgrowths with a rhombohedral habit
are typical of pure KAISi3O8. Some studies emphasised the rarity of feldspar
cement in sandstones (Zamanzadeh et al., 2009). Ehrenberg and Jakobsen
(2001) attributed the rarity of plagioclase in sandstones to diagenetic reaction.

22

In general, many authors have shown the presence of feldspar cement in


sandstones (e.g. Ketzer et al., 2002; Rossi et al., 2002; Ketzer et al., 2003a;
Shah and Bandyopadhyay 2005; El-Ghali et al., 2006a, b; Zhang et al., 2007;
Kim et al., 2007; Wolela and Gierlowski-Kordesch, 2007; Mansurbeg et al.,
2008; Zhang et al., 2008; El-Ghali et al., 2009a; Luo et al., 2009; Mansurbeg et
al., 2009; Tamar-Agha, 2009; Wolela, 2009; Zamanzadeh et al., 2009;
Gonzalez-Acebron et al., 2010; Hammer et al., 2010).

2-3 Diagenesis and reservoir quality


The influence of diagenesis on reservoir quality has been assessed by many
authors in several sedimentary basins focusing on factors such as compaction
and cementation (e.g. Xiaomin et al., 2004; Salem et al., 2005; Higgs et al.,
2007; Molenaar et al., 2007; Zhang et al., 2010).
Several studies have determined the relationship of porosity and permeability in
rocks to diagenesis and sequence stratigraphy (e.g. Tucker, 1993; Morad et al.,
2000; Taylor et al., 2000). Depositional environment and clay mineral contents
are additional factors affecting porosity and permeability (Cocker et al., 2003;
Rahmani et al., 2003).

2-3-1 Compaction
Reservoir quality is decreased by compaction processes in sandstones (e.g.
Harris and Bustin, 2000; Li et al., 2002; Chi et al., 2003; Wolela and GierlowskiKordesch, 2007; Bertier et al., 2008; Ehrenberg et al., 2008; Song et al., 2009;
Dutton and Loucks, 2009; Islam, 2009;). Porosity is influenced by both
mechanical and chemical compaction (Higgs et al., 2007). Physical compaction
causes the reduction of porosity in sandstones by deformation of lithic
fragments (Gecer-Buyukutku and Sahinturk, 2005). Many studies have shown
that mechanical compaction often leads to porosity reduction associated with
increasing depth (Ehrenberg et al., 2008). Paxton et al. (2002) calculated the
influence of compaction on porosity and included an intergranular volume
compaction curve in their study. Pseudomatrix has a negative influence on
23

reservoir quality where the latter decreases with the presence of the former (ElGhali et al., 2009a).
Porosity loss by compaction has been described by many authors (e.g. Ramm
and Bjorlykke, 1994; Marfil et al., 1996; Higgs et al., 2007; dos Anjos et al.,
2000; Zhigang, 2000; Akaegbobi and Adeleye, 2001; Langford and Depret,
2001; Caetano-Chang and Wu, 2003; Ryu, 2003; Kim and Lee, 2004; AlRamadan et al., 2005; Ahmad and Bhat, 2006).

2-3-2 Quartz cementation


Quartz cementation commonly reduces porosity and permeability (Rossi et al.,
2002; Worden et al., 2000; Luo et al., 2009). It fills pores in sandstones, and
porosity is reduced by this process (Vosylius, 1998; Kilda and Friis, 2002;
Molenaar et al., 2007). Walderhaug (2000) introduced a study of modelling
quartz cementation and porosity in sandstones and contrasted petrography,
quartz cement and porosity using the EXEMPLAR program. In many studies,
the relationship between quartz cement and porosity is inverse, where the latter
decreases with increasing quartz cement (Chi et al., 2003; Molenaar et al.,
2007; Zhang et al., 2008).
Porosity loss due to quartz cement was also demonstrated by Cooper et al.
(2000), Giles et al. (2000), Rossi et al. (2002), El-Ghali et al. (2009a) and Wang
et al. (2009).

2-3-3 Carbonate cementation


Carbonate cementation reduces porosity in sandstones (Jia and Gu, 2002;
Wolela, 2009). In some cases, it is also important in the development of porosity
in sandstones, for example the dissolution of carbonate cement in late
diagenesis can conserve porosity (Chi et al., 2003; Zhang et al., 2007, 2008).
Many studies have shown that porosity varies as a result of pore-filling calcite
which can have a variable distribution shown by petrographic and log analysis
(e.g. Vogler and Robison, 1987; Taylor, 1990; Taylor and Land, 1996; Taylor et
al., 2010).
24

2-3-4 Authigenic clay minerals


Many studies have shown that clay minerals affect reservoir quality through
their textures and distribution (e.g. de Waal et al., 1988; Dutton and Diggs,
1990; Nadeau and Hurst, 1991; Worthington, 2003; Tamar-Agha, 2009).
Al-Ramadan et al. (2004) described loss of porosity as a result of pore-filling
illite and chlorite whereas grain coating chlorite preserves porosity in
sandstones. Also, porosity is decreased by clays such as kaolinite and illite
which fill pore throats (Borgohain et al., 2010). Nadeau (1998) found that the
development of smectite led to significant permeability loss in sandstones.
Many studies have shown the importance of grain coatings in the preservation
of porosity in sandstones (e.g. Heald and Larces, 1974; Thomson, 1979;
Pittman et al., 1992; Hillier et al., 1996; Primmer et al., 1997; Anjos et al., 2003;
Cocker et al., 2003; Worden and Morad, 2003; Bakken et al., 2004; Thomas et
al., 2004; Salem et al., 2005; Zhang et al. 2007; Allen, 2008; Smosna and
Sager, 2008; Zhang et al., 2008; Islam, 2009; Ajdukiewicz et al., 2010; Tobin et
al., 2010). Also, Ehrenberg (1993) and Bloch et al. (2002) showed that porosity
and permeability are mostly preserved by chlorite coatings in shallow marine
sandstones. Sandstone porosity has been preserved in other areas by graincoating smectite clays which prevent the development of quartz overgrowths
(Morad et al., 2010). Storvoll et al. (2002) indicated that illite coatings also
inhibited quartz overgrowths and contributed to the preservation of porosity in
sandstones. The influence of grain-coating chlorite in the preservation of
porosity is most important where it suppresses the development of quartz
overgrowths (Zhang et al., 2008; Taylor et al., 2010). Feldspar-bearing
sandstones include permeability which is affected by illite precipitated during
late stage diagenesis (e.g. Ramm and Ryseth, 1996; Midtbo et al., 2000).

2-3-5 Dissolution
Four factors are important to support the development of secondary porosity
including sedimentary facies, diagenesis, faults and shallow depth (Miao et al.,
2000). Dissolution is the major process in the development of secondary
25

porosity in sandstones, enhancing the total porosity in sandstones (Dutton,


2008). Many studies showed that dissolution of feldspars grains (e.g. Milliken,
2004; Lima and De Ros, 2002; Maraschin et al., 2004; Al-Ramadan et al., 2005;
Buyukutku and Bagci, 2005; Zhang et al. 2007; Ajdukiewicz et al., 2010; Tobin
et al., 2010), rock fragments (Luo et al., 2009; Tobin et al., 2010), mica (Luo et
al., 2009; Hammer et al., 2010) and earlier carbonate cement (Forgotson et al.,
2000; Al-Ramadan et al., 2005; Buyukutku and Bagci, 2005; Zhang et al. 2007;
Luo et al., 2009) support the presence of secondary porosity in sandstones.
Also, secondary porosity may result from dissolved interstitial clay material as
shown by Zhang et al. (2010). Other studies indicated that dissolved quartz or
carbonate are important factors in the formation of secondary porosity (e.g.
Lima and De Ros, 2002; Gecer-Buyukutku and Sahinturk, 2005; Borgohain et
al., 2010).
The types of secondary pores include exhumed primary porosity (Xiaomin et al.,
2004; Luo et al., 2009), oversized and mouldic pores (Lima and De Ros, 2002;
Zhang et al., 2007; Luo et al., 2009; Wolela, 2009), elongate pores (Wolela,
2009), micropores and fissures (Zhang et al., 2007).
In general, secondary porosity has been studied by many authors in many
different areas (e.g. dos Anjos et al., 2000; Forkner et al., 2000; McKinley et al.,
2001; Ketzer et al., 2003a; Elias et al., 2004; Reed et al., 2005; KunleDare,
2007; Zhiqian et al., 2007; Abouessa and Morad, 2009; Borgohain et al., 2010;
Saikia et al., 2011).

2-4 Grain size and reservoir quality


Average grain size and sorting are important factors in the determination of
permeability (Mattson and Chan, 2004). Thus, coarse-grained sandstones
which have less diagenetic cements can have higher permeabilities than finegrained sandstones (Buyukutku and Bagci, 2005).
Permeability is clearly influenced by average grain size, lithology and sorting
(Barton 1994; Doyle and Sweet 1995; Liu et al., 1996; Dutton and Willis 1998).
26

Barton (1994) suggested that grain size, mineralogical composition, sorting and
porosity can all affect permeability.

2-5 Detrital grains and reservoir quality


Ductile lithic sand grains contribute to porosity loss (Worden et al., 2000). Also,
the negative influence of ductile grains on permeability has been demonstrated
in some studies (e.g. Worden et al., 2000). Detrital carbonate grains also
reduce porosity in sandstones (Gier et al., 2008). Framework grains such as
feldspars and quartzose lithic fragments have a positive influence on porosity.
They also support the presence of secondary porosity through their dissolution
(Ketzer et al., 2003a). Some studies have shown that pseudomatrix helps in the
occurrence of microporosity (Ketzer and Morad, 2006). The percentage of
porosity reduction in sandstone that includes ductile extrabasinal and
intrabasinal carbonate grains is more than in sandstone which includes
mechanically stable quartz grains (Bloch et al., 2002; Paxton et al., 2002).

2-6 Depositional environment and reservoir quality


Several studies have indicated the importance of depositional environment and
its influence on reservoir quality (e.g. Zhang et al., 2008; Islam, 2009; Wolela,
2009). Coarse-grained sandstone deposited under high energy conditions is
characterized by a high reservoir quality whereas fine-grained sediments
deposited in low energy environment have much less reservoir quality (Zhang et
al., 2008). Some fluvial environments have high reservoir quality because of
their fine to medium grain size and good sorting with the presence of massive to
laminated textures (Islam, 2009). Depositional environments have a strong
influence on porosity and permeability in braided river delta sandstone
reservoirs (Bo and Lu, 2000; Sun et al., 2007). Zhou et al. (2006) showed that
porosity and permeability increase in distributary channel sandstones in a
braided river delta environment.

27

In fluvial-dominated units, permeability is reduced in low energy, fine-grained


deposits but increased in high energy, coarse-grained deposits (Mattson and
Chan, 2004). Permeability in aeolian sandstones is high compared with fluvial
and massive sandstone facies. Also, sand sheets and dune sandstones are
characterised by higher permeability compared with those in sabkha facies
(Bloomfield et al., 2006).
Ramm (2000) studied the relationship of reservoir quality with facies indicating
that the reservoir quality in estuarine channel/mouth bar and foreshore
sandstones is higher than in shoreface and sandy bay-fill facies. Deltaic
sandstones show greater loss of reservoir quality than fluvial sandstones
because of compaction (Luo et al., 2009). Coarse-grained fluvial sandstones
are more strongly affected by dissolution of feldspar compared with finer
grained deltaic sandstones (Luo et al., 2009). Transgressive systems tract
sandstones contain higher reservoir quality than the lowstand systems tract and
highstand systems tract sandstones (El-Ghali et al., 2009a). Many studies have
shown the percentages of porosity in transgressive systems tract sandstones,
lowstand systems tract and highstand systems tract sandstones (e.g. Ketzer et
al., 2002; Ketzer et al., 2003a; El-Ghali et al., 2006b).

2-7 Provenance
Provenance and tectonic setting can be interpreted from sandstone detrital
composition (Dickinson et al., 1983; Cavazza and Ingersoll, 2005; Greene et al.,
2005). Provenance is helpful in identifying the tectonic setting of the source
area (Yoshida and Machiyama, 2004).
Recent studies of sandstone provenance based on chemical composition and
petrography were done by Asiedu et al. (2000), Noda et al. (2004), Wanas and
Abdel-Maguid (2006), Osae et al. (2006), Akarish and El-Gohary (2008), AlJuboury et al. (2009), Dey et al. (2009), Tijani et al. (2010) and Hossain et al.
(2010).

28

Provenance studies have also been conducted using age delineation. For
example, the Quadrilatero Ferri Fero Sandstone in the Sao Francisco Craton,
Brazil, was analysed by Hartmann et al. (2006) through use of zircon U-Pb
isotopes. Also, Pb-Pb zircon ages, palaeocurrent data and mineral chemistry of
detrital tourmaline can be used to interpret provenance in sandstones
(Nascimento et al., 2007). In addition, detrital tourmaline can be used to
determine provenance through its chemistry (Anani, 1999; von Eynatten and
Gaupp, 1999; Li et al., 2004; Morton et al., 2005). Many studies have used
heavy minerals to describe provenance of sandstones (e.g. Hallsworth and
Chisholm, 2008; Fontanelli et al., 2009).
Provenance has been described for different areas by many authors (e.g. Kwon
and Boggs, 2002; Getaneh, 2002; Gonzalez-Acebron et al., 2007; Ghazi and
Mountney, 2011).

2-8 Conclusions
Quartz, feldspar, rock fragment, clay minerals, mica and heavy minerals are
recorded in sandstones in variable percentages.
Compaction, quartz overgrowths, carbonate cements; authigenic clay minerals
and authigenic feldspar may be associated with early and late stages of
diagenesis.
Porosity and permeability are subject to many factors, including alteration,
diagenesis, depositional environment, grain size and type of detrital grains.
Alteration and diagenesis have the most influence on reservoir quality
compared with other factors.
Many factors contribute to the determination of provenance in sandstones.
These are petrography, chemical composition and heavy minerals.

29

30

Chapter Three
Methodology
3-1 Introduction
The methodology for this thesis involved analysis of other pre-existing research
in this area; collection of samples in the field and from drill holes; thin section,
X-ray diffraction and scanning electron microscopy analysis of samples; the use
of well logs to correlate between the wells; and petrography to describe the rock
characteristics and minerals present in the collected samples.
This chapter provides information about the major methodologies used to
interpret the petrology and sedimentology of the Illawarra Coal Measures,
Narrabeen Group and Hawkesbury Sandstone in the southern Sydney Basin.

3-2 Field Work


The field work was divided into two parts. The first part was based on the study
of surface samples, while the second part was based on the study of the
subsurface samples collected from coal exploration wells.
In the first part, field work included the observation and description of
sedimentary rock types and facies, in addition to the description of lithologies in
these rocks, comprising composition, grain size measurement and texture
forms. However, recognition of sedimentary structures and the identification of
palaeocurrent measurements was done in the field.
About 33 surface samples were collected in the field work from the study area
which included the Illawarra Coal Measures, Narrabeen Group and Hawkesbury
Sandstone in the southern Sydney Basin.
In the second part, field work included collection of 309 core samples selected
from 9 wells drilled by BHPBilliton Illawarra Coal in the southern Sydney Basin.
These wells are EAW 18a, EAW 30, EAW 42, EAW 156, EDEN 115, EDEN
31

124, EDEN 125, EDEN 126 and EDEN 127. Samples were collected from the
Illawarra Coal Measures, Narrabeen Group and Hawkesbury Sandstone.
The petrological and sedimentological study of the Illawarra Coal Measures,
Narrabeen Group and Hawkesbury Sandstone in the southern Sydney Basin
depended on a range of data and information being collected from the drilled
wells, whether samples or well logs.
The locations of these wells are shown in (Fig. 1-3). Table 3-1 details the
general information about these wells, including hole name, site ID, drilling
dates, type, easting, northing, collar and total depth.
The samples collected in the field comprised sandstone, siltstone, shale,
mudstone, claystone and conglomerate.

3-3 Laboratory work


Laboratory work depended on the use of many modern techniques such as thin
section, petrographic microscope, X-ray diffraction and scanning electron
microscope.

3-3-1 Thin section


Before preparation of thin sections, 208 samples were vacuum impregnated
with blue dyed resin for the purposes of description and study of the porosity
under the microscope (cf. Lima and De Ros, 2002; Wolela and GierlowskiKordesch, 2007; Mansurbeg et al., 2008; Gier et al., 2008; Wolela, 2009).
The determination of modal composition and porosity proportion for 208
samples was conducted through a thin section examination by point counting
400 points per slide.
The rock characteristics of samples were recognized in terms of sorting, texture,
grain size and roundness. The determination of sorting, texture and roundness
types for samples was evaluated from the thin section under the microscope.
Additionally, the measurement of grain size of samples was identified

32

Table 3-1: The general information of studied wells.


Wells

Site ID

Hole Name

Drilled

Easting

Northing

Collar

TD

Type

EAW 18 a

S1959

EC Appin West DDH 18a

19/11/2008 - 29/11/2008

285463.05

6216881.25

310.32

260.49

Core-Non
Coal

EAW 30

S1975

EC Appin West DDH 30

25/02/2009 - 07/04/2009

283293.14

6219366.32

309.98

881.09

Coal Core

EAW 42

S2037

EC Appin West DDH 42

28/10/2009 - 25/1/2010

286735

6216059.43

136.56

682.1

Coal Core

EAW 156

S1996

EC Appin DDH 156

23/03/2009 - 06/04/2009

298772.26

6207843.32

381.65

491.2

Coal Core

EDEN 115

S1934

DC Dendrobium DDH 115

18/06/2008 - 20/06/2008

292128.84

6192392.99

427.51

114.15

Core-Non
Coal

EDEN 124

S2000

DC Dendrobium DDH 124

01/04/2009 - 06/05/2009

290161.39

6191011.18

441.98

437.15

Coal Core

EDEN 125

S2001

DC Dendrobium DDH 125

23/04/2009 - 07/05/2009

288462.57

6192020.03

413.88

431.5

Coal Core

EDEN 126

S2002

DC Dendrobium DDH 126

22/04/2009 - 07/05/2009

288633.38

6194222.09

399.97

466

Coal Core

EDEN 127

S2003

DC Dendrobium DDH 127

05/05/2009 - 27/05/2009

290571.12

6192478.04

409.42

419.4

Coal Core

33

by measuring 50 grains per thin section. The recognition of diagenetic features


in the sandstone was also conducted through thin section analysis.

3-3-2 X-ray diffraction analysis


About 258 samples were prepared for X-ray diffraction analysis using a Philips
(PW3710) diffractometer (Cu K radiation, 35 kV, 28.5 mA) to determine the
percentage of each mineral in fine samples, and clay minerals in the sandstone
(oriented samples of < 2m clay fractions; cf. Lima and De Ros, 2002; Schmid
et al., 2004; Zhang et al., 2008; Gier et al., 2008). Furthermore, traces, UPDSM
and SIROQUANT software were used to examine resultant diffraction traces for
the purposes of determining the presence and abundance of minerals.

3-3-3 Scanning electron microscope (SEM)


A JEOL JSM T330 scanning electron microscope (SEM) was used to examine
about 84 samples to determine morphology, textural relationships, mineral
composition, porosity, permeability and diagenetic aspects of the sandstone
samples (Wolela, 2009).

3-4 Well logs


Many studies have used well log data to describe porosity (e.g. Kok et al., 2005)
and permeability (e.g. Mohaghegh et al., 1997; Wong, 1999; Bruce and Wong,
2002; Majdi et al., 2010).
Mohaghegh et al. (1997) used well log data to study permeability; depending on
three methods using empirical models, multiple variable regression and virtual
measurement. Also, Wong (1999) used an improved windowing technique
which is effective to describe permeability from well logs. He compared between
conventional well logs and core permeability data in his study. Bruce and Wong
(2002) used an evolutionary neural network as a new method to study
permeability in well logs. Kok et al. (2005) used well logs as the main machine
to study porosity and determine the characteristics of reservoirs.
34

Well logs including gamma ray and neutron logs are important in identification
of depositional sequences (Herbert, 1995). Klett et al. (2002) used well logs to
introduce an important study from facies in the Tertiary Lower Rhine Basin fill.
The features of reservoir analogues are studied by gamma ray logs from
outcrops (Slatt et al., 1992; 2000). Also, the study and description of subsurface
well log correlations depends on correlation with gamma ray logs from outcrop
(Slatt et al., 1992). The determination of lithology, thickness and stacking
patterns in the subsurface sections can be completed through the use of
gamma ray, density and porosity wire line logs (Yang, 2007). Also, spontaneous
potential and resistivity can achieve the same purpose (Yang, 2007). The
hydraulic conditions are studied in boreholes by using flowmeter logs in
conjunction with geophysical logs (Runkel et al., 2006). Evans et al. (2007)
described the stratigraphic framework, sedimentary rocks and sand body
architectures of the Tumblagooda Sandstone by outcrop gamma ray logging.
Also, grain size, geometry, lithology and thickness and also amount of shale in
sedimentary formations are important characteristics, identified by gamma ray
well log (Evans et al., 2007).
In this study, well logs such as gamma ray, calliper, density, density porosity,
neutron porosity and sonic were used (Appendix 4). They are important in the
study of lithology, facies and reservoir quality. In this study, gamma ray well
logs were prepared to determine the facies types and depositional environment.
Density porosity was determined by well logs and compared with thin section
porosity and helium porosity.

3-5 Porosity and Permeability measurements


Twenty four samples from the Illawarra Coal Measures, Narrabeen Group and
Hawkesbury Sandstone were analysed using helium injection porosity, air
permeability and density determinations. This study was completed in
Weatherford Laboratories (Figs 3-1 and 3-2).

35

The test and calculation procedures for samples depends on the sample
preparation, sample drying and base parameters. These samples consisted of
1 inch diameter core plugs and were trimmed to create right cylinders
appropriate for permeability determinations. The samples were dried in an oven
at 60 C and 40% relative humidity. They were cooled to room temperature in
an airtight chamber.
The porosity was calculated as the volume percentage of pore space with
respect to the bulk volume (Fig. 3-1).
P 1 V1 = P 2 V2
P1 Vr = P2 (Vr + Vc Vg)
Vp = Vb - Vg
Ambient porosity % = Vp/Vb 100% where
P1 = initial pressure (psig)
P2 = final pressure (psig)
Vr = reference cell volume (cm)
Vc = matrix cup volume (cm)
Vg = grain volume (cm)
Vp = pore volume (cm)
Vb = bulk volume (cm)
The grain density was calculated by dividing the weight of the plug by the grain
volume determined from the helium injection porosity measurement.
= Wt/Vg where
P = grain density (g/cm)
Wt = weight of sample (g)
Vg = grain volume (cm)

36

Figure 3-1: Porosimeter schematic.


Air permeability for each core sample was calculated using Darcys Law through
knowledge of the upstream pressure, flow rate, viscosity of air and sample
dimensions (Fig. 3-2).
Ka = 2000.BP..q.L/ (P1 P2).A
Ka = air permeability (milliDarcys)
BP = barometric pressure (atmospheres)
= gas viscosity (Cp)
q = flow rate (cm/s)
L = sample length (cm)
P1 = upstream pressure (atmospheres)
P2 = downstream pressure (atmospheres)
A = sample cross-sectional area (cm2)
37

Figure 3-2: Gas permeameter schematic hassler.

38

Chapter Four
Environment of deposition in the southern
Sydney Basin
4-1 Introduction
Sedimentary environments are parts of the Earths surface that are
characterized by particular biological, physical and chemical conditions. In
general, three broad depositional environments were recorded by Boggs
(2001): continental, marginal-marine and marine. Continental environments
include fluvial, desert, lacustrine, glacial and deltaic. Marginal-marine
environments include beach/barrier island, estuarine/lagoonal, tidal flat and
neritic zones. Marine environments include all deeper oceanic zones (Boggs,
2001).
The term facies is common in geology and must be used to interpret
depositional environments in sedimentary rocks (Nichols, 2009). Lithofacies are
recognized based on their sedimentary characteristics. Distinguishing between
lithofacies depends on many characteristics including grain size, texture,
composition and sedimentary structures. Thus, depositional environments can
be determined from lithofacies in many studies. However, different depositional
environments can include the same lithofacies. Lithofacies codes are shown in
Appendix 1.1 modified after Miall 1977, 1978a; from Dehghani, 1994). Facies
systems for the Sydney Subgroup were determined by Kennedy (1999;
Appendix 1.2).
This chapter is subdivided into three parts: (a) previous work on depositional
environments and facies in the southern Sydney Basin, (b) facies analysis in the
present study, and (c) the comparison between previous work and the present
study.

39

4-2 Previous work


4-2-1 Illawarra Coal Measures
A sequence stratigraphic framework for the Illawarra Coal Measures and upper
Shoalhaven Group was provided by Arditto (1987a, 1987b, 1991) and Herbert
(1995). Arditto (1991) showed a sequence stratigraphic subdivision of the Late
Permian strata in the Southern Coalfield, and he indicated six successions or
cycles of Late Permian age that are present in the Southern Coalfield. A
similarity is observed between cycles 3, 4 and 5 in Arditto (1991) and
sequences D, E/F and G of Herbert (1995) who also studied the sequence
stratigraphy of the Late Permian coal measures.

4-2-1-1 Cumberland Sub Group


Bamberry et al. (1989) showed that a Deltaic System A and a Shallow Marine
System were responsible for deposition of the Cumberland Subgroup.
4-2-1-1-1 Pheasants Nest Formation
Bowman (1974) interpreted the coal-bearing sediments of the Pheasants Nest
Formation as fluvial deposits in a deltaic setting. The Pheasants Nest Formation
is characterised by four major alluvial environments and sedimentary structures
as shown in Appendix 1.3 (Bowman, 1980). However, Bamberry et al. (1989)
indicated that volcanolithic conglomerate and sandstone are present in the
Pheasants Nest Formation (Deltaic System A). Also, coal beds and a low
percentage of sandstone are recorded in the succession to the north. The
Pheasants Nest Formation includes a transverse river system (braidplain). The
Gerringong Volcanics is the source of braidplain sediments (Bamberry et al.,
1989). Thus, Deltaic System A was determined to start from the top of the
shallow marine upper Shoalhaven Group and extend into the Pheasants Nest
Formation (Bamberry et al., 1995).
4-2-1-1-2 Erins Vale Formation
Moderately sorted, flat bedded and burrowed, fine- to medium-grained
sandstone, deposited in bays and lagoons, is a common feature in the Erins
40

Vale Formation according to Bowman (1980; Appendix 1.3). In contrast,


Bamberry et al. (1989) interpreted that the bulk of the Erins Vale Formation
represents a shallow marine system that characterizes the upper sediments of
Deltaic System A. However, the shallow marine sandstone in the upper Erins
Vale Formation and fluvial sandstone in the overlying basal Wilton Formation
and Woonona Coal Member represent fluvial/strandplain system deposits. Well
sorted sandstones are recorded in the upper Erins Vale Formation and are
considered to be foreshore deposits (Bamberry et al., 1989). In the northern
part of the coalfield the Erins Vale Formation and Kulnura Marine Tongue were
mainly deposited in a shallow marine system (Bamberry et al., 1995).

4-2-1-2 Sydney Sub-group


The palaeoenvironments of the Sydney Subgroup proposed by Bowman (1980)
are shown in Appendix 1.4. The Sydney Subgroup was deposited mainly in a
fluvial environment on a distributive delta floodplain, indicated by grain size
parameters, sedimentary structures and vertical profiles (Bowman, 1980).
4-2-1-2-1 Wilton Formation
Bunny (1972) and Bowman (1974) interpreted the Wilton Formation as
floodplain deposits associated with a fluvial environment. Jones (1983)
described environments of deposition in the Sydney Subgroup located between
Wollongong and Stanwell Park, and indicated that the deposition of the upper
Wilton Formation was in a very shallow lacustrine or lagoonal environment.
Moreover, the deposits of the Wilton Formation were said to characterize the
lower part of Deltaic System B (Bamberry et al., 1989, 1995). Bamberry et al.
(1995) showed that dark grey claystone and siltstone are present in the Wilton
Formation along with interbedded sandstone. They indicated that the presence
of coarsening upwards successions, sideritic concretions, plentiful bioturbation,
fossil drift-vegetation and thin laterally persistent carbonaceous intervals are
recognised characteristics of the Wilton Formation. Also, the fine grain size,
conservation of the thin tuffaceous laminae, and presence of bioturbation reflect
placid sedimentation during deposition of much of the Wilton Formation.
Additionally, the preservation of widespread coarsening upwards sequences
41

combined with more restricted fining upwards sandstone bodies indicate


shallowing of the depositional environment. Nonetheless, deposition of
suspended sediment by flocculation is indicated by the claystone to siltstone
bay-fill deposits occurring between distributary sandstones in the Wilton
Formation. Interdistributary lakes are rare in the Wilton Formation (Bamberry et
al., 1995).
4-2-1-2-2 Tongarra Coal
The Tongarra Coal formed the upper part of Deltaic System B (Bamberry et al.,
1989, 1995). Banded bright and dull coal interbedded with siltstone,
carbonaceous claystone and tuff are observed in the Tongarra Coal.
Interdistributary bay and channel sequences are indicated by strata that
represent delta top deposits present beneath the Tongarra Coal. The coal also
includes tuffaceous bands, which are attributed to ash falls deposits (Bamberry
et al., 1995).
Thus, the sequence from the top of the Woonona Coal Member to the Tongarra
Coal was deposited as delta front, delta margin and lower delta plain deposits
(Bamberry et al., 1995).
4-2-1-2-3 Bargo Claystone
The deposits of meandering fluvial and floodplain environments were
interpreted for the sequence from the Bargo Claystone to the Allans Creek
Formation (Bunny, 1972; Bowman, 1974). The same succession was
interpreted as the Deltaic System C in the Sydney Subgroup that directly
overlies the Tongarra Coal (Bamberry et al., 1989, 1995). This deltaic system
starts with predominantly Fl lithofacies, and rarely Sh lithofacies, in the Bargo
Claystone. The Fl lithofacies consists of interbedded claystone and siltstone
whereas medium-grained sandstone beds are included in the Sh lithofacies.
Deposition from suspension in a calm environment is indicated by the Fl
lithofacies (Roche, 1997). In the Bargo Claystone, sedimentation is rapid
because of flocculation, which is associated with the convergence of the fresh
water and seawater with the former carrying fine-grained sediment in
suspension (Bamberry, 1992).
42

Late Permian deposits are characterised by seasonally cold conditions that are
indicated by dropstones. Also, seasonal differences in sediment supply are
shown by grain size variations in the Fl lithofacies (Bamberry, 1992).
Deposition of the Bargo Claystone was primarily in an interdistributary bay or
prodelta setting (Roche and Hutton, 1998). Bamberry et al. (1995) showed that
parallel laminated claystone and siltstone, sideritic bands, tuffaceous laminae
and the sporadic distribution of bioturbation are characteristics that indicate
deposition of the Bargo Claystone in a placid prodelta setting.
4-2-1-2-4 Darkes Forest Sandstone
The Darkes Forest Sandstone contains sandstone and fine-grained lithofacies,
plentiful finely comminuted plant debris and plentiful bioturbation and
synaereses cracks in shale beds. Also, fine- to medium-grained Sp, Sr and Sh
facies are widespread in the Dark Forest Sandstone and are interbedded with
composite and finer grained lithofacies (Fl; Bamberry et al., 1995). As a result of
crevasse splay deposits, the Sh lithofacies are stacked in the east and
deposition of the Darkes Forest Sandstone was recognised as a river mouth
setting by Roche (1997).
4-2-1-2-5 Allans Creek Formation
Two carbonaceous intervals with sequences of claystone-siltstone are
described in the Allans Creek Formation according to Bamberry et al. (1995).
Bamberry et al. (1989) showed that interdistributary lakes and bays, crevasse
subdeltas, distributary channels and swamps are the dominant facies in the
coastal plain sediments forming the Allans Creek Formation, which were
deposited behind the prograding river mouth successions of the Darkes Forest
Sandstone (Bamberry et al., 1989). Fl and C lithofacies are widespread in the
Allans Creek Formation; thus the deposition occurred in a calm environment
(Roche, 1997). According to Roche (1997) and Roche and Hutton (1998) this
fine-grained formation was deposited in a poorly drained floodplain setting
accompanied by coaly intervals.

43

4-2-1-2-6 American Creek Coal Member


This member at the top of the Allans Creek Formation is dominated by coal with
thin interbeds of carbonaceous claystone (Bamberry et al., 1995). Its deposition
occurred in a poorly drained floodplain setting according to Roche (1997) and in
shallow lakes according to Bamberry et al. (1995).
4-2-1-2-7 Kembla Sandstone
The sequence from the base of the Kembla Sandstone to the Wongawilli Coal
represents upward fining from a fluvial channel to a coal swamp environment
(Jones, 1983). Bamberry et al. (1995) showed that the Kembla Sandstone
included coarse and fine members as the main lithofacies. Coarse members
consist of St, Sp and Sh lithofacies. The St lithofacies are widespread while Sp
and Sh lithofacies are present in small amounts. The coarse members contain
predominantly medium- to large-scale channel-fill characteristics. The fine
members include common fine-grained lithofacies (Fl, Fsc), which are also
present as lenses and interbeds in the coarse members but are less common.
Sedimentary structures such as ripple cross-lamination, dewatering structures
and horizontal lamination are observed in the fine members. Additionally, the
fine members are characterised by rootlets, mats of fossil drift vegetation and
vitrinous coal fragments. A lobate distribution is indicated by the coarse member
that was created by in-channel processes and lateral migration. Floodplain
sequences are represented by the fine member and comprise levee and
crevasse splay deposits, abandoned channel fill and floodplain lake
successions. The deposition of floodplain sequences occurred by vertical
accretion of fine-grained sediment (Bamberry et al., 1995) while deposition of
the whole Kembla Sandstone was by a meandering fluvial system (Roche and
Hutton, 1998).
4-2-1-2-8 Wongawilli Coal
Carbonaceous claystone lenses and tuff bands are observed in the upper and
lower parts of the Wongawilli Coal (Bamberry et al., 1995) while the coal is
associated with Sx channel lithofacies in drill hole New Belanglo 30 (Roche,
1997). Jesus (1999) recorded that the Wongawilli Coal is widespread and it
44

contains high concentrations of vitrinite (Shibaoka and Smyth, 1975). A


distinctive thick tuff bed in the lower Wongawilli Coal was termed the
Farmborough Claystone Member by Bamberry et al. (1995).
4-2-1-2-9 Lower Eckersley Formation
The lower section of the Eckersley Formation includes the Novice Sandstone
Member, Woronora Coal Member, Hargrave Coal Member, Cape Horn Coal
Member and Burragorang Claystone Member (Bamberry et al., 1995).
Bamberry et al. (1995) showed that thin interbeds of very fine-grained
sandstone and siltstone are prominent in the Novice Sandstone. Thus,
deposition of the Novice Sandstone was attributed to meandering distributary
channels. Also, laminated carbonaceous shale and siltstone, coal and tuff are
recorded in the Woronora Coal Member. The depositional environment for this
unit was in a lacustrine setting. Plentiful drift-vegetation and synaereses cracks,
sparse burrows and sedimentary structures were recorded in the lower
Eckersley Formation (Bamberry et al., 1995).
The Burragorang Claystone Member commonly occurs as a massive bed or
medium beds characterised by very thinly laminated tuffaceous claystone that
shows well developed graded bedding. Stylolites, rill marks and symmetrical
ripple marks are also observed in this unit (Bamberry et al., 1995). The
Burragorang Claystone Member represents an air-fall ash deposit (Bamberry et
al., 1995; Grevenitz et al., 2003). Partial reworking of the Burragorang
Claystone Member occurred in a shallow water environment (Bamberry et al.,
1995). Bamberry and Doyle (1987) interpreted that the Burragorang Claystone
Member was deposited in a shallow lacustrine setting on a widespread
floodplain near to a meandering river system, which is shown by some
sedimentary structures important in revealing the low energy currents of this
environment, for example, sedimentary structures such as graded bedding
(Bamberry and Doyle, 1987).

45

4-2-1-2-10 Upper Eckersley Formation


The upper Eckersley Formation comprises the Lawrence Sandstone and
Balgownie Coal Member (Bamberry et al., 1995).
The Lawrence Sandstone consists of Sh, Sx, Fl and Fx lithofacies. The basal
unit includes Sh and Sx lithofacies. Sh and Fl lithofacies are present in the
middle unit while Fl lithofacies dominate the uppermost unit (Roche, 1997). The
Balgownie Coal Member is composed of banded coal and dull coal. Also, noncoal beds are present in this unit and include massive to laminated
carbonaceous claystone and siltstone (Bamberry et al., 1995). Deposition of the
Lawrence Sandstone occurred in meandering fluvial system (Roche and Hutton,
1998) giving way to floodplain and swamp environments for the Balgownie
Coal.
4-2-1-2-11 Loddon Sandstone
The Loddon Sandstone includes Sh, Sx, Fx and Fl lithofacies. The Sh and Sx
lithofacies are widespread in the basal unit and are interpreted as channel-fill
deposits. Their thickness is up to 4 m in the east. The middle unit is
characterized by Fx and Fl lithofacies while Fl lithofacies are recorded in the
part of the upper unit (Roche, 1997). Deposition of the Loddon Sandstone
occurred as a fining upward succession in a meandering fluvial system (Roche
and Hutton, 1998).
4-2-1-2-12 Bulli Coal
Jones (1983) showed that the Bulli Coal contains carbonaceous siltstone and a
thick well banded coal seam. Deposition of the Bulli Coal was in a very broad
shallow swamp or lacustrine environment, indicated by the presence of many
bands in the coal seam (Jones, 1983). Deposition of the Bulli Coal occurred as
an unconfined peatland coal over a large area of the Southern Coalfield and it
blankets the underlying strata (Johnson, 1995). Non-coal beds occur in the Bulli
Coal as carbonaceous claystone and the coal is more than 3 m thick in the
Campbelltown to Camden area (Bamberry et al., 1995). Roche and Hutton
(1998) indicated that deposition of the Balgownie Coal and Bulli Coal mainly
occurred during regional tectonic-controlled subsidence. The roof rocks of the
46

Bulli seam are clastic lacustrine, floodplain and fluvial sediments (Diessel,
1966).

4-2-2 Narrabeen Group


Loughnan et al. (1964) indicated that deposition of the red beds in the
Narrabeen Group occurred in a piedmont environment. Hamilton and Galloway
(1989) interpreted that the Narrabeen Group was deposited in a fluvio/lacustrine
environment whereas, according to Herbert (1993a,b, 1996, 1997), deposition
of the Narrabeen Group occurred on an alluvial plain.

4-2-2-1 Clifton Sub-Group


4-2-2-1-1 Coalcliff Sandstone
According to Deen (1999), the Coalcliff Sandstone consists of a gravel facies
assemblage, a sandstone facies assemblage and a fine-grained facies
assemblage. Gm lithofacies are widespread in the gravel facies assemblage
while Gms and Gis lithofacies are rare. The thickness of Gm lithofacies reaches
up to 1.2 m. The sandstone facies assemblage comprises Sh, Se, Sm, Sr, St
and Ss lithofacies. The St lithofacies have thicknesses ranging between 10 and
50 cm. The Sr lithofacies are present with St, Fl and Fm lithofacies and have a
maximum thickness of 40 cm. The fine-grained facies assemblage includes Fl,
Fsc, Fib, Fm and C lithofacies. Fl lithofacies are most widespread in the finegrained facies assemblage. The Fl lithofacies have thicknesses ranging from 2
mm to 1 cm. Suspended load deposition on a floodplain environment is
characterized by the Fl lithofacies. The deposition of Fm lithofacies occurred in
proximal parts of the floodplain and on bar-tops. The deposition of Fib
lithofacies was in shallow pond-like environments in a floodplain setting.
Ward (1972), Jones (1986), Hamilton et al. (1987), Reynolds (1988), Dehghani
and Jones (1994b) and Deen (1999) suggested a fluvial origin for the Coalcliff
Sandstone. In addition, Dehghani and Jones (1994b) showed that the Coalcliff
Sandstone was deposited in a fluvial environment based on: fining-upward
channel-fill deposits; closely spaced erosional surfaces present in multistorey
sandbodies; channel-floor scouring and associated lag deposits; a unidirectional
47

palaeocurrent

for

individual

mesoforms

and

macroforms

shown

by

palaeocurrent data; occurrence of overbank deposits with crevasse splays; lack


of body fossils and paucity of trace fossils; and plentiful coalified and petrified
logs and plant fragments.
4-2-2-1-2 Wombarra Claystone
Dehghani (1994) showed that planar cross-bedded sandstone (Sp), ripple
cross-laminated sandstone (Sr), plane bedded sandstone (Sh), massive
sandstone (Sm), parallel laminated mudstone and fine-grained sandstone (Fl),
massive mudstone (Fm), ironstone bands (Fib) and palaeosols (P) are present
in the Wombarra Claystone as lithofacies. The Sp and Sr lithofacies have
thicknesses varying between 10 cm and 40 cm and characterise crevasse splay
deposition in a proximal floodplain environment. The thickness of Sh and Sm
beds ranges from 10 cm to 40 cm (Dehghani, 1994). The Fl lithofacies are
widespread and have thicknesses reaching up to 2.5 m. The presence of the Fl
lithofacies characterise vertically accreted deposits in a floodplain environment.
The Fl lithofacies were deposited from suspension. The presence of the Fl
lithofacies, associated with Sp and Sr lithofacies, characterize proximal
floodplain deposition. The Fm lithofacies is common and indicates deposition in
a distal floodplain or flood basin environment. The Fib lithofacies are observed
varying in thicknesses between 1 cm and 10 cm (Dehghani, 1994). Additionally,
carbonaceous mudstone (C) and laminated to massive siltstone and mudstone
(Fsc) are present in the Wombarra Claystone (Deen, 1999). The Wombarra
Claystone was deposited as a floodplain deposit (Dehghani, 1994; Deen, 1999).
4-2-2-1-3 Scarborough Sandstone
A fluvial depositional environment for the Scarborough Sandstone was indicated
by Ward (1972), Bunny (1972), Bowman (1974), Hamilton et al. (1987) and
Dehghani and Jones (1994c). This accounts for the conglomerate, sandstone
and fine-grained facies assemblages that were determined in the Scarborough
Sandstone by Dehghani and Jones (1994c). The conglomerate facies
assemblage contains massive to crudely bedded conglomerate (Gm), planar
cross-bedded conglomerate (Gp), trough cross-bedded conglomerate (Gt) and
scour fill conglomerate (Ge). Furthermore, Gm lithofacies are widespread in the
48

Scarborough Sandstone while Gt lithofacies are rare. The Gp lithofacies are


present with thicknesses varying between 0.5 m and 1.2 m. Sandstone facies
assemblages are dominant in the Scarborough Sandstone and are more
common than conglomerate or fine-grained facies assemblages. Four facies are
recognised in the sandstone facies assemblage, composed of planar crossbedded sandstone (Sp), ripple cross-laminated sandstone (Sr), trough crossbedded sandstone (St) and plane bedded sandstone (Sh). The Sp lithofacies
have thicknesses varying between 10 cm and 1.3 m. The St lithofacies are rare
and has a thickness ranging from 20 cm to 50 cm. The Sh lithofacies are
present in thicknesses between 20 cm and 50 cm. The Sr lithofacies are rare
and comprise non-climbing ripple types (Dehghani and Jones, 1994c). Finegrained facies assemblages constitute 0 - 10% of the formation and are rare in
the Scarborough Sandstone. It includes parallel laminated mudstone, finegrained sandstone (Fl) and massive mudstone (Fm; Dehghani and Jones,
1994c). Thus, the Scarborough Sandstone was deposited as bed load fluvial
deposits (Hamilton et al., 1987; Dehghani and Jones, 1994c; Deen, 1999).
4-2-2-1-4 Stanwell Park Claystone
Red brown and grey green massive to poorly laminated claystone, and a few
beds of friable granule conglomerate are recognised in the Stanwell Park
Claystone (Ward, 1972). The granule conglomerate beds include rounded shale
clasts. According to Deen (1999), the Stanwell Park Claystone includes gravel,
sandstone and fine-grained facies assemblages. The gravel facies assemblage
includes Gms lithofacies. Sedimentary structures are missing in the gravel
facies assemblage. Deposition of these gravels occurred as lag deposits in local
crevasse channel environments. The sandstone facies assemblage has a
thickness range from 10 cm to 1.8 m and includes volcano-lithic fragments.
They represent subordinate channels in the floodplain environment. However,
the Fm, Fsc and Fl lithofacies are widespread in the fine-grained facies
assemblage and they represent deposition from suspension. Fm lithofacies are
the most common and have thicknesses reaching up to 3 m. The distal
floodplain depositional environment is indicated by thick beds of Fl facies. The
Fl lithofacies are associated with Sr and Sh facies indicating deposition in the
proximal parts of a floodplain environment (Deen, 1999). Dehghani (1994)
49

indicated that palaeosols (P) are also recorded in the Stanwell Park Claystone
as a lithofacies. He interpreted that the Stanwell Park Claystone was deposited
as a floodplain deposit.
4-2-2-1-5 Bulgo Sandstone
Ward (1980) indicated that the Bulgo Sandstone in coastal regions contains
three distinctive facies: pebbly facies, volcanic facies and shaly facies. Friable
granule conglomerate and coarse volcanic sandstone are recorded in the
volcanic facies. He noted that the Bulgo Sandstone consists of fluvial deposits.
Deen (1999) described the lithofacies in the lower, middle and upper Bulgo
Sandstone. The lower Bulgo Sandstone includes gravel and sandstone facies
assemblages. The gravel facies assemblage is characterised by Gm, Gms and
Gld lithofacies. The Gm lithofacies has a thickness of up to 0.7 m. The
sandstone facies assemblage consists of St, Sp, Sm, Sr and Sh lithofacies. The
St lithofacies (75 cm thick) are common and represent deep channel deposits.
The Sm lithofacies are characterised by thicknesses ranging from 20 cm to 1 m.
Sp and Sh lithofacies occur in thicknesses of up to 1.1 and 1 m, respectively
(Deen, 1999). Moreover, mud and clay intraclasts are prevalent in the
sandstone facies and were interpreted to result from bank failure because of the
deep channel system (Dehghani, 1994). Sr and Sh lithofacies represent bar top
or crevasse splay deposits in a floodplain environment (Deen, 1999). Based on
gravel and sandstone facies assemblages, the lower Bulgo Sandstone was
deposited as bed-load fluvial deposits (Dehghani, 1994; Deen, 1999).
The middle Bulgo Sandstone includes gravel, sandstone and fine-grained facies
assemblages. The gravel facies assemblage contains Gms and Gld lithofacies
and both are present as thin layers (Deen, 1999). Gms lithofacies were
interpreted as debris flow deposits by Miall (1978a, b). On the other hand, Sp,
St, Sl, Sh, Sm and Sr lithofacies are recorded in the sandstone facies
assemblage. Sh lithofacies are the most common in the sandstone facies
assemblage. The fine-grained facies assemblage includes Fl and Fm
lithofacies. Fl lithofacies has a thickness reaching up to 50 cm and represents
deposition from suspension in an overbank setting (Deen, 1999). The coarser
material overlies fine-grained lithologies, which were reworked by channels. The
50

coarser material was transported onto the floodplain during large floods
(Dehghani, 1994). In general, based on the common occurrence of Sp and St
lithofacies, the middle Bulgo Sandstone represents a braidplain deposit (Deen,
1999).
The upper Bulgo Sandstone includes sandstone and fine-grained facies
assemblages. St, Sp, Sh and Sr lithofacies are present in the sandstone facies
assemblage. The Fl lithofacies is widespread in the fine-grained facies
assemblage. The Fl lithofacies is observed as thin layers with thicknesses
varying between 10 and 50 cm (Deen, 1999). The upper Bulgo Sandstone
represents a mixed-load fluvial deposit (Dehghani, 1994; Deen, 1999).
4-2-2-1-6 Bald Hill Claystone
The red-brown Bald Hill Claystone sediments indicate a lacustrine environment
deficient in organic material (Loughnan, 1962). The Bald Hill Claystone red
beds were also interpreted as marine bay deposits by Conolly (1969). Goldberry
and Holland (1973) suggested that the Bald Hill Claystone was deposited in a
piedmont deltaic setting. Hamilton et al. (1987) showed that the red kaolinitic
mudstone facies are a common characteristic in the Bald Hill Claystone. In
addition, channel-fill and channel-margin deposits are the other facies in this
unit. The occurrence of channel-fill deposits is determined by discrete finingupward sandstone bodies and as amalgamated sandstone units. Also, the bed
load deposits include medium- to very coarse-grained, dominantly quartzose
sediment. Levee and crevasse splay facies are recognised in the channel
margin deposits, which contain intercalated siltstone and fine-grained quartzrich sandstone (Hamilton et al., 1987). Dehghani (1994) indicated that the
lithofacies identified in the Bald Hill Claystone include sandy conglomerate
(Gm), fretted sandstone (Sm and Sh up to 3 m thick), dark grey laminated shale
(Fl), light grey claystone (Fm) and palaeosols (P). Thus, the Bald Hill Claystone
was deposited as a floodplain deposit (Dehghani, 1994), probably in a welldrained floodplain environment based on the presence of Fm, Fsc and Fl
lithofacies (Deen, 1999). However, the Gm, Sh and Sm lithofacies indicate
floodplain channel and crevasse splay deposits (Deen, 1999).

51

4-2-2-2 Gosford Sub Group


4-2-2-2-1 Garie Formation
Baker (1956), Loughnan (1969, 1970), Ward (1971 a, b), Goldberry and Holland
(1973), Loughnan et al. (1974), Dehghani (1994) and Deen (1999) have all
studied the Garie Formation. Recognised lithofacies in the Garie Formation are
massive tuff with sparse accretionary lapilli, accumulated accretionary lapilli,
and brecciated claystone (Dehghani, 1994; Deen, 1999). Massive tuff with
sparse accretionary lapilli characterizes an ash-fall deposit (Deen, 1999). The
accumulated accretionary lapilli vary in size between 1 mm and 6 mm
(Dehghani, 1994). In this unit, a series of similar beds were noted by Bunny and
Herbert (1971), and vary in thickness between 15 and 30 cm. Breccia is rare in
the Garie Formation and indicates partial reworking of these tuffaceous
materials after deposition (Dehghani, 1994).
4-2-2-2-2 Newport Formation
Osborne (1948), Crook (1957), Branagan (1969) and Bunny and Herbert (1971)
studied the characteristics of the Newport Formation. Branagan et al. (1966)
and Branagan (1969) studied the equivalent Gosford (Newport) Formation
between Palm Beach and Long Reef to the north of Sydney. The results
obtained were not detailed. However, they indicated successions of thinly
bedded and laminated sediments. Between these beds, and at times erosive
into them, are zones of cross-bedded units, scour-and-fill deposits and
sandstone channels (Branagan et al., 1966; Branagan, 1969). Also, they
determined four breaks in sedimentation. Bunny and Herbert (1971) interpreted
that a mixture of deltaic swamps, interdistributary bays and sub-aerial levee
facies are represented in the lower shaly part of the Newport Formation. The
upper part of the Newport Formation is characterised by fluvial sedimentation
and includes point-bars. In the upper part of the Gosford Formation in the
Bouddi area, a similar succession was present and has eight main sandy units,
which were interpreted by McDonnell (1969) as fluvio-deltaic, with interruptions
and vacillations occurring between associated facies. At North Head, Herbert
(1980b) noted that shale in the Newport Formation contained a plentiful variety
of sedimentary structures suggesting complex lower deltaic facies. However,
52

laminae characterise the uppermost shale bed of the Newport Formation and
include finely interbedded, very fine- to fine-grained sandstone and siltstone.
Other

common

characteristics

are

flaser

bedding,

micro-cross-bedded

sandstone laminae, load casting, ball-and-pillow structures and bioturbation. In


addition, in the Newport Formation, some sandstone consists of widespread
cross-beds but rarely massive sandstone (Herbert, 1980b). These sandstones
are similar to the sheet and massive sandstone facies that were determined by
Conaghan and Jones (1975) in the Hawkesbury Sandstone. These sandstones
are distributary mouth bars (Packham, 1976). In a more recent study, Deen
(1999) indicated that the lithofacies in the Newport Formation contain gravel,
sandstone and fine-grained lithofacies. Gravel lithofacies include conglomerate
that represents periodic channel involvement in the environment. Sandstone
lithofacies consist of symmetrical ripple marks and rare planar and trough crossbedding. The sandstone characterizes deposition by high-energy events. Finegrained lithofacies include predominantly shales and mudstones. The welllaminated mudstones are widespread and deposited in shallow subaqueous
environments. The deposition of dark grey shales occurred in a poorly drained
environment (Deen, 1999). Deposition of the Newport Formation was recorded
as lagoonal deposits according to Dehghani (1994) and Deen (1999).

4-2-3 Hawkesbury Sandstone


4-2-3-1 Lithofacies
Three dominant facies assemblages have been determined in the Hawkesbury
Sandstone massive sandstone facies, stratified sandstone facies and
mudrock facies (Rust and Jones, 1987; Miall and Jones, 2003).
4-2-3-1-1 Massive sandstone facies
Massive sandstone facies in the Hawkesbury Sandstone is based on its gross
aspect (Conaghan and Jones, 1975). Massive sandstone is widespread in the
Hawkesbury Sandstone, occurring in structureless to faintly parallel-stratified
sandstone beds (Rust and Jones, 1987; Miall and Jones, 2003). The presence
of the structureless and faintly laminated sandstone is usually recorded above
an erosion surface (Jones and Rust, 1983).
53

Also, the sorting of massive sandstones is described as poor to very poor and
these units are generally medium- to coarse-grained according to Jones and
Rust (1983). Granules and pebbles of quartz are present in the massive
sandstone (Jones and Rust, 1983). These components increase towards the
base. Scattered carbonized leaves are also recorded in the massive sandstone.
The massive sandstone commonly contains angular blocks of shale, mudstone,
siltstone and interbedded siltstone/fine sandstone. These are more common at
the base and top (Rust and Jones, 1987).
Mudstone intraclasts are common and range in length up to 5 m (Jones and
Rust, 1983) and 10 m (Conaghan, 1980). Mudstone intraclasts represent about
3 10% of the units (Jones and Rust, 1983). Finally, the deposition of massive
sandstone facies occurred from bank collapse during falling flood conditions
(Rust and Jones, 1987).
4-2-3-1-2 Stratified sandstone facies
Stacked sets of cross-stratified sandstone are characteristic of the stratified
sandstone assemblage and have a thicknesses varying from < 1 m to > 5 m
(Pells, 2002). Massive sandstone beds, horizontally stratified and ripple-crosslaminated are included in the stratified sandstone assemblage. Massive
sandstone has a thickness of 0.5 m in sheet-like units. Cross-stratal sets occur
in both trough and planar forms (Rust and Jones, 1987).
Trough cross-strata have a thickness range from 10 cm to 1.5 m. Individual sets
include sharply curved bases. Also, the preservation of convex-downstream
cross-strata indicated that the bedforms were sinuous-crested dunes (Rust and
Jones, 1987). Small-scale trough cross-stratified sandstone facies include
siltstone and fine- to medium-grained sandstone. They are well sorted (Liu et
al., 1996).
The planar sets have a thickness range from 10 cm to 7.5 m. They vary from
planar-tabular to planar-curved (Rust and Jones, 1987). Individual planar sets
extend for many tens of metres (Conaghan, 1980). The presence of planarcurved cross-strata is rare in the Hawkesbury Sandstone (Conolly, 1965;
Conaghan, 1980). Concave-downstream, sinuous and convex plan forms are
54

recorded in the curved-planar sets. Concave-downstream forms are the most


common. Large planar sets are composed of intrasets with reactivation
surfaces. In the Hawkesbury Sandstone, large-scale planar cross-strata often
include soft-sediment deformation (Rust and Jones, 1987). The large-scale
planar/tabular cross-stratified sandstone facies consists of moderately to well
sorted, fine- to very coarse-grained sandstone. Foresets have high angles
exceeding 20 (Liu et al., 1996).
The Hawkesbury Sandstone commonly includes small- to medium-scale sets of
planar cross-strata with thickness ranges from 10 cm to 1 m (Rust and Jones,
1987). The small-scale planar/tabular cross-stratified sandstone facies are fineto coarse-grained sandstone and are well sorted. Foresets have a thickness
rarely exceeding 1 m and are characterized by dip angle that is more than 20
(Liu et al., 1996).
Low-angle stratified sandstone facies are trough or planar cross-sets and are
very well sorted. This facies has low angle dips between 5 and 20 (Liu et al.,
1996).
4-2-3-1-3 Mudrock facies
Mudrock facies are minor in the Hawkesbury Sandstone. Standard (1969)
showed that the thickness of units of the mudrock assemblage reach up to 12
m. Near Sydney, he noted a 9 m thick mudrock unit extends for approximately
2.5 km. In another study, Herbert and Uren (1972) studied a mudrock about 35
m thick. Rust and Jones (1987) indicated that the mudrock assemblage units
generally have a thickness ranging between 1 and 2 m and commonly are
sheet-like (Rust and Jones, 1987). Ripple cross-laminated, fine-grained
sandstone to siltstone and horizontally laminated fine-grained sandstone to
siltstone and shale are widespread in the mudrock assemblage. These facies
grade into massive mudstone and carbonaceous shale. Plant material is
preserved in ripple troughs by climbing ripples (Rust and Jones, 1987).
Mudrock sequences have preserved small dune bedforms with internal planar
cross-stratification along the Garie-Bundeena coast. The deposits of the
mudrock facies association overlie and underlie these dune bedforms.
55

Bioturbation, load structures and sand-filled cracks are observed in the mudrock
facies assemblage. Bioturbation is uncommon and consists of faunal and root
disturbances. Load structures are prevalent whereas sand-filled cracks are
assigned to desiccation (Rust and Jones, 1987).
Deposition of the mudrock facies was suggested to occur in a lower energy
environment (Conaghan, 1980) and in abandoned channels (Herbert, 1983).
Rust and Jones (1987) interpreted that the mudrock assemblages were
deposited on the floodplain and as abandoned channel fills.

4-2-3-2 Sedimentary structures


Sedimentary structures observed in the Hawkesbury Sandstone include crossbedding and ripple marks. Cross-bedding occurs commonly in the Hawkesbury
Sandstone and has a thickness range from 0.3 to 1.5 m. The presence of ripple
marks is rare in the Hawkesbury Sandstone (Standard, 1969). Slump structures
in the Hawkesbury Sandstone contain contorted stratification, irregularly
contorted bedding and convolute bedding (Standard, 1964). Contorted
stratification is present in some cross-bedded sandstone units and in thin
interbedded shale units and is common in the Hawkesbury Sandstone.
Irregularly contorted bedding is recorded in shale units. Convolute bedding is
prevalent in the highly cross-bedded sandstone units. Other types of
sedimentary structures in the Hawkesbury Sandstone are intraformational
recumbent folds, shale breccias, channels, lenses, scour and fill structures and
sedimentary dykes (Standard, 1964). Other studies indicated sedimentary
structures in the Hawkesbury Sandstone, such as Miall and Jones (2003) who
showed cross-beds, ripples and climbing-ripple cosets.

4-2-3-3 Depositional environment


Conolly (1969) and Conolly and Ferm (1971) suggested that the Hawkesbury
Sandstone was deposited in a marine barrier to tidal delta system. Ashley and
Duncan (1977) proposed that the Hawkesbury Sandstone was deposited in an
aeolian environment. Conaghan and Jones (1975), Conaghan (1980), Jones
and Rust (1983), Rust and Jones (1987) and Miall and Jones (2003) indicated
the deposition of the Hawkesbury Sandstone was by a large braided river,
56

similar to the modern Brahmaputra River. Rust and Jones (1987) showed that
fluvial deposition for the Hawkesbury Sandstone is determined by unidirectional
palaeoflow in the sandstone, scour surfaces and plentiful mudrock intraclasts.
However, a braided fluvial system for the Hawkesbury Sandstone is shown by
low palaeocurrent variance, sheet-like morphology, plentiful erosion surfaces,
and the small amounts of in situ mudrocks.

4-3 Facies analysis


4-3-1 Illawarra Coal Measures
The study of the formation thicknesses and facies descriptions were obtained
from well logs for the drill cores. The average thicknesses of the formations
were determined from all drill cores, including EAW 18a, EAW 30, EAW 42,
EAW 156, EDEN 115, EDEN 124, EDEN 125, EDEN 126 and EDEN 127. The
facies were described from two wells EAW 42 and EDEN 125.

4-3-1-1 Sydney Sub-group


4-3-1-1-1 Wilton Formation
The Wilton Formation is an average of 5.2 m thick in the study area and is
characterised by shale, carbonaceous mudstone, sandstone and siderite bands
(Appendixes 3 and 4).
Floodplain facies: The thicknesses of three sequences of floodplain facies
range from 0.6 to 3.2 m with an average of 1.5 m thick, and they include shale,
carbonaceous mudstone and medium-grained sandstone.
Shale is the most common feature in all parts of the formation, occurring in
sequences between 0.2 and 1.6 m thick at an average of 0.6 m (Appendix 21a). The thickly laminated shale includes silty laminae, sandy wisps in places,
plus some tuffaceous, carbonaceous and sandy beds.
Black carbonaceous mudstone beds are up to 0.3 m thick (finer than the shale
beds; Appendix 2-1a). These beds usually have a diffuse base with the
underlying unit, and are thickly laminated.
57

Grey medium-grained sandstone (Appendix 2-1b) is mainly very thinly bedded


but in the upper part of the formation beds are up to 0.6 m thick. It contains silty
laminae in places and has a diffuse base with the underlying mudstone unit.
In the well logs, both the sandstone and shale units show an increase in gamma
ray and sonic logs and a decrease in resistivity log between 423.3 and 426.6 m
(Well 125, Fig. 4-1). The shale contains common clay minerals indicated by
relatively high gamma ray log. At 422.9, the gamma ray log showed distinct
base which occurs between fine-grained palaeosol deposits in the Wilton
Formation and the overlying Tongarra Coal (Well 125, Fig. 4-1).
Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-1: Well 125 (red lines indicates the boundaries of


the Wilton Formation.
4-3-1-1-2 Tongarra Coal
The thickness of the Tongarra coal is up to 6.4 m and it consists of shale,
carbonaceous shale, coal, calcite, siltstone and sandstone (Appendixes 3 and
4). The formation gradationally overlies the Wilton Formation. Note the main
coal beds had been removed for testing and were not available for facies
descriptions. The carbonaceous shale is very fine- to fine-grained in beds
between 0.01 and 0.1 m thick (average of 0.04 m; Appendix 2-1c). Coal beds
are more important, ranging from 0.01-0.2 m thick with average of 0.1 m
(Appendix 2-1d). Shale units (0.01-3 m thick, average of 0.3 m) are mostly very
fine- to fine-grained and range from black to greyish brown, brownish buff and
brownish grey.
4-3-1-1-3 Bargo Claystone
The Bargo Claystone is 11.4 m thick and includes shale, siltstone, sandstone
and minor siderite (Appendixes 3 and 4).
58

Shallow marine facies: Six facies are present ranging in thickness between
0.9 and 4.6 m at an average of 2.2 m, and they include carbonaceous shale,
coal, sandstone, siltstone and shale.
Brownish buff, brownish grey to grey of shale is distinguished in the middle and
upper parts of the Bargo Claystone (Appendix 2-2a; average of 0.7 m thick).
Shale units are thickly laminated in the Bargo Claystone and have tuffaceous
and mudstone matrix in the middle part of formation whereas sideritic bands,
sandy lenses, muddy lenses and sandy wisps occur up-sequence in the upper
part of formation.
Also, thickly laminated siltstone (Appendix 2-2b) is represented commonly in the
Bargo Claystone in sequences that average 1.1 m in thickness. The lower
contact of this facies shows a distinct base with the underlying claystone. The
siltstone is interbedded with sandstone and tuffaceous claystone bands in the
lower part of formation, whereas sandy wisps, and clayey and sandy lenses
occur in the middle part of the formation. Sideritic bands are characteristic in the
siltstone in the upper part of formation.
Fine-grained sandstone varies between 0.6 and 0.9 m in thickness (average of
0.8 m; Appendix 2-2c). Bedding is very thinly laminated in this facies. Finegrained sandstone includes silty partings in the lower part of formation whereas
in the upper part of formation, this facies includes sideritic and muddy laminae
in places and coarse sandstone bands over the top of shaly lenses.
In well logs, the fine-grained sediments are associated with irregular but higher
gamma ray responses (Well 125, Fig. 4-2). Both gamma ray and sonic logs also
indicate the dominance of shale between 407.8 and 410 m and between 411.5
and 413 m (Well 125, Fig. 4-2). At 415.5, the gamma ray log is very high and
indicates that the shale has a high clay mineral content (Well 125, Fig. 4-2).
Sandstone occurs between 410.3 and 411.5 m with a blocky shaped gamma
ray response (Well 125, Fig. 4-2). At 407.77 m, a sharp boundary is observed
between the shallow marine shale of the Bargo Claystone and the channel
deposits of the Darkes Forest Sandstone, indicated by gamma ray logs (Well
125, Fig. 4-2).
59

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-2: Well 125


4-3-1-1-4 Darkes Forest Sandstone
In this study, the Darkes Forest Sandstone is 5.7 m thick and is characterised
by the presence of siltstone and sandstone (Appendixes 3 and 4).
Two units of channel facies range in thickness from 1.8 to 2.5 m at an average
of 2.2 m and they contain siltstone and sandstone. This facies shows a sharp
base with the underlying Bargo Claystone. The channel facies are represented
by basal sequences of medium-grained sandstone (Appendix 2-2d) that are an
average of 0.9 m thick. These units include sparse silty and muddy wisps and
are thinly bedded to medium bedded. The overlying fine-grained sandstone
(Appendix 2-3a, b) usually has a diffuse base with the underlying unit. It is very
thinly bedded in sequences averaging 0.8 m thick, and includes silty lenses and
muddy wisps.
Also, two units of siltstone are recognised with an average thickness of 0.5 m
(Appendix 2-3c). Both units show diffuse bases with the underlying unit and are
thickly laminated. Sandy bands occur in the siltstone units.
In well logs, the gamma ray log showed a blocky shape trend with a sharp base
and top indicating channel deposits. Sandstone is the dominant facies and
occurs between 403.5 and 405.6 m and between 406.7 and 407.8 m with a
blocky shaped gamma ray response (Well 125, Fig. 4-3). Siltstone is present
between 405.7 and 406 with irregular and higher gamma ray responses (Well
125, Fig. 4-3). At the base of unit, well logs showed the upper contact as sharp
with a decrease to lower gamma ray intensity (Well 125, Fig. 4-3).
60

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
Resistivity
VL2F VL6F FE1 FE2

Fig. 4-3: Well 125

4-3-1-1-5 Allans Creek Formation


The Allans Creek Formation has an average thickness of 5.5 m and includes
shale, carbonaceous mudstone, siltstone and sandstone (Appendixes 3 and 4).
This formation has a diffuse base with the underlying Darkes Forest Sandstone.
Five units of channel facies range in thickness between 0.5 and 2.7 m at an
average of 1.3 m. They contain shale, carbonaceous mudstone, siltstone and
sandstone. One unit of fine-grained sandstone is present in the formation (1.3 m
thick) and it includes silty bands and wisps (Appendix 2-3d). This unit is
recorded as thickly laminated.
Brownish grey to grey of siltstone units (Appendix 2-4a) are widespread in the
formation with thicknesses ranging between 0.1 and 1.2 m (average of 0.6 m).
These units show diffuse bases with underlying units and are mostly thickly
laminated. They are carbonaceous and muddy, and contain clayey matrix and
sandy wisps throughout the units. They are interbedded with sandstone and
may have coal bands over the top.
Beds of shale are present as individual layers and range from 0.03 to 0.07 m
thick (average of 0.05 m; Appendix 2-4b) and have diffuse bases with the
underlying units. They are brownish black, grey, brownish grey to brownish buff
and are very fine- to fine-grained. The shale beds are clayey and tuffaceous
and they are interbedded with siltstone and mudstone. Bedding is thickly
laminated in most described units.
61

The formation is characterised by carbonaceous mudstone (Appendix 2-4c),


which is black to brownish black and very fine- to fine-grained. These units are
thickly laminated and have thicknesses of up to 0.3 m (average of 0.2 m). The
carbonaceous mudstone is clayey and includes claystone and tuffaceous
bands.
Gamma ray and sonic logs showed that this unit consists of channel deposits.
Fine-grained deposits, such as fine sandstone and siltstone, occur with irregular
and higher gamma ray responses (Well 125, Fig. 4-4). Gamma ray and sonic
logs showed a number of sandstone intervals. It is recognised by a relative
increase in gamma ray intensity and a relatively decrease in sonic and
resistivity logs (Well 125, Fig. 4-4). Siltstone is recorded in well logs with
irregular and higher gamma ray responses (Well 125, Fig. 4-4). The shale unit
occurs with a high gamma ray signature (Well 125, Fig. 4-4). At the top of this
formation, the contact is distinct with an increase in gamma ray log intensity
(Well 125, Fig. 4-4).
Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-4: Well 125 (red lines indicate the top and the base of
the Allans Creek Formation.
4-3-1-1-6 American Creek Coal Member
Units of shale, carbonaceous mudstone, coal and nepheline syenite are
included in the American Creek Coal Member that is 5.9 m thick (Appendixes 3
and 4). Shale units appear in this formation, varying between 0.2 and 0.4 m
thick (Appendix 2-4d). Carbonaceous mudstone (Appendix 2-5a) is included in
the lower and upper parts of formation (0.3 m thick). Also, the American Creek
Coal Member is characterised by an abundance of coal (0.2-0.4 m thick).
62

4-3-1-1-7 Kembla Sandstone


Our study showed that the Kembla Sandstone has an average thickness of 6.2
m, including shale, siltstone and sandstone (Appendixes 3 and 4).
The channel facies ranges from 0.3 to 1.7 m in thickness at an average of 1.1 m
and contains shale, siltstone and sandstone. Medium-grained sandstone is rare
and occurs at the base the Kembla Sandstone (0.03 m thick). It contains
pebbles lenses and is massive bedded. Rare beds show fining upwards. Finegrained sandstone is grey, 0.1-1.5 m thick (average of 0.9 m) and includes
carbonaceous laminae, bands and is muddy (Appendix 2-5b, c). This facies is
characteristically thinly to medium bedded in places and grades upwards to
mudstone.
A bed of grey siltstone varies between 0.2 and 1.5 in thickness at an average of
0.9 m.
Shale facies (Appendix 2-5d) occurs at the top of the Kembla Sandstone,
having average thickness of 0.2 m and is grey to greyish brown. The shale has
a gradational lower contact with the underlying unit. Bedding in this facies is
thinly to thickly laminated.
Sandstone, siltstone and shale units are also recognised by well logs in the
Kembla Sandstone (Well 42, Fig. 4-5). Sandstone beds occur with high gamma
ray and sonic logs values with stable resistivity values (Well 42, Fig. 4-5). These
beds have blocky shaped gamma ray responses. Between 681.2 and 681.3 m
and between 678.2 and 678.5 m and also at the top of formation, shale beds
are observed by irregular gamma ray and sonic logs (Well 42, Fig. 4-5). Gamma
ray values and sonic log values in shale units are less than those in sandstone
units (Well 42, Fig. 4-5). At the base of the formation, gamma ray and sonic logs
show fining upward channel deposits with blocky shaped gamma ray and sonic
log responses. At the top of formation, the shale has a gradational lower contact
and a decrease in gamma ray responses. These features are similar to the
sedimentology description and indicate channel deposits.

63

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-5: Well 42

4-3-1-1-8 Wongawilli Coal


In this study area, the Wongawilli Coal has a thickness of 10.7 m and contains
coal, tuff, carbonaceous shale and shale (Appendixes 3 and 4). Shale facies
ranges from 0.01 to 0.3 m in thickness at an average of 0.1 m. Carbonaceous
shale is less common and is distributed in beds varying between 0.01 and 0.1 m
thick (Appendix 2-6a; average of 0.04 m). Coals beds observed in the
Wongawilli Coal average 0.1 m in thickness (Appendix 2-6b). Tuff beds are
more important in the Wongawilli Coal (0.2 m thick) than in other formations and
usually are brownish buff (Appendix 2-6c, d).
4-3-1-1-9 Unnamed Member Three
The Unnamed Member Three is 5.5 m thick, and includes shale and sandstone
(Appendixes 3 and 4). The shale of the Unnamed Member Three shows a sharp
planar base with the underlying Wongawilli Coal.
This unit consists mainly of floodplain facies in beds varying between 0.3 and
1.1 m in thickness at an average of 0.6 m and contains shale and sandstone.
Fine-grained sandstone (0.2 m thick) is present and is very thinly bedded. It is
grey and is silty in places.
Shale units (0.1-1.1 m thick; average of 0.4 m) are the dominant lithology in the
formation and vary in colour between grey, greyish brown to brownish grey
(Appendix 2-7a). The shale is thinly to thickly laminated and includes
carbonaceous, silty laminae in places, sandy and silty lenses and a silty muddy
matrix.
64

In well logs, gamma ray and sonic logs showed sandstone and shale units in
the Unnamed Member Three (Well 125, Fig. 4-6). Sandstone and shale beds
are observed with high gamma ray and sonic logs values (Well 125, Fig. 4-6).
Sandstone occurs with blocky shaped gamma ray responses between 375.2
and 375.3 m whereas shale occurs with irregular higher gamma ray logs (Well
125, Fig. 4-6). At the base of the formation, the shale has a sharp contact with
an increase to higher gamma ray responses (Well 125, Fig. 4-6).
Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-6: Well 125 (red lines indicate the top and the base of the
Unnamed Member Three.

4-3-1-1-10 Unnamed Member Two


In this study area the Unnamed Member Two is 10.7 m thick. This unit
comprises tuff, carbonaceous mudstone, shale, siltstone and sandstone
(Appendixes 3 and 4). At the base of the Unnamed Member Two, shale has a
distinct base with the underlying unit.
Five units of floodplain facies range in thickness from 2.1 to 8.4 m at an average
of 5.1 m and contain shale, carbonaceous mudstone, tuff and sandstone. Shale
is recorded between 0.02 and 3.4 m thick at an average of 0.8 m (Appendix 27b) and usually has a distinct base with underlying units. Bedding in the shale
units is usually thinly laminated. The shale contains coal wisps, calcite, silty and
tuffaceous lenses, sideritic bands, tuffaceous and sandstone bands.
Fine-grained sandstone (Appendix 2-7c, d) is only present in middle and upper
parts of the Unnamed Two Member (0.4-1.9 m thick, average of 1.1 m) and are
grey. These units are well sorted and thinly to medium bedded. They have
65

lenses of mudstone throughout the sandstone and have distinct bases with the
underlying mudstone units.
Carbonaceous mudstone facies is limited to the middle part of the Unnamed
Member Two and ranges from 0.1 to 1.4 m in thickness (average of 0.8 m;
Appendix 2-8a). It is thinly laminated. Most carbonaceous mudstone units have
a distinct base with the underlying units.
Tuff is observed in the lower and middle parts of the Unnamed Member Two in
beds between 0.03 and 0.3 m thick with an average of 0.09 m (Appendix 2-8b).
These units are characterised by a distinct base with underlying units whereas
bedding in all tuff units is thinly laminated.
In this formation, well logs confirmed that shale units are more common with
irregular and higher gamma ray and sonic log responses (Well 42, Fig. 4-7).
Shale units typically show a distinct base and an upward increase to higher
gamma ray responses (Well 42, Fig. 4-7). Sandstone units were shown by well
logs at some intervals with blocky shaped gamma ray responses (Well 42, Fig.
4-7). Sandstone is recorded in well logs with stable gamma ray and sonic logs.
Between 646.2 and 646.8 m, the sandstone is characterised by a distinct base
and an upward increase in the gamma ray logs (Well 42, Fig. 4-7).
Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-7: Well 42.


4-3-1-1-11 Cape Horn Coal Member
This unit is totally composed of black coal that varies between banded bright
and banded dull (Appendix 2-8c). Thickness of this unit ranges from 0.1 to 0.2
m at an average of 0.1 m.
66

4-3-1-1-12 Laurence Sandstone


In this study area, the Lawrence Sandstone has a thickness of 6.8 m and
includes siltstone and sandstone (Appendixes 3 and 4).
Two channel facies occur between 0.8 and 6.5 m thick at an average of 3.4 m
and contain fine- to medium-grained sandstone. The two units of mediumgrained sandstone (Appendix 2-8d; 2.4-4.1 m thick, average 3.3 m) include a
few carbonaceous laminae. The medium-grained sandstone fines upwards to
fine-grained sandstone. Bedding in these units ranges from thinly bedded to
massive.
At the top of the Lawrence Sandstone, fine-grained sandstone (Appendix 2-9a)
is described as grey and thickly laminated (0.8 m thick). It contains
carbonaceous laminae and has a gradational lower contact with the underlying
unit.
In this formation, well logs showed that sandstone is the dominant facies with
blocky shaped gamma ray and sonic log responses (Well 42, Fig. 4-8). Between
636.6 and 639 m, the well logs showed a fining upward trend at least 3 m thick
associated with irregular gamma ray responses (Well 42, Fig. 4-8). Also, the
well logs showed that the upper part of the formation has a gradational lower
contact and an upward decrease in gamma ray intensity (Well 42, Fig. 4-8,
635.9-636.6 m). Heterolithic facies was observed by serrate curves shown by
the gamma ray logs (Well 42, Fig. 4-8). These heterolithic facies indicate fluvial
deposits.

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-8: Well 42.


4-3-1-1-13 Balgownie Coal Member
This study showed that the Balgownie Coal Member occurs as a 1.9 m thick
unit

containing

carbonaceous

shale,
67

coal

and

an

igneous

intrusion.

Carbonaceous shale and coal occurs in beds with thicknesses of 0.6 m and 0.3
m, respectively.
4-3-1-1-14 Loddon Sandstone
The Loddon Sandstone is 7.5 m thick and has mudstone, carbonaceous
mudstone, sandstone and coal (Appendixes 3 and 4). At the base of the Loddon
Sandstone, carbonaceous mudstone shows a distinct base with the underlying
Balgownie Coal Member.
Three channel facies units range from 0.3 to 3.1 m in thickness, at an average
of 1.5 m, and include carbonaceous mudstone, shale and sandstone. Mediumgrained sandstone (Appendix 2-9b, c) is present at an average of 1.1 m thick in
the Loddon Sandstone and is thinly bedded. It has an erosional lower contact
with the underlying mudstone unit and contains granules, cobbles and rare
carbonaceous lenses.
Fine-grained sandstone (0.9 m thick; Appendix 2-9d) gradationally overlies the
medium-grained sandstone and is very thinly bedded. Carbonaceous laminae
bands are widespread in this unit. Shale units have a thickness varying between
0.2 and 0.8 m (average of 0.4 m) in the Loddon Sandstone and also include
sandy lenses throughout the unit (Appendix 2-10a). The shale beds are thinly
laminated.
In this formation, carbonaceous mudstone beds are 0.02 and 0.1 m thick
(average of 0.1 m) and are brownish black, dark grey to black, and are thinly
laminated (Appendix 2-10b). At the top of the Loddon Sandstone, carbonaceous
mudstone contains coaly lenses.
Gamma ray and sonic log responses confirmed that the Loddon Sandstone
consists of channel deposits (Well 42, Fig. 4-9). Well logs showed sandstone
with blocky shaped gamma ray responses between 628.7 and 631.8 m (Well
42, Fig. 4-9). Sandstone occurs with a decrease in gamma ray and an increase
in sonic log responses (Well 42, Fig. 4-9). Between 628.7 and 631.8 m,
medium-grained sandstone occurs with a blocky shaped gamma ray response
and is overlain by fine-grained sandstone with irregular and higher gamma ray
68

responses (Well 42, Fig. 4-9). Also, shale is observed with higher gamma ray
responses (Well 42, Fig. 4-9). Gamma ray values in shale units are higher than
those in the sandstone (Well 42, Fig. 4-9). At 632.1 m, well logs showed that the
Loddon Sandstone has a distinct base and an upward increase to higher
gamma ray and decrease to lower sonic log values (Well 42, Fig. 4-9). In this
formation, heterolithic facies were also observed by serrate curves shown by
the gamma ray logs, indicating fluvial deposits (Well 42, Fig. 4-9).
Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-9: Well 42.

4-3-1-1-15 Bulli Coal


In this study area the Bulli Coal has a thickness of 2.7 m and includes mostly
coal, which ranges in beds from 0.2 to 0.6 m thick (Appendix 2-10c; average of
0.4 m).

4-3-2 Narrabeen Group


The study of the formation thicknesses and facies descriptions were obtained
from well logs of the drill cores. The average thicknesses of the formations were
determined from all drill cores, including EAW 18 a, EAW 30, EAW 42, EAW
156, EDEN 115, EDEN 124, EDEN 125, EDEN 126 and EDEN 127. The facies
was described from EAW 42.

69

4-3-2-1 Clifton Sub-Group


4-3-2-1-1 Coalcliff Sandstone
The Coalcliff Sandstone is present at the base of the Narrabeen Group and has
a thickness of 14.9 m, including mudstone, siltstone and sandstone, with the
latter being the most abundant lithofacies (Appendixes 3 and 4). This facies is
characterised by a sharp erosional basal contact with the underlying Bulli Coal.
Eight channel facies units range from 1 to 6.7 m in thickness at an average of
3.4 m thick. They include shale, siltstone and sandstone. Five units of very
coarse-grained sandstone are distributed within the lower and upper parts of
formation and are present in beds up to 1.2 m thick (average 0.4 m; Appendix
2-10d). In the lower part, coarse-grained sandstone overlies medium-grained
sandstone whereas in the upper part, very coarse-grained sandstone has a
distinct base with the underlying mudstone. Bedding ranges from thin to
massive in this facies. Mudstone clasts and rare pebbles are observed in this
facies in the lower part of the formation.
Sets of medium-grained sandstone (Appendix 2-11a, c) usually overlie the
coarse-grained sandstone and bed thickness does exceed 4 m with an average
thickness of 1.4 m. Rare beds show fining upwards. Mudstone clasts occur
frequently in this facies in the lower and middle parts of the formation. Also, the
beds varied between very thinly bedded and massive.
Fine-grained sandstone beds are between 0.2 and 2.8 m thick (average of 0.9
m; Appendix 2-11b, d). This facies ranges from thinly bedded to massive. Silty
carbonaceous laminae and rare mudstone clasts are distributed through this
facies in the lower part of the formation. Carbonaceous wisps and rare granules
are present in the middle and upper parts of the formation. At the top of the
Coalcliff Sandstone, this facies includes intraclasts of mudstone.
The average thickness of the siltstone units is 0.06 m. These units are rare
components in the lower part of formation. They usually have a diffuse base
and are thinly laminated and muddy.

70

In the middle and upper parts of formation, mudstone occurs in thin layers
where bed thickness varies between 0.1 and 0.7 m with an average of 0.4 m
(Appendix 2-12a). This facies is grey and contains sandy lenses. The mudstone
ranges from thinly laminated to thinly bedded. The lower contact usually shows
a gradational or distinct base with the underlying unit.
Also, well logs showed common sandstone, siltstone and shale indicating that
this formation consists of channel deposits (Well 42, Fig. 4-10b). An incisedvalley was shown in well logs as sharp based blocky sandstone (Well 30, Fig. 410a). Sandstone is common with blocky shaped gamma ray responses (Well
42, Fig. 4-10a, b). Between 602.9 and 613.3 m, coarse and medium-grained
sandstone occurs with blocky shaped gamma ray responses and is typically
overlain by fine-grained deposits also with blocky shaped gamma ray responses
(Well 42, Fig. 4-10b). Well logs showed that shale is rare in channel deposits
(Well 42, Fig. 4-10b). An upward fining trend was observed with irregular
gamma ray logs between 598.9 and 599.3 m (Well 42, Fig. 4-10b). Also,
channel deposits were confirmed in the Coalcliff Sandstone by serrate curves
which indicate heterolithic facies (Well 42, Fig. 4-10a, b). At the base of the
formation, well logs showed a sharp base followed by an upward increase in
gamma ray and decrease in sonic log intensity (Well 42, Fig. 4-10b).
a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

b)

Fig. 4-10: a) Well 30 and b) Well 42.


71

Resistivity
FE1 FE2

4-3-2-1-2 Wombarra Claystone


In this study area, the Wombarra Claystone is characterised by shale and
sandstone and is 30.4 m thick (Appendixes 3 and 4).
Four sequences of floodplain facies range from 4.3 to 21 m in thickness at an
average of 9 m. They mostly consist of shale and fine-grained sandstone.
Grey to greenish grey shale is the dominant lithology in the formation and
occurs in sequences up to 5 m thick with an average of 3.3 m (Appendix 2-12b).
Organic matter content is the reason for the grey colour. Shale commonly has a
gradational base with the underlying fine-grained sandstone in lower part of the
formation. Also, bedding is thinly laminated in the lower part of formation
whereas it is medium bedded, thinly bedded and thinly laminated in the upper
part of formation. Sandy bands and interbeds are recorded in some shale units
in the Wombarra Claystone.
The lower and upper parts of the Wombarra Claystone contain grey fine-grained
sandstone beds (Appendix 2-12c), which have a maximum thickness of 2.7 m
(average 1.6 m). In the upper part of the formation, rare beds show coarsening
upwards. Carbonaceous laminae and carbonaceous wisps occur in some
sandstone beds, although they are rare. Bedding ranges from thinly bedded to
massive.
Also, well logs confirmed that the Wombarra Claystone is rich in shale units with
a minor presence of sandstone (Well 42, Fig. 4-11). Shale and sandstone units
are observed in well logs with blocky shape gamma ray responses (Well 42,
Fig. 4-11). Between 564.2 and 569.1 m, sandstone beds are common with
lower gamma ray responses (Well 42, Fig. 4-11). Gamma ray logs showed a
coarsening upward trend between 566.4 and 567.1 m and also showed a fining
upward trend between 554.3 and 557.3 m (Well 42, Fig. 4-11). Gamma ray logs
showed that channel deposits and floodplain deposits are interbedded at depths
between 563.5 and 564.2 m (Well 42, Fig. 4-11). Between 559.2 and 560.9 m
and between 563.5 and 564.2 m, well logs are clearly very serrated in the
Wombarra Claystone and this may as result of the common occurrence of
interbedded sandstone and shale beds (Well 42, Fig. 4-11).
72

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-11: Well 42.


4-3-2-1-3 Scarborough Sandstone
In this study area, the Scarborough Sandstone has a thickness of 36.4 m and
overlies the Wombarra Claystone. It includes shale, siltstone and sandstone
(Appendixes 3 and 4). At the base of the Scarborough Sandstone, fine-grained
sandstone shows a distinct base with the underlying Wombarra Claystone.
The eight recognised channel facies units vary between 2 and 11.7 m thick at
an average of 5 m. Coarse-grained sandstone is recognised in the lower and
upper parts of formation (Appendix 2-12d). It is grey, up to 0.6 m thick and is
poorly sorted. Rare beds show fining upwards. Coarse-grained sandstone has
common granules and pebbles in the lower part of formation whereas it
contains mudstone clasts and muddy pebbles in the upper part of formation.
The units are medium bedded.
In the Scarborough Sandstone, medium-grained sandstone is the dominant
lithology with thicknesses not less than 3.6 m (Appendix 2-13a, b; average 1.4
m). Channel action is indicated by the thicker beds. The abundance of this
facies indicates little variation in grain size and bedding. The sandstone is
moderately to well sorted with rare beds show fining upwards. Medium-grained
73

sandstone contains coarse granule bands and scattered granules and pebbles
in the lower part of formation, whereas it contains mudstone clasts, mudstone
pebbles and mudstone band in the middle and upper parts of the formation.
Bedding varies between thinly bedded and massive in this facies. Most
medium-grained sandstone beds have a distinct base with the underlying unit.
The channel facies includes fine-grained grey sandstone (Appendix 2-13c, d)
that ranges from 0.1 to 2.6 m thick (average of 1 m). Fine-grained sandstone
rarely contains granules and mudstone clasts and ranges from very thinly
bedded to massive.
Siltstone is rare in the Scarborough Sandstone and is observed as thin layers
up to 0.3 thick (Appendix 2-14a). It is grey and characterised by thick laminae.
The siltstone has a distinct base with the underlying medium-grained
sandstone.
Beds of grey shale range from 0.07 to 1.5 m in thickness with an average of 0.5
m (Appendix 2-14b). It is most common in the middle part of the formation.
Sandy and silty lenses occur in places in the shale units. The majority of shale
units are thinly laminated and but some units are thickly laminated.
Sandstone, siltstone and shale are determined by well logs (Well 42, Fig. 412a). Gamma ray values in sandstone, siltstone and shale are relatively similar
(Well 42, Fig. 4-12a). Sandstone beds are recorded with blocky shaped gamma
ray responses (Well 42, Fig. 4-12a). Between 545.9 and 548.4 m, coarse and
medium-grained sandstone is present with blocky shape gamma ray responses
and is overlain by fine-grained deposits with irregular and higher gamma ray
responses (Well 42, Fig. 4-12a). Channel incisions occur at different depths
indicated by gamma ray, sonic logs and resistivity logs (Well 42, Fig. 4-12a). In
general, channel deposits are observed in the Scarborough Sandstone by
serrate curves which indicate heterolithic facies (Well 42, Fig. 4-12a). An
upward fining succession was shown by gamma ray curves between 520.7 and
521.1 m and between 523.7 and 525.7 m and between 537.3 and 538.3 m (Well
42, Fig. 4-12a). Gamma ray logs showed that channel deposits and floodplain
deposits are interbedded at depths between 531.5 and 534 m and between
74

536.8 and 537.3 m (Well 42, Fig. 4-12a). Within these depths, well logs are
clearly very serrated in this formation as result of the occurrence of interbedded
sandstone and shale beds (Well 42, Fig. 4-12a). At the base, well logs
displayed that the Scarborough Sandstone is characterised by a distinct base
and an upward relative increase in gamma ray logs (Well 42, Fig. 4-12a). The
presence of a blocky shaped trend and a sharp base by gamma ray logs, in
addition to the presence of conglomerate (Well 125, 309.3-320.4 m, Fig. 4-12b),
indicate that channel base sediments were erosive with high energy flows being
recognised in the Scarborough Sandstone.
a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

b)

Fig. 4-12: a) Well 42 and b) Well 125 (red arrows


indicate sharp base.

75

Resistivity
FE1 FE2

4-3-2-1-4 Stanwell Park Claystone


In this study area, shale and sandstone are observed in the Stanwell Park
Claystone, which has a thickness of 14.2 m (Appendixes 3 and 4).
This formation mainly consists of floodplain facies that have an average bed
thickness of 0.7 m and include common shale and fine-grained sandstone. At
the base of the Stanwell Park Claystone, shale shows a distinct basal contact
with the underlying Scarborough Sandstone.
Shale is observed as individual layers at the top and base of the Stanwell Park
Claystone (Appendix 2-14c). It occurs as grey and greyish brown beds up to 1.3
m thick (average 0.9 m). The shale is thinly laminated at the top and base of the
Stanwell Park Claystone and it usually has a gradational lower contact with the
underlying unit.
Grey to greyish cream fine-grained sandstone (Appendix 2-14d) is present in
beds varying in thickness between 0.3 and 0.5 m (average of 0.4 m). The
sandstone is very thinly bedded or thickly laminated in this facies. Fine-grained
sandstone contains granule bands and typically has a distinct base with the
underlying unit.
Well logs confirmed that the Stanwell Park Claystone consists of floodplain
deposits (Well 42, Fig. 4-13). They showed common shale and sandstone units
on the gamma ray and sonic logs (Well 42, Fig. 4-13). Shale units occur with
irregular and higher gamma ray responses whereas sandstone is present with
blocky shaped gamma ray responses (Well 42, Fig. 4-13). Thus, gamma ray
values in shale are higher than those in sandstone units (Well 42, Fig. 4-13). At
504.3 m, the gamma ray log showed a distinct contact between the floodplain
deposits in the Stanwell Park Claystone and the overlying channel deposits in
the Bulgo Sandstone (Well 42, Fig. 4-13). Between 506.4 and 506.7 m, the
shale units have a gradational lower contact shown by an increase in the
gamma ray logs (Well 42, Fig. 4-13).

76

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-13: Well 42.


4-3-2-1-5 Bulgo Sandstone
In this study area, the Bulgo Sandstone has a thickness of 135.8 m and
includes shale, siltstone, sandstone and igneous intrusions (Appendixes 3 and
4).
Forty five units of channel facies are present and they vary in thickness
between 0.4 and 8.7 m with an average of 4.7 m. The channels contain shale,
siltstone and sandstone. The basal unit of medium-grained sandstone has a
distinct base with the underlying Stanwell Park Claystone.
Coarse-grained sandstone (Appendix 2-15a, d) is present in the lower and
middle parts of the Bulgo Sandstone with bed thickness varying between 0.02
and 3.4 m (average 1.1 m). It is grey, greyish cream, whitish grey, cream and
brownish cream and is poorly to moderately sorted in most units. Rare beds
show coarsening upward. Coarse-grained sandstone is thickly to medium
bedded and has a distinct base with the underlying units. In general, these units
include granules, pebbles, cobbles, fine sandy bands, mudclasts, muddy bands
and carbonaceous bands.
Medium-grained sandstone has a maximum bed thickness of 5 m with average
of 1.9 m and is the dominant lithology in the Bulgo Sandstone (Appendix 2-15b,
2-16a). It is cream, grey, whitish grey, pinkish grey, greyish white and greyish
cream and is poorly to well sorted. Rare beds show fining upwards or
coarsening upwards. Bedding ranges from thinly bedded to massive in the
medium-grained sandstone. This facies contains pebbles, granules, granule
and coarse-grained bands, fine-grained and muddy bands, mudstone clasts and
mudstone interbeds in the lower part of formation. Also, cobbles, pebbles,
77

pebbly bands, mudstone pebbles, silty interbeds, and carbonaceous and muddy
bands are observed in medium-grained sandstone in the middle and upper
parts of formation. The medium-grained sandstone usually has a distinct base
with the underlying unit.
Fine-grained sandstone occurs throughout the formation (Appendix 2-15c, 216b) in beds which vary in thickness between 0.1 and 4 m thick with an average
of 0.9 m. Multi colours of grey, greyish cream, whitish grey occur in fine-grained
sandstone. Rare beds show fining upward. Bedding varies between thinly
bedded and massive but is generally thinly bedded in the fine-grained
sandstone. Pebbles cobbles, mudclasts, silty and carbonaceous bands,
carbonaceous laminae, and mudstone interbeds are distributed in the middle
and upper parts of fine-grained sandstone. Also, fine-grained sandstone
generally has a distinct base with the underlying unit.
Grey siltstone is not common in the Bulgo Sandstone. The thickness of this
facies reaches up to 1.2 m. Bedding is described as thickly laminated whereas
disturbed bedding is present as sedimentary structures in this facies.
Throughout the formation, shale units (Appendix 2-16c) occur much more than
siltstone and are grey to greyish maroon having a bed thickness of 0.1-1.8 m at
an average of 0.5 m. These units have a distinct base with the underlying units
and are mostly thinly laminated.
Gamma ray logs showed sandstone, siltstone and shale units in the Bulgo
Sandstone (Well 42, Fig. 4-14). Sandstone is more common in this formation
with blocky shaped gamma ray responses (Well 42, Fig. 4-14). In some
intervals, shale is clear with irregular and higher gamma ray responses (Well
42, Fig. 4-14). In some intervals, medium-grained sandstone is present with
blocky shaped gamma ray responses and is overlain by fine-grained deposits
with irregular and higher gamma ray responses (Well 42, Fig. 4-14, between
404.6 and 408 m). Also, sandstone units generally have a distinct base and an
upward increase in gamma ray logs (Well 42, Fig. 4-14, for example at 446.1
m). Well logs showed heterolithic facies with serrate curves in gamma ray (Well
42, Fig. 4-14). This supports that the Bulgo Sandstone consists of channel
78

deposits. Between 492.1 and 493.2 m, 496.7 and 497.7 m, gamma ray logs are
very serrated in some cases and this indicates the presences of interbedded
sandstone and shale beds (Well 42, Fig. 4-14). This means that channel
deposits and floodplain deposits are interbedded in some intervals. Incisedvalley is recorded by gamma ray, sonic logs and resistivity logs (Well 42, Fig. 414). Valley incision in the well logs is characterised by a sharp base and a
blocky sand body. Also, upward fining is shown by gamma ray logs in the
incised channel deposits (Well 42, Fig. 4-14, at 363.1 and 363.6 m). Also, an
upward coarsening succession was also displayed by gamma ray logs in the
Bulgo Sandstone between 480.3 and 480.9 m and between 480.9 and 481.6 m
in Well 42 (Fig. 4-14). A sharp base is recorded between the channel deposits
in the Bulgo Sandstone and the underlying floodplain deposits in the Stanwell
Park Claystone by gamma ray logs at 297.7 m in Well 42 (Fig. 4-14). A
hummocky clinoform pattern is shown in well logs with a coarsening upward
shape pattern in the gamma ray log between 480.3 and 481.6 m (Well 42, Fig.
4-14). Debris flow deposits are clear in well logs by fining upward trend with a
sharp base between 447.1 and 449.2 m in Well 42 (Fig. 4-14). Some individual
beds are characterised by sharp upper and lower boundaries and this is
interpreted to represent high energy gravity flows according to Van Wagoner et
al. (1990).

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Fig. 4-14: Well 42.

79

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Fig. 4-14 (cont): Well 42.

80

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Fig. 4-14 (cont): Well 42.


81

Sonic
VL2F VL6F

Resistivity
FE1 FE2

4-3-2-1-6 Bald Hill Claystone


The Bald Hill Claystone is 21.7 m thick in the study area and includes claystone,
siltstone and sandstone (Appendixes 3 and 4). Claystone occurs at the base of
the Bald Hill Claystone and has a gradational lower contact with the uppermost
fine facies of the underlying Bulgo Sandstone.
Four sequences of floodplain facies range in thickness from 0.7 to 12.9 m at an
average of 7.6. These facies include common shale and siltstone.
Maroon and greyish maroon shale (Appendix 2-16d) is prevalent in beds
ranging from 0.7 to 4 m thick (average 3.4 m). This facies is predominantly
thinly laminated but includes sandy and clayey bands, plus disseminated
sideritic patches.
Grey siltstone is uncommon in the Bald Hill Claystone (Appendix 2-17a). Its bed
thickness does not exceed 0.7 m. It is thinly bedded and shows a distinct base
with the underlying shale.
Well logs showed shale units with irregular gamma ray and sonic log responses
in the Bald Hill Claystone (Well 42, Fig. 4-15). Between 294.9 and 297.7 m,
shale occurs with irregular and higher gamma ray and sonic log responses
(Well 42, Fig. 4-15).
Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Fig. 4-15: Well 42.

82

Sonic
VL2F VL6F

Resistivity
FE1 FE2

4-3-2-2 Gosford Sub-Group


4-3-2-2-1 Garie Formation
In this study area, the Garie Formation is a floodplain facies and has a
thickness of 2.7 m. It consists of greyish cream tuffaceous shale or airfall tuff
(Appendix 2-17b) with a few sandy interbeds and is thinly laminated.
4-3-2-2-2 Newport Formation
In this study area, the Newport Formation has a thickness of 14 m and includes
shale, siltstone, and sandstone (Appendixes 3 and 4).
Six sequences of lagoonal facies occur in the Newport Formation and range in
thickness from 0.7 to 5.5 m at an average of 2.8 m. They include shale, siltstone
and fine-grained sandstone.
Fine-grained sandstone (Appendix 2-17c, d) is distributed in the lower and
middle parts of the formation as individual layers varying between 0.1 and 1.7 m
thick with an average of 0.8 m. Bedding ranges from thinly bedded to thickly
bedded. These beds include silty interbeds, silty laminae and carbonaceous
laminae, with the latter being abundant in the middle part of the formation. Finegrained sandstone includes sedimentary structures such as disturbed bedding
that may be interbedded with fine-grained sandstone in some sequences.
Siltstone facies (Appendix 2-18a) occurs in the Newport Formation with
thickness ranges from 0.3 m to 0.7 m (average 0.5 m). This facies is thinly
bedded and may also be interbedded with disturbed bedding.
The lower and upper parts of the Newport Formation contain grey shale facies
(Appendix 2-18b) in beds ranging from 0.7 to 0.9 m thick (average 0.8 m). This
facies is thinly laminated to thickly laminated and includes disturbed bedding as
sedimentary structures.
Lagoonal facies are observed in well logs indicated by gamma ray responses
which show distinct muddy intervals (Well 42, Fig. 4-16). Between 259.5 and
263.5 m, upward coarsening intervals are observed in well logs and indicate
wash-over fans over the muddy lagoonal facies. Sandstone beds are clear with
83

irregular gamma ray responses whereas shale beds are clear with irregular and
higher gamma ray responses (Well 42, Fig. 4-16). At the top of the formation,
the shale shows a distinct contact with an upward decrease in the gamma ray
logs (Well 42, Fig. 4-16).
Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 4-16: Well 42.

4-3-3 Hawkesbury Sandstone


The formation thickness and facies descriptions were obtained from well logs of
the drill cores. The thickness was determined from all drill cores, including EAW
18a, EAW 30, EAW 42, EAW 156, EDEN 115, EDEN 124, EDEN 125, EDEN
126 and EDEN 127. The facies was described from EAW 42.
In this study, the Hawkesbury Sandstone has a thickness of about 138.2 m and
includes shale, siltstone, sandstone and conglomerate (Appendixes 3 and 4).
Thirty seven recognised channel facies range in thickness from 1.4 to 8.9 m at
average of 5.5 m. It includes shale, siltstone, common sandstone and
conglomerate. The basal unit of this formation has an erosional lower contact
with the underlying Newport Formation.
Three units of conglomerate (Appendix 2-18c) are present in the Hawkesbury
Sandstone (1.1-1.6 m thick; average 1.3 m) and generally have a coarse
granule to pebble grain size. They are grey to brownish grey and poorly sorted.
Rare beds show fining upwards. The conglomerate is medium bedded or
massive and contains cobbles and pebbly bands with a sandy matrix.

84

Throughout the Hawkesbury Sandstone, coarse-grained sandstone (Appendix


2-18d and 2-19a) occurs in beds between 1 and 5.1 m thick (average 2.2 m). It
is grey, cream, greyish grey, brownish cream, whitish grey and greyish white
and is commonly moderately sorted. Rare beds show fining upwards or
coarsening upwards. Coarse-grained sandstone is thinly bedded to massive
and includes carbonaceous laminae, minor pebbles, pebble bands, granule
bands, mudclasts and a muddy matrix. Coarse-grained sandstone has a distinct
base or gradational lower contact with the underlying unit.
Medium-grained sandstone (Appendix 2-19b, c) also occurs throughout the
formation and is mostly moderately to well sorted. It is grey, cream, whitish
grey, greyish cream, brownish grey and whitish cream and the beds range from
0.1 to 5.5 m in thickness with an average of 2.6 m. Bedding is mostly thinly
bedded but ranges to massive. Carbonaceous laminae and carbonaceous
wisps are present in most medium-grained sandstone units. Also, mudstone
lenses, mudclasts, clayey matrix and cobbles are rarely observed. Mediumgrained sandstone commonly has a distinct base with the underlying units.
Grey to cream fine-grained sandstone is recorded in the formation. Its thickness
does not exceed 2.8 m (average 0.6 m; Appendix 2-19d) and it is poorly to well
sorted. Also, rare beds show fining upwards. This facies is thinly bedded to
massive and mostly has a gradational lower contact or distinct base with the
underlying units. Carbonaceous laminae occur in most fine-grained sandstone
units along with occasional pebbles. Clayey matrix, carbonaceous wisps, mud
clasts, granular bands, mudstone and silty bands are also observed in the finegrained sandstone.
Siltstone is mostly grey and has a thickness of between 0.1 and 0.6 m (average
0.4 m; Appendix 2-19e). It occurs in the upper preserved part of the formation
and is thickly laminated.
Grey shale is recognised in the middle and upper parts of the preserved
Hawkesbury Sandstone and has a bed thickness from 0.03 to 0.8 m (average
0.3 m). Thin lamination is dominant in the shale and some units of shale include
sand and pebbles.
85

Gamma ray logs showed conglomerate, sandstone, siltstone and shale units in
the Hawkesbury Sandstone (Well 42, Fig. 4-17). Sandstone beds occur
dominantly with blocky shaped gamma ray responses (Well 42, Fig. 4-17).
Conglomerate beds are observed with low gamma ray logs (Well 42, Fig. 4-17).
Also, the Hawkesbury sandstone units have a sharp base followed by an
increase in gamma ray logs (Well 42, Fig. 4-17, for example at 242 m). Well
logs showed heterolithic facies with serrate curves in the gamma ray log,
indicating channel deposits (Well 42, Fig. 4-17). In some cases, gamma ray
logs are very serrated, showing that interbedded sandstone and shale beds are
present in this formation (Well 42, Fig. 4-17, between 97.1 and 97.8 m, 123.6
and 124.4 m as examples). This means that channel deposits and floodplain
deposits are interbedded in some intervals. Well logs such as gamma ray, sonic
logs and resistivity logs showed incised channels which are characterised by
sharp bases and blocky sand bodies (Well 42, Fig. 4-17). Also, upward fining
and upward coarsening successions were shown by gamma ray logs in the
Hawkesbury Sandstone between 203.4 and 206.4 m and between 236.4 and
245.3 m (Well 42, Fig. 4-17). Between 236.4 and 245.3 m, well logs showed a
hummocky clinoform pattern with a coarsening upward shape pattern in the
gamma ray log (Well 42, Fig. 4-17). Also, between 447.1 and 449.2 m, well logs
showed debris flow deposits with a fining upward trend above a sharp base
(Well 42, Fig. 4-17). Some individual beds are characterised by sharp upper
and lower boundaries and this is interpreted as high energy gravity flows
according to Van Wagoner et al. (1990).

86

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Fig. 4-17: Well 42.

87

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Fig. 4-17 (cont): Well 42.

88

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
Sonic
DEPO RPOR VL2F VL6F

Fig. 4-17 (cont): Well 42.

89

Resistivity
FE1 FE2

4-4 The comparison between previous work and


the present study
The present study, together with Bunny (1972) and Bowman (1974), interprets
the Wilton Formation as floodplain deposits, which is in disagreement with
Jones (1983) who suggested that part of the Wilton Formation was deposited in
a shallow lacustrine environment. The present study and Bamberry et al. (1995)
indicated that shale is the most common lithology in the Wilton Formation.
The present day and previous work all describe the Tongarra Coal as being rich
in coal, whether bright or dull, with the added presence of siltstone,
carbonaceous shale and tuff bands. Also, this study showed that sandstone is
interbedded in the Tongarra Coal. Bamberry et al. (1995) interpreted that
deposition of the sequence from the Woonona Coal Member to the Tongarra
Coal represents delta front, delta margin and lower delta plain deposits and this
study confirms that deposition of the Tongarra Coal occurred in the floodplain
environment possibly on a lower delta plain.
The present study does not differ from previous work by Bamberry et al. (1995)
and Roche (1997) in the description of the Bargo Claystone. All show that
claystone and siltstone are commonly represented in the Bargo Claystone.
Also, the present day confirms the previous work by Bunny (1972) and Bowman
(1974) in the interpretation of the Bargo Claystone as floodplain deposits.
There is no difference between Bamberry et al. (1995), Roche (1997) and the
present study that all describe sandstone and fine-grained deposits in the
Darkes Forest Sandstone. Bamberry et al. (1995) and Roche (1997) showed a
gradational contact between the Darkes Forest Sandstone and the underlying
Bargo Claystone whereas this study showed sharp erosional contact between
the two units.
This study, together with Bamberry et al. (1995) and Roche (1997), described
sandstone, siltstone and shale in the Allans Creek Formation. This study agrees
with Roche (1997) that the Allans Creek Formation was deposited on a
floodplain due to the common occurrence of fine-grained deposits. This study
90

also showed that channel facies are represented by sandstone beds in the
Allans Creek Formation.
The present study, like that of Bamberry et al. (1995), noted that coal is the
dominant lithology in the American Creek Coal Member with a lesser
occurrence of carbonaceous shale. This study also concurs with Roches (1997)
interpretation of a floodplain setting for deposition of the American Creek Coal
Member.
The Kembla Sandstone is characterised by the presence of sandstone facies
associated with minor fine-grained facies, as previously recorded by Bamberry
et al. (1995). Roche and Hutton (1998) and Arditto et al. (2003) stated that the
Kembla Sandstone was deposited in a fluvial environment.
Previous works, represented by Bamberry et al. (1995), Roche (1997) and
Jesus (1999), are not different from the present study in the description of the
Wongawilli Coal. They all record tuff, coal and carbonaceous shale in this unit,
although some previous works indicated the presence of a sandstone facies
which was not detected in this study. Jesus (1999) described the contact
between the Wongawilli Coal and the underlying Kembla Sandstone as
gradational whereas this study has shown that the Wongawilli Coal interfingers
with the top of the underlying Kembla Sandstone.
The present study and previous works indicated that the Lawrence Sandstone
Member is comprised of sandstone and siltstone. This study and the study by
Hutton and Roche (1998) recorded that the Lawrence Sandstone Member was
deposited in a fluvial channel environment.
The Balgownie Coal Member includes coal and carbonaceous shale. Bamberry
et al. (1995) recorded the presence of siltstone in Balgownie Coal Member
whereas this study has noted the presence of igneous intrusions in the
Balgownie Coal Member.
Roche (1997) and this study noted that the Loddon Sandstone is characterized
by channel-fill sandstone deposits together with some overlying fine-grained
sediments. The present study differs from previous works since it recognises a
91

small amount of coal in the Loddon Sandstone. The difference in total thickness
between the present study and previous work is minor. However, this study
showed that medium-grained sandstone occurring as channel-fill deposits is
only 1.4 m thick whereas the thickness of the Sh and Sx lithofacies was 4 m
farther east (Roche, 1997). This study and Hutton and Roche (1998) recorded
that the Loddon Sandstone was deposited in a fluvial environment.
The Bulli Coal is mostly dominated by coal according to the present study and
previous work such as Jones (1983) and Bamberry et al. (1995). In the study
area the coal is 2.7 m thick whereas Bamberry et al. (1995) showed that the
thickness of the coal is more than 3 m in the Campbelltown to Camden area.
The present study and previous works agree that the Coalcliff Sandstone is rich
in sandstone facies including coarse-, medium- and fine-grained sandstone. In
this study area the thickness of the sandstone is up to 14.9 m and Deen (1999)
determined that individual St and Sr lithofacies in the Coalcliff Sandstone do not
exceed 50 cm and 40 cm thick, respectively. Siltstone and shale facies are
observed in the Coalcliff Sandstone based on both the present study and
previous works but with differences in thickness. Dehghani and Jones (1994)
and Deen (1999) showed that lower contact of the Coalcliff Sandstone with the
underlying Illawarra Coal Measures is an erosional contact and this study
confirmed that it has a sharp base with the underlying Illawarra Coal Measures.
The present study and previous works (such as Ward, 1972; Jones, 1986;
Hamilton et al. 1987; Reynolds, 1988; Dehghani and Jones, 1994b; Deen 1999)
came to a similar conclusion that the Coalcliff Sandstone was deposited in a
fluvial environment.
This study and previous works showed that sandstone, siltstone and shale are
all present in the Wombarra Claystone. Shale is dominant in this unit. The
thickness of sandstone sequences in the present study (up to 2.7 m) is different
from the thickness of individual sandstone beds in the study by Dehghani
(1994) who recorded up to 40 cm of individual Sp, Sr, Sh and Sm lithofacies. In
the Wombarra Claystone, previous work by Dehghani (1994) confirmed the
presence of ironstone bands and palaeosols, which were not recognised in the
92

present study. Finally, this study together with Dehghani (1994) and Deen
(1999) showed that the Wombarra Claystone represents a floodplain deposit.
The present study is concurs with previous works by Ward (1972), Bunny
(1972), Bowman (1974), Hamilton et al. (1987) and Dehghani and Jones (1994)
that the Scarborough Sandstone was deposited in a fluvial environment.
Sandstone is the main facies in the Scarborough Sandstone with only rare
occurrences of fine-grained facies assemblages. This study showed that the
maximum thickness of sandstone sequences in this unit is 2.6 m whereas
Dehghani and Jones (1994c) showed that the maximum thickness of individual
Sp, St, and Sh lithofacies are 1.3 m, 50 cm and 50 cm. The lower contact of the
Scarborough Sandstone with the underlying Wombarra claystone is an
erosional contact according to Dehghani and Jones (1994c) and Deen (1999)
and this study confirms that the Scarborough Sandstone is characterized by a
distinct base with the underlying Wombarra Claystone.
In the Stanwell Park Claystone, this study showed that sandstone facies vary
between 0.01 and 3.9 m in thickness whereas Deen (1999) indicated that
sandstone facies varies between 10 cm and 1.8 m thick. Fine-grained facies
represented by siltstone and shale are abundant in the Stanwell Park
Claystone. The thickness of fine-grained facies in the present study is
essentially similar to those in previous works. Shale reaches up to 1.3 m thick in
the present study whereas massive mudstone does not exceed 3 m thick in the
study by Deen (1999). Ward (1972) and Deen (1999) noted that the Stanwell
Park Claystone contains conglomerate, which is absent in the present study
area. Also, Dehghani (1994) indicated that the Stanwell Park Claystone
contains palaeosols which are also absent in the present study. Dehghani
(1994) and Deen (1999) showed that the Stanwell Park Claystone is
characterized by a gradational contact with the underlying Scarborough
Sandstone whereas the present day indicated that the Stanwell Park Claystone
has a distinct base with the underlying Scarborough Sandstone. Studies by
Dehghani (1994), Deen (1999) and this work agree that the Stanwell Park
Claystone represents a floodplain deposit.

93

The present study and previous works by Dehghani (1994) and Deen (1999)
support that sandstone is the dominant facies in the Bulgo Sandstone, although
the thickness of sandstone in the present study is different from those in
previous works. In this study, sandstone sequences are up to 20.5 m thick
whereas in the study by Deen (1999) individual Sm, Sp, St and Sh lithofacies up
to 1 m, 1.1 m, 75 cm and 1 m thick. All these studies show that siltstone and
shale are observably less common. The present study determined that siltstone
and shale facies have thicknesses of up to 1.6 m whereas Deen (1999)
determined than Fl lithofacies has a thickness up to 50 cm. The present study
differs from previous works since it contains no conglomerate. The present
study agrees with the study by Ward (1980), Dehghani (1994) and Deen (1999)
that the Bulgo Sandstone was deposited in a fluvial environment.
In the Bald Hill Claystone, the present study and previous works (Hamilton et
al., 1987; Dehghani, 1994; Deen, 1999) recognized that claystone is the most
widespread lithology. All studies also recognized the presence of sandstone in
the Bald Hill Claystone. There was a minor difference between the thickness of
sandstone determined by the present study (2.7 m thick) and previous work by
Dehghani (1994) where Sm and Sh lithofacies are up to 3 m thick. The present
study did not indicate the occurrence of conglomerate in the Bald Hill Claystone
whereas Dehghani (1994) and Deen (1999) noted that conglomerate occurs in
this unit. The present study and previous work by Dehghani (1994) and Deen
(1999) showed that there is a gradational contact between the Bald Hill
Claystone and the underlying Bulgo Sandstone. The present study and previous
works confirmed that the Bald Hill Claystone was deposited on a floodplain.
In the study area claystone dominates the Garie Formation whereas previous
works by Bunny and Herbert (1971), Dehghani, (1994) and Deen (1999)
indicated that the unit consists of brecciated claystone associated with massive
tuff with some accretionary lapilli.
The present study is similar to previous works by Herbert (1980b), Dehghani
(1994) and Deen (1999), which indicated that the Newport Formation includes
common shale facies. Dehghani (1994), Deen (1999) and this study showed
that conglomerate is present in this unit. Branagan et al. (1966) and Branagan
94

(1969) and the present study noted that the sediments are thinly bedded and
laminated. In addition the present day also showed that some of the sediments
are thickly laminated. Dehghani (1994) and Deen (1999) showed that the
Newport Formation is characterized by a sharp basal contact with the
underlying Garie Formation whereas contact between the Newport Formation
and the underlying Garie Formation was not determined in the present study.
This study agrees with Dehghani (1994) and Deen (1999) that the Newport
Formation was deposited in a lagoonal environment.
The Hawkesbury Sandstone in the present study is very similar to previous
works by Conaghan and Jones (1980), Jones and Rust (1983), Rust and Jones
(1987) and Miall and Jones (2003). Sandstone facies are widespread and thick
in this unit and are mostly medium- to coarse-grained. Minor fine-grained
sandstone, siltstone and shale are also recorded in the Hawkesbury Sandstone
according to the present study and previous works. In this study recorded shale
beds 1 to 3 m thick whereas Rust and Jones (1987) showed that mudrock is 1
to 2 m thick. All these studies have indicated a fluvial environment of deposition
dominated by channel facies.

95

96

Chapter Five
Petrology of Southern Sydney Basin
5-1 Introduction
Three hundred and thirty two samples from the southern Sydney Basin were
selected for petrographic analysis. These samples include one hundred and
ninety seven samples of sandstone, forty four samples of siltstone, seventy
seven samples of shale, nine samples of tuff, three samples of igneous rocks,
one sample of siderite, one sample of carbonate cement and one sample of
coal. Thin section, X-ray diffraction (XRD) and scanning electron microscope
(SEM) methods were used to analyse these samples.
Three units in the southern Sydney Basin were described, including the
Illawarra Coal Measures, Narrabeen Group and Hawkesbury Sandstone. The
Illawarra Coal Measures includes both coarse-grained and fine-grained deposits
in the southern Sydney Basin. Coarse-grained deposits are recorded in the
Kembla Sandstone, Lawrence Sandstone and Loddon Sandstone whereas finegrained deposits are recognized in the Wilton Formation, Tongarra Coal, Bargo
Claystone, Darkes Forest Sandstone and Allans Creek Formation.
The Narrabeen Group also comprises coarse- and fine-grained deposits. The
coarse-grained

deposits

include

the

Coalcliff

Sandstone,

Scarborough

Sandstone and Bulgo Sandstone. The fine-grained include the Wombarra


Claystone, Stanwell Park Claystone, Bald Hill Claystone and Newport
Formation.
This chapter aims to interpret the petrographic features, including the detrital
composition of sandstone, siltstone and shale in the southern Sydney Basin.

79

5-2 Description of Components


5-2-1 Quartz
Quartz is main mineral in the sandstone and plays a role in the determination of
source area. Detrital quartz is composed of both monocrystalline and
polycrystalline quartz grains (Appendix 6-1a-b and 6-3a-b). This classification
depends on inclusions, extinction and grain shape.
5-2-1-1 Monocrystalline quartz
Most sandstone is dominated by monocrystalline quartz grains, which are seen
as individual sand-size crystals (Appendix 6-1a-b). They include plutonic quartz,
volcanic quartz and vein quartz that may have non-undulatory extinction
(extinction angle 5) to undulatory extinction (extinction angle > 5). The
former are more common than the latter, indicating that study area is sourced
from plutonic and volcanic source rocks (Basu, 1985; Ullah et al., 2006). Nonundulatory monocrystalline quartz is unstrained in most samples whereas
undulatory monocrystalline quartz mainly shows a slightly undulose extinction
with the latter indicating a plutonic igneous origin (Al-Juboury et al., 2009).
Monocrystalline quartz is relatively more common than polycrystalline quartz in
all units except the Newport Formation also indicating a plutonic origin for the
quartz grains. According to grain size, monocrystalline quartz grains are
classified as fine-, medium- or coarse-grained. A felsic volcanic igneous
provenance is the main source of fine-grained monocrystalline quartz
(Dickinson, 1970; Ghazi and Mountney, 2011). The weathering of granite is the
main source of medium to coarse monocrystalline quartz grains (Basu et al.,
1975; Pettijohn et al., 1987; Datta, 2005; Dutta, 2007; Ghazi and Mountney,
2011).
In the current study, monocrystalline quartz varies between 40 and 300 m in
most samples. A cloudy appearance of many monocrystalline quartz grains was
observed. The monocrystalline quartz grains show long to concavo-convex and
sutured contacts (Appendix 6-2c) and they also show overgrowths (Appendix 61a). Incipient quartz overgrowths are represented on some quartz grains in the
studied sequence, especially in the upper part of the succession. The
79

monocrystalline quartz grains range from rare angular grains to mostly wellrounded grains. This evidence indicates recycling and a sedimentary origin (AlJuboury et al., 2009). The highly angular quartz grains in a few samples show a
short transport distance or high relief according to the interpretation provided by
Rieser et al. (2005) and Sonel et al. (2009).
Plutonic quartz (common quartz) is characterized by a straight to slightly
undulose extinction while composite quartz is rare. The boundaries of plutonic
quartz grains are commonly straight in most samples. Some plutonic quartz
grains include mineral inclusions, such as muscovite and heavy minerals,
indicating a granitic component (Appendix 6-2a-c). When fluid inclusions are
present in the plutonic quartz they indicate hydrothermal vein quartz. In most
samples, inclusions are scattered. According to Folk (1968), granite batholiths
and granite-gneisses are recorded as the main source of plutonic quartz.
According to Folk (1968), erosion of volcanic rocks provides the source of
volcanic quartz grains which exist as single glassy clear and equant idiomorphic
grains that have straight extinction. The margins of volcanic quartz grains
contain corrosion embayments whereas inclusions are missing. The volcanic
quartz grains have no included diagenetic characteristics. Vacuoles are not
observed in the volcanic quartz while they occur in the vein quartz. This type of
quartz occurs as rounded grains in most samples and sometimes it contains
shards.
The vein quartz has straight to slightly undulose extinction but the majority of
grains show straight extinction. This type of quartz is prevalent in the pebble
size range with an absence of microlites. Pegmatites and hydrothermal deposits
are the main sources of vein quartz grains according to Folk (1968).
5-2-1-2 Polycrystalline quartz
Polycrystalline quartz grains (composite quartz) occur rarely in all units
(Appendix 6-3a-c) suggesting long distance transportation (Dabbagh and
Rogers, 1983; Ullah et al., 2006). They are less common than monocrystalline
quartz grains and are typically elongate in shape. Two forms of polycrystalline
quartz grains are represented in sequences as follows: (1) polycrystalline quartz
77

grains consisting of two to three crystals; and (2) polycrystalline quartz grains of
more than three crystals that are sourced from metamorphic source rocks (Blatt
et al., 1980; Asiedu et al., 2000; Ullah et al., 2006). Polycrystalline quartz grains
do not contain silica overgrowths in all samples and occur mostly associated
with sutured crystal boundaries. Recrystallised, schistose and stretched
metamorphic quartz are the three types of polycrystalline quartz in the southern
Sydney Basin.
Recrystallised metamorphic quartz grains are the most prevalent and show
straight extinction and include single individuals or composite grains which are
more common. Vacuoles are rare in the recrystallised metamorphic quartz,
occurring as inclusions. Also, mica and heavy minerals occur rarely as
inclusions in this type of quartz. In addition, straight crystal contacts are
observed in the recrystallised metamorphic quartz grains.
Schistose metamorphic quartz grains are elongated, have straight extinction
and also consist predominantly of composite grains. This quartz type may
contain common mica flakes. Other inclusions are absent, while vacuoles are
rare in the schistose metamorphic quartz. In some cases, schistose
metamorphic quartz grains are cloudy due to prolonged abrasion.
Stretched metamorphic quartz is clear and is also called sheared quartz. It
consists of single grains that are moderately to strongly undulose, occurring as
elongate crystals. Also, lensoidal sub-crystals are present and are more
common than single grains. Also, inclusions are observed in the stretched
metamorphic quartz.
In general, the majority of quartz grains are subequant to equant, and
moderately to well sorted. They vary in size between 15 to 700 m. Some
quartz grains are partly to completely coated by carbonate cement and clay
(Appendix 6-3a, 6-4c and 7-2f). The quartz grains show straight or sutured to
partly sutured contacts between grains. Sources of detrital quartz including all
three rock types (igneous, metamorphic and sedimentary) with the latter being
the main source according to the criteria of Boggs (2009).

011

Rounded and pitted recycled sedimentary quartz occurs in some samples with
an absence of quartz overgrowth (Appendix 6-4a). Corrosion rims are
characteristic in this type of quartz and are brown in colour.
In this study, chalcedony was observed as fibrous to feathery texture (Appendix
6-4b). Inclusions are present in chalcedony although they are very tiny when
they occur.

5-2-2 Feldspar
Feldspar is the second most important mineral in sandstones, following quartz.
Feldspar importance lies in determining the provenance of the sandstone.
Feldspar grains include fine-grained K-feldspar and plagioclase (Appendix 6-4c
and 6-5a-c). Feldspar grains are less stable than quartz because of their
cleavage. Some large feldspar grains are mainly altered to clay, thus further
reducing feldspar abundance. Also, they may be partly replaced by carbonate
cement and clays (Appendix 6-5b and 6-6a-b), thus determination of the origin
of these grains may be difficult. Alteration of feldspar may occur along the
cleavage planes in some samples. Fresh feldspar dominates altered feldspars
as a result of uplifted and less weathered source rocks (Dey et al., 2009).
However, alteration of feldspars occurs commonly in the Illawarra Coal
Measures and the Narrabeen Group. Mineral inclusions are totally absent in the
feldspar grains. Most feldspar grains are angular and vary in size between 12
and 550 m. Dissolution of detrital feldspar contributes to the formation of
secondary porosity. In general, erosion and climate affect feldspar grains and
they are typically destroyed during extensive weathering in a humid climate
(Tucker, 1991). Also, weathering and diagenetic alteration affects feldspar
grains with the latter causing corrosion of grains boundaries.
5-2-2-1 K-feldspar
K-feldspars grains consist of orthoclase and microcline (Appendix 6-4c) with the
former being more common than the latter. Both grain types are always optically
negative. The orthoclase is untwined, similar to quartz, and its recognition in
thin section is difficult. Staining can help to distinguish it from quartz. It shows a
lower birefringence compared with quartz and is typically characterized by a
010

dusty appearance. Microcline grains are characterized by cross-hatch twinned


with untwined crystals being rare. Thus, the determination of microcline grains
using a petrographic microscope is easier than orthoclase. In some cases Kfeldspar grains are affected by grain-coating carbonate (Appendix 6-4c).
Untwined microcline grains are generally similar to orthoclase and quartz grains
when they are present in thin section, and it is difficult to distinguish between
them. In general, K-feldspar occurs as subrounded to rounded grains.
Orthoclase and microcline typically have a cloudy appearance as a result of
alteration products (Boggs, 2009).
5-2-2-2 Plagioclase
Plagioclase occurs in most sandstone samples in the study area and is
characterised by its lath-shaped grains. It is composed of twinned and untwined
grains where the former are more common than latter (Appendix 6-5a-c and 66a-b). When there is no twinning in the plagioclase grains it is difficult to
distinguish from quartz. Untwined plagioclase is prevalent in the finer particles
because they break along cleavage planes (Basu, 1976). The feldspar grains
may include well developed zoning. Plagioclase grains are grey in colour and
are seen as euhedral in most samples. They are of volcanic origin. They are
present as angular to sub-angular grains and are optically positive to negative
under the microscope. Plagioclase grains typically show more partial to
extensive alteration than K-feldspar grains and they may be coated by
carbonate cement (Appendix 6-5b and 6-6a-b). Various sources of plagioclase
include volcanic and plutonic igneous rocks according to Boggs (2009).
Untwined plagioclase grains are less common than twinned plagioclase grains
in the study area and in some cases, they show extensive diagenetic alteration.
In general, detrital plagioclase is present in sandstones, deriving from basic and
intermediate lavas (Boggs, 2009).

5-2-3 Rock fragments


Rock fragments are generally common in sandstones (Boggs, 2009). They
comprise a variety of volcanic and sedimentary rock fragments but lack
metamorphic rock fragments in most units in the southern Sydney Basin. Rock
011

fragments are mostly unstable and can be good provenance indicators (Tucker,
1991).
5-2-3-1 Volcanic rock fragments
Four types of volcanic rock fragments are recognised in this study. They are
felsic, vitric, microlithic and lathwork (Appendix 6-6c and 6-7a). In both the
Illawarra Coal Measures and Narrabeen Group, carbonate-replaced fragments
are classified as volcanic rock fragments since these two units were
predominantly derived from a volcanic source which does not include limestone;
thus the carbonate fragments are most probably a replacement of volcanic rock
fragments. In the Hawkesbury Sandstone, they are classified as sedimentary
rock fragments since this unit is derived from the Lachlan Fold Belt which
includes limestone.
Felsic and vitric grains are the most common volcanic rock fragments.
Microlithic grains are rarely observed in some samples while lathwork grains are
very rare or missing in all samples. The identification of volcanic rock fragment
types depends on their texture. Volcanic rock fragments include silicic to
intermediate flows and tuff. Also, porphyritic and trachytic textures are common
in volcanic rock fragments, while original volcanic glass fragments are dark in
colour.
Felsic volcanic rock fragments are usually similar to chert grains, so it is often
difficult to distinguish between them. They occur as granular or foliated grains
and may include quartz and feldspar crystals. Silicic volcanic rocks are the
source of felsic grains. Clay, feldspar and quartz crystals are recorded in vitric
volcanic rock fragments which are sourced from glassy flows and tuffs. Glass
and altered glass is often included in the vitric volcanic rock fragments. The
microlithic volcanic rock fragments are less common than felsic and vitric grains
and include very small feldspar laths. A lava of intermediate composition was
the source of the microlithic grains which also show several textures, such as
pilotaxitic, trachytic and hyalopilitic.
In general, volcanic rock fragments are subrounded to rounded and are
characterised by grain size ranges from 20 to 1200 m. Carbonate cement
011

occurs in most samples as coatings on the margins of volcanic rock fragments


(Appendix 6-6c and 6-7a).
5-2-3-2 Sedimentary rock fragments
Sedimentary rock fragments consist of sandstone, siltstone and shale in the
Illawarra Coal Measures and Narrabeen Group (Appendix 6-7b-c and 6-8a). In
the Hawkesbury Sandstone, carbonate fragments are classified as sedimentary
rock fragments because of the reason given above.
Sandstone fragments are the most common sedimentary rock fragments. Sand
and silt fragments are characterized by fragmental textures and commonly
contain quartz and feldspar. Clay and silt size particles are common in the shale
fragments which are similar to vitric volcanic lithic grains, since both are finegrained. Thus, the distinction between shale fragments and vitric volcanic lithic
fragments is difficult and some may have been counted as igneous rock
fragments. Microgranular textures are widespread features for carbonate
fragments.
Sedimentary rock fragment grain sizes range from 22 to 1020 m and they are
sub-rounded to rounded. Also, carbonate cement is seen in thin section on the
margins of sedimentary rock fragments (Appendix 6-7b). Chemical weathering
may be the reason for the scarcity of sedimentary lithic grains in most samples
(Caracciolo et al., 2011).

5-2-4 Chert
Chert grains are common in the southern Sydney Basin and have a
microgranular texture (Appendix 6-8b-c and 6-9a). In the Illawarra Coal
Measures and Narrabeen Group chert grains are counted as volcanic rock
fragments because it is difficult to distinguish between felsic volcanic rock
fragments and chert grains (Bai and Keene, 1996). Also, chert grains are similar
to polycrystalline grains in the thin section. Uniform crystal size of chert is
important for distinguishing chert grains from polycrystalline quartz grains. In the
Hawkesbury Sandstone, chert grains occur clearly (Appendix 6-9a) and are
classified separately.
011

Chert grain size ranges from 15 to 950 m and may be coated partially by
carbonate cement (Appendix 6-8b).

5-2-5 Mica
In this study, mica includes biotite and muscovite, with the latter being more
common than the former (Appendix 6-9b-c and 6-10a). Muscovite grains are
deformed, and usually occur as bent flakes between quartz grains because of
mechanical compaction. They are sourced either from metamorphic or
deformed assemblages (Michaelsen and Henderson, 2000; Ullah et al., 2006).
Also, Folk (1968), Tucker (1991) and Boggs (2009) showed that metamorphic
rocks are the most common source for muscovite flakes. Also, they showed that
muscovite is more stable than biotite. In this study, physical compaction is
mainly indicated by bent mica flakes. Muscovite grains exist as inclusions in
some quartz grains but are typically aligned parallel to bedding laminae.
Muscovite grains range from green in colour to colourless, with weak
birefringence.
In a few samples, muscovite grains are squeezed by adjacent grains and they
vary in size between fine- to coarse-grained in the thin section. Fine-grained
sediments are richer in detrital muscovite than coarse-grained sediments
because the sheet-like shape of muscovite grains means they are winnowed
out of sandstone (Doyle et al., 1983; Boggs, 2009).

5-2-6 Heavy minerals


Galloway (1972) compared the presence of rutile in the Hawkesbury Sandstone
and the Narrabeen Group in terms of colour. Rutile in the Narrabeen Group is
dark brown to opaque, whereas in the Hawkesbury Sandstone it is red and light
honey brown. Also, he showed that the size of rutile and zircon is larger in the
Hawkesbury sandstone than in the Narrabeen Group. According to Galloway
(1972), rutile is more common in the Hawkesbury Sandstone than in the
Narrabeen Group whereas zircon is more common in the Narrabeen Group
than in the Hawkesbury Sandstone.

011

In this study, the heavy minerals are present in trace amounts and consist of
hornblende, rutile, zircon and tourmaline (Appendix 6-10b-c). The last three
minerals indicate plutonic source rocks according to Wanas and Abdel-Maguid,
(2006) and are also stable heavy minerals (Boggs, 2009). Heavy minerals are
angular to sub-rounded and are observed in all samples as fine grains. Some
heavy minerals are present as inclusions in some of the quartz grains. Most
heavy minerals show parallel extinction.
In the current study, tourmaline is mainly brown whereas hornblende varies
from green to brown. Zircon is very small and varies in colour between yellow to
grey whereas rutile is red to brownish red. Rutile is euhedral to anhedral and
has irregular grain boundaries. Coarse-grained sediments and fine-grained
sediments have very angular rutile and rounded rutile grains, respectively. Also,
tourmaline grains are angular-subangular and rounded grains and the former
are more common. In contrast, the rounded zircon grains are more common
than angular-subangular zircon grains and the latter are larger crystals.
Tourmaline and zircon are harder than rutile. Tourmaline and zircon show
mainly strong birefringence. Hornblende is observed as having moderate to
strong birefringence whereas rutile has extreme birefringence.

5-2-7 Clay minerals


In the Hawkesbury Sandstone, Standard (1969) indicated that the argillaceous
matrix included illite, kaolinite and mixed-layered clay minerals. Dickite also
occurs in the Hawkesbury Sandstone as rouleaux and vermicular crystals
(Bayliss et al., 1965). Spry (2000) showed that clay minerals characteristically
occur as pellets, pore fillings, films and books of dickite in the Hawkesbury
Sandstone. In general, sandstones typically include both detrital and authigenic
clay minerals and, according to Tucker (1991), it is known that climate,
weathering and source area are determined mostly by the detrital clay minerals.
In this study, clay minerals are most common in the finer-grained rock types and
are the main components of the <2 m clay size fraction. The description of clay
minerals using the polarizing microscope is somewhat difficult due their very
fine grain size. X-ray diffraction and scanning electron microscope were used as
011

techniques to describe and determine the percentages of clay minerals in this


study. Kaolinite and chlorite are noticeable using the polarizing microscope in
some samples. Kaolinite is observed as small cavity-filling platelets and is
present in the matrix of most samples. Also, it occurs as a dull grey mineral and
is recorded as deformed patches among rigid grains.
Chlorite occurs as grain coatings and as pore-fillings. Grain coating chlorite is
represented by a greenish colour; although in some cases, it is reddish in colour
where oxidised. Distinguishing between clays and cement by the microscope is
difficult, thus clays were not determined accurately by this means. Silts and
muds commonly include kaolinite and rarely chlorite, illite and smectite. This
indicates harsh weathering and temperature (van de Kamp, 2010; Ghazi and
Mountney, 2011).

5-2-8 Matrix
Matrix is recognised in sandstones as pore-fillings and is always fine-grained.
Clay minerals are found as clay matrix in thin sections, whether detrital or
authigenic, but most of the matrix is detrital clay. Silt-sized quartz is also
included in the matrix. Also, fine components such as quartz, feldspar, clays
and mica may be included in the matrix in some cases. Matrix is common in
samples which are very poorly sorted. It is difficult to differentiate between
matrix and some cements in the thin sections. Also, there is a complexity in
distinguishing between matrix and pseudomatrix in thin section. Thus, the
percentage of matrix may be increased due to the pseudomatrix being counted
as matrix. In some cases, lithic grains are deformed and unclear, thus they may
also be counted as matrix. These cases are uncommon. In general, compaction
and unstable grain alteration are indicated as the source for matrix in
sandstones according to Tucker (1991).

019

5-3 Lateral and vertical variations of composition


Illawarra Coal Measures
5-3-1 Introduction
In the Illawarra Coal Measures, about 142 samples were selected from outcrop
and from six wells including EAW 30, EAW 156, EDEN 124, EDEN 125, EDEN
126 and EDEN 127. These samples include seventy one samples of sandstone,
twenty one samples of siltstone, thirty nine samples of shale, nine samples of
tuff, one sample of coal and one sample of carbonate cement. Petrographic
microscope studies were completed for seventy six samples of siltstone, finegrained sandstone, medium-grained sandstone and coarse-grained sandstone,
while X-ray diffraction analysis was used to study one hundred and twenty two
samples of fine to medium-grained sandstone, fine-grained sandstone,
siltstone, shale, tuff, coal and carbonate cement. Also, twenty eight samples
were described using scanning electron microscope analysis.
The petrology of the Illawarra Coal Measures has previously been studied by a
number of authors (e.g. Cusack, 1991; Bamberry, 1992; Jesus, 1999; Grevenitz
et al., 2003). Bamberry (1992) described the petrography in the Sydney
Subgroup showing quartz and lithic suites. He noted that the sandstones of
Marrangaroo and Blackmans Flat Conglomerates and the Wilton Formation
represent the quartz suite. The quartz suite is also characterized in the
uppermost Erins Vale Formation from the Thirroul area and sandstone in the
basal Wilton Formation. On the other hand, all sandstones which overlie the
Tongarra Coal represent the lithic suite (Bamberry, 1992). In the Wongawilli
Seam, petrography and composition of kaolinitic claystones were described by
Loughnan (1971).

5-3-2 Thin section


5-3-2-1 Detrital composition and texture
Point count size analysis showed that mean grain size ranges from siltstone
(58 m) to coarse sandstone (854 m; Appendix 5-7 and 8-14). The Loddon
019

Sandstone has the largest grain size at an average of 464.2 m whereas the
Bargo Claystone has the lowest grain size at an average of 55.7 m (Appendix
8-13). Sorting varies from 0.5 to 1.1 (Appendix 5-1). There is a moderate
positive correlation between sorting and grain size in the Kembla Sandstone (r 2
= 0.4; Fig. 5-1). Also, there is a good negative correlation between sorting and
grain size in the Lawrence Sandstone (r2 = 0.6; Fig. 5-2). Thus, sorting
increases with decreasing grain size in this unit. The sorting and roundness are
mainly low in the litharenites (von Eynatten and Gaupp, 1999). Some samples
include both rounded and angular grains in one sample, which is evidence of
mixing grains with very different transport histories (Maher et al., 2004).

y = 1127.7x - 367.29
R = 0.3618

900

Grain size (micron)

800
700
600
500
400
300
200
100
0
0.0

0.2

0.4

0.6

0.8

1.0

Sorting (phi)

Figure 5-1: A moderate positive correlation occurs between


sorting and grain size in the Kembla Sandstone.

017

y = -548.85x + 573.33
R = 0.5643

450
400
Grain size (micron)

350
300
250
200
150
100
50
0
0.0

0.2

0.4

0.6
Soring (phi)

0.8

1.0

1.2

Figure 5-2: Good negative correlation between sorting and grain


size in the Lawrence Sandstone.

001

The framework composition of the Illawarra Coal Measures was classified using
the QFL diagram according to Folk (1968). The Illawarra Coal Measures
consists mostly of litharenite with rare sublitharenite (Q29.5%, F2.9%, R67.6%; Fig. 53a). The framework composition of the units of the Illawarra Coal Measures was
recorded in Figures Fig. 5-3b, 5-4a, b, 5-5a, b and 5-6. The average contents of
quartz, feldspar and rock fragments in units of the Illawarra Coal Measures are
recorded in Table 5-1. The origin of the Illawarra Coal Measures was
recognized based on the diagram from Dickinson (1985). The Illawarra Coal
Measures were derived from lithic recycled to transitional recycled and
quartzose recycled provenance areas (Qm25.1%, F3.2%, Lt71.8%; Fig. 5-7a). The
origin of the units of the Illawarra Coal Measures was recorded in Figs 5-7b, 58a, b, 5-9a, b and 5-10. The averages of monocrystalline quartz, feldspar and
rock fragment for units of the Illawarra Coal Measures were recorded in Table
5-1. The classification of samples in the Illawarra Coal Measures according to
Folk (1968) and Dickinson (1985) is recorded in Appendix 5-4.
Table 5-1: Average classification data for Folk (1968) and Dickinson
(1985) based on petrographic analyses of rocks from the southern
Sydney Basin.
Formation

Average (F : Q : R)

Average (F : Qm : Lt)

Hawkesbury Sandstone

32

0.3 : 90.1 : 9.7

0.3 : 89.7 : 10

Newport Formation

0.6 : 89 : 10.4

0.7 : 88.3 : 11

Bald Hill Claystone

1.7 : 53.2 : 45.1

2.8 : 22.7 : 74.5

Bulgo Sandstone

12

1.2 : 84.3 : 14.4

1.4 : 82.4 : 16.2

Stanwell Park Claystone

3.7 : 33.4 : 62.9

4.4 : 21: 74.5

Scarborough Sandstone

19

2.1 : 47.4 : 50.5

2.3 : 42.2 : 55.5

Wombarra Claystone

43

1.2 : 45.4 : 53.4

1.3 : 40.6 : 58

Coalcliff Sandstone
Narrabeen Group

20

0.9 : 39 : 60.1

1 : 34.3 : 64.7

132

1.4 : 50.3 : 48.3

1.5 : 45.5 : 52.9

Loddon Sandstone

19

2.6 : 20.3 : 77.1

2.8 : 14.9 : 82.3

Lawrence Sandstone

18

3.6 : 39.3 : 57

3.9 : 35.3 : 60.9

Unnamed Member Two

2.8 : 52.8 : 44.4

3.1 : 47.4 : 49.5

Unnamed Member Three

5.3 : 40.7 : 54

5.8 : 36.1 : 58.1

Wongawilli Coal

1.9 : 27.6 : 70.6

1.9 : 24.9 : 73.2

Kembla Sandstone

19

2.1 : 19.6 : 78.3

2.2 : 16.3 : 81.5

Allans Creek Formation

2.7 : 36.2 : 61

2.9 : 32.5 : 64.6

Dark Forest Sandstone

3.8 : 31.6 : 64.6

4.2 : 24.7 : 71

Bargo Claystone

0.9 : 79.7 : 19.4

0.9 : 79.5 : 19.5

Wilton Formation
Illawarra Coal Measures

2.9 : 41.9 : 55.2

3.1 : 38.9 : 58

76

2.9 : 29.5 : 67.6

3.2 : 25.1 : 71.8

000

Quartz
arenite

a)

Illawarra Coal Measures


Sublitharenite

Subarkose

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite

Bargo Claystone
Wilton Formation

Quartz
arenite

b)

Sublitharenite

Subarkose

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite
R

Figure 5-3: (a) Classification of the Illawarra Coal Measures. (b)


Classification of the Wilton Formation and Bargo Claystone (after Folk,
1968). Q=quartz, F=feldspar R=rock fragment.

001

Allans Creek Formation

Quartz
arenite

a)

Darkes Forest Sandstone


Sublitharenite

Subarkose

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite

Quartz
arenite

b)

Kembla Sandstone
Sublitharenite

Subarkose

Arkose

Litharenite

Lithic
arkose

Feldspathic
litharenite

Figure 5-4: (a) Classification of the Darkes Forest Sandstone and


Allans Creek Formation. (b) Classification of the Kembla
Sandstone (after Folk, 1968). Q=quartz, F=feldspar R=rock
fragment.

001

Quartz
arenite

a)

Unnamed Two Member


Unnamed Three Member
Wongawilli Coal
Sublitharenite

Subarkose

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite

Lawrence Sandstone
Q

Quartz
arenite

b)

Sublitharenite
Subarkose

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite

Figure 5-5: (a) Classification of the Wongawilli Coal, Unnamed


Three Member and Unnamed Member Two. (b) Classification of
the Lawrence Sandstone (after Folk, 1968). Q=quartz, F=feldspar
R=rock fragment.

001

Loddon Sandstone
Q

Quartz
arenite

Sublitharenite
Subarkose

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite

Figure 5-6: Classification of the Loddon Sandstone (after Folk,


1968). Q=quartz, F=feldspar R=rock fragment.

001

Qm

a)

Illawarra Coal Measures


Craton
interior

Quartzose
recycled

Transitional
continental
Transitional
recycled

Mixed
Basement
uplift

Dissected Ac
Lithic
continental

Lt

F
Undissected

Ac

Qm
Bargo Claystone

b)

Wilton Formation

Craton
interior

Quartzose
recycled

Transitional
continental

Transitional
recycled

Mixed
Basement
uplift

Dissected Ac
Lithic
continental

Lt
Undissected

Ac

Figure 5-7: (a) The provenance of the Illawarra Coal Measures.


(b) The provenance of the Wilton Formation and Bargo
Claystone (after Dickinson, 1985). Qm=monocrystalline quartz,
F=feldspar, Lt=rock fragment + chert.

001

Allans Creek Formation

Qm

a)

Darkes Forest Sandstone

Craton
interior

Quartzose
recycled

Transitional
continental

Transitional
recycled

Mixed
Basement
uplift

Dissected Ac
Lithic
continental

Lt
Undissected

Ac

Kembla Sandstone

Qm

b)
Craton
interior

Quartzose
recycled

Transitional
continental

Transitional
recycled

Mixed
Basement
uplift

Dissected Ac
Lithic
continental

Lt
Undissected

Ac

Figure 5-8: (a) The provenance of the Darkes Forest Sandstone and
Allans Creek Formation. (b) The provenance of the Kembla Sandstone
(after Dickinson, 1985). Qm=monocrystalline quartz, F=feldspar,
Lt=rock fragment + chert.
009

Qm

a)

Unnamed Two Member


Unnamed Three Member

Craton
interior

Wongawilli Coal
Quartzose
recycled

Transitional
continental
Mixed
Basement
uplift

Transitional
recycled

Dissected Ac
Lithic
continental

Lt
Undissected

Ac

Qm

b)

Lawrence Sandstone

Craton
interior

Quartzose
recycled

Transitional
continental

Mixed
Basement
uplift

Transitional
recycled

Dissected Ac
Lithic
continental

Lt
Undissected

Ac

Figure 5-9: (a) The provenance of the Wongawilli Coal, Unnamed


Three Member and Unnamed Two Member. (b) The provenance
of the Lawrence Sandstone (after Dickinson, 1985).
Qm=monocrystalline quartz, F=feldspar, Lt=rock fragment + chert.

009

Qm

Loddon Sandstone

Craton
interior
Quartzose
recycled

Transitional
continental

Mixed
Basement
uplift

Transitional
recycled

Dissected Ac
Lithic
continental

Lt
Undissected

Ac

Figure 5-10: The provenance of the Loddon Sandstone (after


Dickinson, 1985). Qm=monocrystalline quartz, F=feldspar,
Lt=rock fragment + chert.

007

5-3-2-1-1 Quartz
Quartz occurs in the Illawarra Coal Measures ranging in abundance from 1.2%
to 49.4% (Appendix 5-7 and 8-2). Detrital quartz is recorded in the Wilton
Formation at an average of 14.8%, increasing to 17.3% in the Bargo Claystone
(Appendix 5-13, 5-14 and 8-1). Up-sequence, quartz decreases from 17.3% in
the Bargo Claystone to 11.4% in the Kembla Sandstone which contains the
lowest percentage of quartz (Appendix 5-13, 5-14 and 8-1). These decreases
are related to an increased amount of lithic grains in these units. Up-sequence,
quartz content fluctuates between 12.8% and 19% from the Wongawilli Coal to
the Unnamed Member Two but then decreases to 13.7% in the Loddon
Sandstone (Appendix 5-13, 5-14 and 8-1). In general, more quartz is present in
siltstone (18.1%) than in sandstone (14.2%; Appendix 5-10).
Monocrystalline quartz is more widespread compared to polycrystalline quartz
in all units of the Illawarra Coal Measures (Appendix 5-13, 5-14 and 8-1).
Monocrystalline quartz grains fluctuate between 8.9% and 17.1% from the
Wilton Formation to the Loddon Sandstone whereas polycrystalline quartz
grains fluctuate between 0.2% and 4% from the Wilton Formation to the Allans
Creek Formation, decreasing to 2.1% in the overlying Wongawilli Coal
(Appendix 5-13, 5-14 and 8-1). Up-sequence, they increase to 3.7% in the
Unnamed Member Two but then fluctuate between 3.4% and 4.3% in the
Lawrence Sandstone and the Loddon Sandstone (Appendix 5-13, 5-14 and 81). In two samples from the Kembla Sandstone, polycrystalline quartz was
absent (Appendix 5-7).
A high negative correlation is recorded between quartz and grain size in the
Kembla Sandstone (r = 0.7; Fig. 5-11a). Also, a moderate positive correlation is
determined between quartz and feldspar in the Lawrence Sandstone (r = 0.4;
Fig. 5-11b). There is no correlation between monocrystalline quartz and grain
size and also there is no correlation between polycrystalline quartz and grain
size in the Illawarra Coal Measures. In the Kembla Sandstone, quartz is
moderately negatively correlated with lithic grains (r = 0.3; Fig. 5-12)

011

a
y = -27.698x + 679.08
R = 0.709

900

Grain Size (micron)

800
700
600
500
400
300
200
100
0
0

10

15

20

25

Quartz (%)

y = 0.0438x + 0.9
R = 0.3961

3.5
3

Feldspar (%)

2.5
2
1.5
1
0.5
0
0

10

20

30

40

50

60

Quartz (%)

Figure 5-11: a) High negative correlation between quartz and


grain size in the Kembla Sandstone, b) Moderate positive
correlation between quartz and feldspar in the Lawrence
Sandstone.

010

y = -1.6018x + 72.684
R = 0.2815

100
90

Lithic grains (%)

80
70
60
50
40
30
20
10
0
0

10

Quartz (%)

15

20

25

Figure 5-12: Moderate negative correlation between quartz and


lithic grains in the Kembla Sandstone.

5-3-2-1-2 Feldspar
Feldspar is present in the Illawarra Coal Measures but the percentages do not
exceed 3.2% (Appendix 5-7 and 8-4). The Wilton Formation has feldspar at an
average of 1.1% but it then decreases to 0.2% in the Bargo Claystone as the
lowest percentage (Appendix 5-13, 5-14 and 8-3). Up-sequence, it increases to
1.8% in the Darkes Forest Sandstone but then decreases to 1.1% in the
Wongawilli Coal (Appendix 5-13, 5-14 and 8-3). Feldspar grains fluctuate
between 1% and 1.8% from the Unnamed Member Three to the Loddon
Sandstone (Appendix 5-13, 5-14 and 8-3). The Darkes Forest Sandstone and
Loddon Sandstone include the highest percentages of feldspar at average of
1.8% (Appendix 5-13, 5-14 and 8-3). Sandstone contains feldspar at an
average of 1.6% whereas siltstone has feldspar at an average of 1% in the
Illawarra Coal Measures (Appendix 5-10).
Plagioclase grains form 0.8% and 0.2% in the Wilton Formation and Bargo
Claystone, respectively, and then reach up to 1.6% in the Darkes Forest
Sandstone, decreasing to 0.8% in the Wongawilli Coal (Appendix 5-13, 5-14
and 8-3). Up-sequence, they fluctuate between 0.9% and 1.4% from the
Unnamed Member Three to the Loddon Sandstone (Appendix 5-13, 5-14 and 8011

3). K-feldspar grains occur at 0.2% in the Wilton Formation and then disappear
in the Bargo Claystone (Appendix 5-13, 5-14 and 8-3). Up-sequence, K-feldspar
increases to 0.3% in the Wongawilli Coal, decreasing to 0.1% in the Unnamed
Member Two and then it varies between 0.3% and 0.4% in the Lawrence
Sandstone and Loddon Sandstone (Appendix 5-13, 5-14 and 8-3).
In all units of the Illawarra Coal Measures, plagioclase is more common than Kfeldspar (Appendix 5-13, 5-14). K-feldspar occurs in coarse-grained samples
more than in fine-grained samples. No correlation is recorded between total
feldspar, K-feldspar and plagioclase and grain size in the Illawarra Coal
Measures.
5-3-2-1-3 Lithic grains
Lithic grains are composed of rock fragments and chert in the Illawarra Coal
Measures. Chert grains are included within rock fragments because
distinguishing between felsic volcanic rock fragments and chert grains is not
easy. So, chert grains are classified with volcanic rock fragments in this unit.
Lithic grains are present in the Illawarra Coal Measures and range from 1.9% to
90.5% (Appendix 5-7 and 8-6). They are observed in the Wilton Formation
(24.7%) whereas the lowest percentages are determined in the Bargo
Claystone (4.2%) and then they increase up section to 54.4% in the Kembla
Sandstone (Appendix 5-13, 5-14 and 8-5). These increases are associated with
a decreased amount of quartz grains in these units. Up-sequence, lithic grains
decrease to 16% in the Unnamed Member Two but then increase to 52.4% in
the Loddon Sandstone (Appendix 5-13, 5-14 and 8-5). In the Illawarra Coal
Measures, the coarse-grained samples contain more lithic grains than in the
fine-grained samples. Lithic grains are rare in siltstone samples (6.1%) but, they
are prevalent in the sandstone (44.8%; Appendix 5-10).
Lithic grains show good positive correlation with grain size in the Illawarra Coal
Measures (r = 0.5; Fig. 5-13a). There is a moderate positive correlation
between lithic grains and grain size in the Kembla Sandstone (r = 0.4; Fig. 5-

011

13b) and in the Lawrence Sandstone (r = 0.5; Fig. 5-14a). This correlation is
very low in the Loddon Sandstone (r = 0.01; Fig. 5-14b).

y = 5.8031x + 23.593
R = 0.4848

900
800
Grain size (micron)

700
600
500
400
300
200
100
0
0

20

40
60
Lithic grains(%)

80

100

y = 6.9085x - 13.014
R = 0.4021
900

Grain size (micron)

800
700
600
500
400
300
200
100
0
0

20

40

60

80

100

Lithic grains(%)

Figure 5-13: (a) and (b) Moderate positive correlation between


lithic grains and grain size in the Illawarra Coal Measures and
Kembla Sandstone.

011

y = 4.0612x + 62.12
R = 0.5559

450
400

Grain size (micron)

350
300
250
200
150
100
50
0
0

20

40

60

80

Lithic grains(%)

y = 1.8154x + 282.19
R = 0.0124
700
600
Grain size (%)

500
400
300
200
100
0
0

20

40

60

80

Lithic grains(%)

Figure 5-14: (a) There is a good positive correlation between


lithic grains and grain size in the Lawrence Sandstone, (b) Lithic
grains show low positive correlation with grain size in the
Loddon Sandstone.

011

5-3-2-1-3-1 Volcanic Rock Fragment plus chert and carbonate fragments


The volcanic rock fragment plus chert and carbonate fragments occur in the
Wilton Formation at an average of 23.5% and then decreases to 4.2% in the
Bargo Claystone (Appendix 5-13 and 5-14). Up-sequence, they increase to
52.8% in the Kembla Sandstone and then decreases to 16% in the Unnamed
Member Two, increasing to 48.6% in the Loddon Sandstone (Appendix 5-13
and 5-14). More volcanic rock fragments plus chert and carbonate fragments
occur in sandstone (42.7%) than in siltstone samples (6%; Appendix 5-10).
5-3-2-1-3-2 Sedimentary Rock Fragments
The Wilton Formation includes sedimentary rock fragments at an average of
1.1% (Appendix 5-13 and 5-14). Sedimentary rock fragments disappear in the
Bargo Claystone and then appear in the Darkes Forest Sandstone (1.1%;
Appendix 5-13 and 5-14). They fluctuate between 0.7% and 3.5% from the
Allans Creek Formation to the Unnamed Member Three but then disappear in
the Unnamed Member Two (Appendix 5-13 and 5-14). They occur in the
Lawrence Sandstone (1.2%) and Loddon Sandstone (3.9%), the latter having
the highest percentage (Appendix 5-13 and 5-14).
5-3-2-1-3-3 Chert
Chert grains are recorded in amounts up to 90.5% (Appendix 5-7 and 8-6).
They are present in the Wilton Formation at an average of 16.7% (Appendix 513, 5-14 and 8-5). They decrease up sequence to 4.2% in the Bargo Claystone
and then increase to 45.7% in the Kembla Sandstone which has the highest
percentage (Appendix 5-13, 5-14 and 8-5). Up-sequence, they decrease to 3%
in the Unnamed Member Two but then increase to 42.5% in the Loddon
Sandstone (Appendix 5-13, 5-14 and 8-5). Chert grains exist in the sandstone
at an average of 37.4% whereas it is less common in siltstone samples at an
average of 4.5% (Appendix 5-10).
A moderate positive correlation is shown between chert and grain size in the
Illawarra Coal Measures (r = 0.4; Fig. 5-15a). A good positive correlation exists
between chert and grain size in the Lawrence Sandstone (r = 0.5; Fig. 5-15b).
011

y = 5.8597x + 62.259
R = 0.4178

900
800

Grain size (micron)

700
600
500
400
300
200
100
0
0

20

40

60

80

100

Chert (%)

y = 4.3607x + 72.168
R = 0.5348

450
400
Grain size (micron)

350
300
250
200
150
100
50
0
0

20

40

60

80

Chert (%)

Figure 5-15: a) A moderate positive correlation is observed


between chert and grain size in the Illawarra Coal Measures, b) A
good positive correlation exists between chert and grain size in the
Lawrence Sandstone.

019

5-3-2-1-4 Muscovite
Amounts of muscovite vary between 0 and 4.2% (Appendix 5-7 and 8-8) with
average of 0.7% and 1.1% in both sandstone and siltstone respectively
(Appendix 5-10). More muscovite is recorded in the Wilton Formation (0.6%)
than in the Bargo Claystone (0.2%; Appendix 5-13, 5-14 and 8-7). It increases
to 0.8% in the Allans Creek Formation and then fluctuates between 0% and
1.1% from the Kembla Sandstone to the Loddon Sandstone (Appendix 5-13, 514 and 8-7). The Unnamed Member Two and Loddon Sandstone contain the
highest percentage, at an average of 1.1% (Appendix 5-13, 5-14 and 8-7).
Muscovite is moderately negatively correlated with grain size in the Loddon
Sandstone (r = 0.4; Fig. 5-16).
5-3-2-1-5 Heavy minerals:
Heavy minerals are very minor in the Illawarra Coal Measures, varying between
0% and 2.7% (Appendix 5-7 and 8-8). They occur in the Wilton Formation and
Bargo Claystone at average of 0.7% and 0.2% respectively, increasing to 1.2%
in the Darkes Forest Sandstone (Appendix 5-13, 5-14 and 8-7). Up-sequence,
the percentages of heavy minerals decrease to 1% in the Kembla Sandstone
and then increase to 1.3% in the Wongawilli Coal (Appendix 5-13, 5-14 and 87). Up-sequence, they decrease to 0.3% in the Unnamed Member Two but then
increase to 1% in the Loddon Sandstone (Appendix 5-13, 5-14 and 8-7). The
Wongawilli Coal includes the highest percentage of heavy minerals at an
average of 1.3% (Appendix 5-13, 5-14 and 8-7). Heavy minerals are identified in
the Bargo Claystone at an average of 0.2% but these values are lower than
those in other units (Appendix 5-13, 5-14 and 8-7). The Wongawilli Coal
contains hornblende and tourmaline (Appendix 5-13 and 5-14).
In the Illawarra Coal Measures, hornblende is the most common heavy mineral,
whereas rutile and zircon are not observed in most units of the Illawarra Coal
Measures (Appendix 5-14 and 5-15).

019

y = -57.997x + 440.25
R = 0.3839

700
Grain size (micron)

600
500
400
300
200
100
0
0

Muscovite (%)

Figure 5-16: A moderate negative correlation is present between


muscovite and grain size in the Loddon Sandstone.

5-3-2-1-6 Chlorite
Chlorite is rare to absent in some units of the Illawarra Coal Measures
(Appendix 5-7 and 8-8). The Loddon Sandstone includes chlorite at average of
0.2% as the highest percentage in the coal measures (Appendix 5-13, 5-14 and
8-7).
5-3-2-1-7 Matrix
Matrix is widespread in most units of the Illawarra Coal Measures and ranges
from 0% to 78.8% (Appendix 5-7 and 8-12). Matrix is present in the Wilton
Formation (37.7%, Appendix 5-13, 5-14 and 8-11). The Bargo Claystone is rich
in matrix with an average of 39% and it then decreases to 13% in the Darkes
Forest Sandstone (Appendix 5-13, 5-14 and 8-11). Up-sequence, it is observed
in the Allans Creek Formation and Kembla Sandstone at averages of 29.9%
and 17.1% and then increases to 40.3% in the Unnamed Member Three
(Appendix 5-13, 5-14 and 8-11). Up-sequence it decreases gradually to 20.2%
in the Loddon Sandstone (Appendix 5-13, 5-14 and 8-11).

017

A moderate negative correlation occurs between matrix and grain size in the
Lawrence Sandstone (r = 0.4; Fig. 5-17).

y = -3.4363x + 296.12
R = 0.4047

450
400
Grain size (micron)

350
300
250
200
150
100
50
0
0

20

40

60

80

100

Matrix (%)

Figure 5-17: A moderate negative correlation occurs between


matrix and grain size in the Lawrence Sandstone.

5-3-3 X-ray diffraction (XRD)


In this study, fine- to medium-grained sandstone, siltstone, shale, mudstone,
claystone, tuff, coal and siderite samples were also analysed by X-ray
diffraction.
5-3-3-1 Quartz
X-ray diffraction indicated that quartz ranges from 1.7% to 77.4% in the finegrained samples in the Illawarra Coal Measures (Appendix 5-15 and 8-16).
Quartz content fluctuates between 23.8% and 39.9% from the Wilton Formation
to the American Coal Member but then increase to 51.6% in the Wongawilli
Coal (Appendix 5-21, 5-22 and 8-15). Up-sequence, quartz occurs in the
Unnamed Member Two and the Unnamed Member Three at averages of 40.1%
and 46.3%, decreasing to 18.6% in the Balgownie Coal (Appendix 5-21, 5-22
and 8-15). Up-sequence, it reaches 41.9% in the Loddon Sandstone (Appendix
5-21, 5-22 and 8-15). The percentages of quartz are similar in sandstone
011

(40.9%) and siltstone (42.7%) and are greater than that in shale (31.2%;
Appendix 5-18).
A moderate negative correlation is observed between quartz and clay minerals
in the Illawarra Coal Measures as a whole (r = 0.5; Fig. 5-18), in the Kembla
Sandstone (r = 0.4; Fig. 5-19a) and in the Lawrence Sandstone (r = 0.4; Fig.
5-19b). In addition, quartz shows strong negative correlation with total clay
minerals (r = 0.9; Fig. 5-20a) and a good negative correlation with kaolinite (r =
0.6; Fig. 5-20b) in the Loddon Sandstone.

100
y = -0.887x + 67.638
R = 0.5247

90

Clay minerals (%)

80
70
60
50
40
30
20
10
0
0

20

40

60

80

100

Quartz (%)

Figure 5-18: A moderate negative correlation occurs between


quartz and clay minerals in the Illawarra Coal Measures (XRD
results).

010

a
45

y = -0.5468x + 55.8
R = 0.448

40
Clay minerals (%)

35
30
25
20
15
10
5
0
0

10

20

30

40

50

60

70

Quartz (%)

b
50
y = -0.5908x + 51.786
R = 0.387

45

Clay minerals (%)

40
35
30
25
20
15
10
5
0
0

10

20

30

40

50

60

70

80

Quartz (%)

Figure 5-19: (a) and (b) These are a moderate negative


correlation between quartz and clay minerals in the Kembla
Sandstone and the Lawrence Sandstone respectively (XRDresults).

011

a
60

y = -1.0866x + 78.021
R = 0.9157

Clay minerals (%)

50
40
30
20
10
0
0

10

20

30

40

50

60

Quarz (%)

y = -0.7027x + 33.501
R = 0.5713

30
25
Kaolinite (%)

20
15
10
5
0
0

10

20

30

40

50

60

Quartz (%)

Figure 5-20: (a) Quartz show high negative correlation with clay
minerals in the Loddon Sandstone (XRD results), (b) There is a
good negative correlation between quartz and kaolinite in the
Loddon Sandstone (XRD results).

011

5-3-3-2 Feldspar
X-ray diffraction demonstrated that feldspar is included in the fine-grained
samples varying between 1.1% and 28.4% (Appendix 5-15 and 8-18). Feldspar
gradually increases from 8.5% in the Wilton Formation to 16.9% in the Darkes
Forest Sandstone (Appendix 5-21, 5-22 and 8-17). Feldspar is observed in the
Allans Creek Formation, American Creek Coal Member, Kembla Sandstone and
the Wongawilli Coal at averages of 12.2%, 11%, 13.3% and 8.6% respectively
(Appendix 5-21, 5-22 and 8-17). Up-sequence, it increases to 16.5% in the
Laurence Sandstone but then decreases to 9% in the Balgownie Coal,
increasing to 13.7% in the Loddon Sandstone (Appendix 5-21, 5-22 and 8-17).
Plagioclase grains are present in the Wilton Formation and the Tongarra Coal at
6% and 5.5% and then increase to 14.9% in the Darkes Forest Sandstone
(Appendix 5-21, 5-22 and 8-17). Up-sequence, they fluctuate between 6.2%
and 10.2% from the Allans Creek Formation and the Wongawilli Coal,
increasing to 13.7% in the Lawrence Sandstone (Appendix 5-21, 5-22 and 817). Up-sequence, they are recorded at averages of 6% and 10.9% in the
Balgownie Coal and the Loddon Sandstone (Appendix 5-21, 5-22 and 8-17). Kfeldspar grains are 2.6% in the Wilton Formation but then increase to 3.1% in
the Tongarra Coal, decreasing to 2% in the Darkes Forest Sandstone
(Appendix 5-21, 5-22 and 8-17). Up-sequence, these grains are fluctuated
between 2.3% and 4.2% from the Allans Creek Formation to the Loddon
Sandstone (Appendix 5-21, 5-22 and 8-17).
Orthoclase is missing in the Darkes Forest Sandstone, American Creek Coal
Member, Unnamed Member Two and Loddon Sandstone (Appendix 5-15).
Feldspar occurs in sandstone, siltstone and shale at averages of 15.3%, 13.8%
and 10.4% (Appendix 5-18).
5-3-3-3 Clay minerals
Bamberry (1992) showed that the composition of clay minerals is variable in
both the quartz and lithic suites. The major clay constituent of the sandstone is

011

illite in both suites, whereas the subordinate matrix components are kaolinite
and mixed-layer clays.
In this study, clay minerals are present throughout the succession, particularly
kaolinite and mixed-layer illite-smectite. The content of clay minerals ranges
from 1.2% to 94.9% (Appendix 5-15 and 8-16). Clay minerals increase upsequence from 43.2% in the Wilton Formation to 44.5% in the Bargo Claystone
but then decrease to 21% in the Darkes Forest Sandstone (Appendix 5-21, 5-22
and 8-15). Up-sequence, they increase to 43% in the American Creek Coal
Member. Up-sequence, the contents of clay minerals fluctuate between 26.4%
and 63.4% from the Kembla Sandstone to Loddon Sandstone (Appendix 5-21,
5-22 and 8-15). The highest percentages of clay minerals were determined in
the Balgownie Coal at an average of 63.4% whereas the smallest percentage
exists in the Darkes Forest Sandstone at an average of 21% (Appendix 5-21, 522 and 8-15).
Clay minerals in the Illawarra Coal Measures include illite, kaolinite, dickite,
chlorite and mixed-layer illite-smectite. Kaolinite and mixed-layer illite smectite
are the most common clay minerals in the Illawarra Coal Measures, whereas
dickite and chlorite are uncommon (Appendix 5-15). Shale samples have more
clay minerals than sandstone and siltstone samples (Appendix 5-18).
5-3-3-4 Muscovite/illite
The, percentages of muscovite are generally low ranging from 0 to 11.1% in the
Illawarra Coal Measures (Appendix 5-15 and 8-20). Muscovite exists in the
Wilton Formation (2.6%) but is less common than in the tuff beds from the
Tongarra Coal (4.4%) and in the Bargo Claystone (5.6%; Appendix 5-21, 5-22
and 8-19). It then decreases to 2.5% in the Darkes Forest Sandstone (Appendix
5-21, 5-22 and 8-19). Up-sequence, muscovite fluctuates between 0% and
5.1% from the Allans Creek Formation to the Unnamed Member Three
(Appendix 5-21, 5-22 and 8-19). It is recorded in the Unnamed Member Two at
average of 3.1% but then increases to 3.9% in the Loddon Sandstone (3.9%;
Appendix 5-21, 5-22 and 8-19). The proportion of muscovite in the Bargo
Claystone is the highest at an average of 5.6% (Appendix 5-21, 5-22 and 8-19)
011

whereas in the American Creek Coal Member muscovite is absent (Appendix 515). Muscovite occurs in shale, siltstone and sandstone at average
concentrations of 5.4%, 4.1% and 2%, respectively (Appendix 5-18).
5-3-3-5 Heavy minerals
Heavy minerals are observed in all samples in different proportions ranging
from 0.2 to 10.5% (Appendix 5-15 and 8-20). Heavy minerals fluctuate between
3.3% and 5.3% from the Wilton Formation to the Allans Creek Formation but
then decrease to 4.4% in the American Creek Coal Member (Appendix 5-21, 522 and 8-19). Up-sequence, they increase to 5.4% in the Unnamed Member
Three and then decrease to 3.3% in the Unnamed Member Two (Appendix 521, 5-22 and 8-19). From the Laurence Sandstone up-sequence to the Loddon
Sandstone, heavy minerals fluctuate between 3.6% and 4.4% (Appendix 5-21,
5-22 and 8-19).
The Bargo Claystone and Unnamed Member Two have the lowest percentage
of heavy minerals (3.3%) whereas the Bulli Coal has the highest percentage
(9.5%; Appendix 5-21, 5-22 and 8-19). Hornblende is the dominant heavy
mineral in the fine-grained samples.
5-3-3-6 Carbonate minerals
Variable amounts of carbonate minerals occur in the Illawarra Coal Measures.
They range between 0% and 64.8% and consist of siderite, calcite, dolomite
and ankerite (Appendix 5-15 and 8-20). Ankerite is dominant in the sandstone
samples (4.5%; Appendix 5-18) whereas siderite is more common in siltstone
(2.7%) and shale (1.8) samples (Appendix 5-18). Carbonate minerals are more
common in the Wilton Formation (18.1%) than in other units (Appendix 5-21, 522 and 8-19). This percentage decreases up-sequence to 2.9% in the Tongarra
Coal but then increases to 15.5% in the Darkes Forest Sandstone (Appendix 521, 5-22 and 8-19). Up-sequence, carbonate minerals fluctuated between 1.7%
and 7.6% from Allans Creek Formation to the Loddon Sandstone (Appendix 521, 5-22 and 8-19).

011

The Wilton Formation and Darkes Forest Sandstone include the highest
percentages of carbonate minerals at average of 18.1% and 15.5% (Appendix
5-21, 5-22 and 8-19). The lowest percentages are determined in the Wongawilli
Coal and Balgownie Coal at average of 1.7% and 1.8% (Appendix 5-21, 5-22
and 8-19). Carbonate minerals are more common in sandstone (8.8%) than in
siltstone (4.2%) and shale (5.5%; Appendix 5-18).
One sample of coal, one sample of carbonate cement and nine samples of tuff
were analysed in the sequence and are described following X-ray diffraction.
The results are shown in Appendix 5-23.

Narrabeen Group
5-3-1 Introduction
The petrography of the Narrabeen Group was based on one hundred and fifty
three samples. These samples were selected from outcrop and from seven
wells: EAW 30, EAW 42, EAW 156, EDEN 124, EDEN 125, EDEN 126 and
EDEN 127. These samples include ninety seven samples of sandstone,
eighteen samples of siltstone, thirty four samples of shale and three igneous
rocks. Thin sections, X-ray diffraction and scanning electron microscope were
used to analyse these samples.
Fine-grained sediment in the Narrabeen Group was previously studied by
several authors, such as Baker (1956), Loughnan et al. (1964, 1974); Goldberry
and Holland (1973), and Retallack (1977). Ward (1972) noted quartzose,
quartz-lithic and volcanic sandstones as the three sediment suites in the
Narrabeen Group. Provenance markers of foreland basin-fill sediments were
determined by Cowan (1993). The composition and texture of the Narrabeen
Group were described by McElroy (1954) in a systematic manner. Descriptions
of the petrology in the Narrabeen Group have been conducted by numerous
authors (e.g. Loughnan, 1963; Ward, 1971a; b; Bai, 1991; Dehghani, 1994; Bai
and Keene, 1996). These studies showed that quartz, rock fragments, feldspar
and clay minerals are recorded in the Narrabeen Group. Also, siderite, heavy
019

minerals, mica and iron oxide are present in this unit (Dehghani, 1994; Bai and
Keene, 1996).

5-3-2 Thin section


5-3-2-1 Detrital Composition and Texture
Point count size analysis was used to describe grain size and sorting in the
Narrabeen Group. The measured mean grain sizes range from siltstone (56 m)
to very coarse sandstone (1576 m; Appendix 5-8) whereas sorting is recorded
between 0.4 and 1.2 in this succession (Appendix 5-2). Litharenite is mostly
characterized by poor sorting and roundness (von Eynatten and Gaupp, 1999).
According to Maher et al. (2004), the mixing of grains with very different
transport histories can be indicated by the occurrence of rounded and angular
grains in one sample.
The Narrabeen Group is rich in rock fragments and is classified as litharenite to
sublitharenite, and rarely quartzarenite (Q50.3%, F1.4%, R48.3%; Fig. 5-21a, Table 51). The classification of units in the Narrabeen Group according to Folk (1968)
is shown in Figures 5-21b, 5-22a, b and 5-23a, b. The average percentages of
quartz, feldspar and rock fragments for units in the Narrabeen Group are
recorded in Table 5-1.
The Dickinson diagram (1985) shows that samples from the Narrabeen Group
plot in the lithic recycled, transitional recycled to quartzose recycled and rarely
craton interior provenance fields (Qm45.5%, F1.5%, Lt52.9%; Fig. 5-24a, Table 5-1).
The source of the units in the Narrabeen Group is shown in Figures 5-24b, 525a, b and 5-26a, b. Table 5-1 includes the average percentages of
monocrystalline quartz, feldspar and lithic fragments for units in the Narrabeen
Group. The classification of samples in the Narrabeen Group according to Folk
(1968) and Dickinson (1985) is recorded in Appendix 5-5.
5-3-2-1-1 Quartz
In the Narrabeen Group, the lowest percentage of detrital quartz (5.7%) was
observed in sample CCSS-Surface2 in the Coalcliff Sandstone (Appendix 5-8
and 8-2). The highest percentage of detrital quartz (85.5%) is present in sample
019

a)

Narrabeen Group

Quartz
arenite

Sublitharenite
Subarkose

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite

Quartz
arenite

b)

Coalcliff Sandstone
Sublitharenite

Subarkose

Arkose

Litharenite

Lithic
arkose

Feldspathic
litharenite

Figure 5-21: (a) Classification of the Narrabeen Group, (b)


Classification of the Coalcliff Sandstone (after Folk, 1968). Q=quartz,
F=feldspar R=rock fragment.

017

Quartz
arenite

a)

Wombarra Claystone
Sublitharenite

Subarkose

Arkose

Litharenite

Lithic
arkose

Feldspathic
litharenite

Quartz
arenite

b)

Scarborough Sandstone
Sublitharenite

Subarkose

Litharenite

Arkose

Lithic
arkose
F

Feldspathic
litharenite
R

Figure 5-22: (a) Classification of the Wombarra Claystone, (b)


Classification of the Scarborough Sandstone (after Folk, 1968).
Q=quartz, F=feldspar R=rock fragment.

011

a)

Bulgo Sandstone

Quartz
arenite

Stanwell Park Claystone


Sublitharenite

Subarkose

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite

b)

Newport Formation

Quartz
arenite

Bald Hill Claystone

Subarkose

Sublitharenite

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite
R

Figure 5-23: (a) Classification of the Stanwell Park Claystone and


Bulgo Sandstone, (b) Classification of the Bald Hill Claystone and
Newport Formation (after Folk, 1968). Q=quartz, F=feldspar
R=rock fragment.

010

Qm

a)
Craton
interior

Narrabeen Group
Quartzose
recycled

Transitional
continental

Mixed
Basement
uplift

Transitional
recycled

Dissected Ac
Lithic
continental
Lt

F
Undissected

Ac

Qm

b)
Craton
interior

Coalcliff Sandstone
Quartzose
recycled

Transitional
continental

Mixed
Basement
uplift

Transitional
recycled

Dissected Ac
Lithic
continental

F
Undissected

Ac

Lt

Figure 5-24: (a) The provenance of the Narrabeen Group, (b) The
provenance of the Coalcliff Sandstone (after Dickinson, 1985).
Qm=monocrystalline quartz, F=feldspar, Lt=rock fragment + chert.

011

Qm
a)
Craton
interior

Wombarra Claystone

Quartzose
recycled

Transitional
continental
Mixed
Basement
uplift

Transitional
recycled

Dissected Ac
Lithic
continental

Lt
Undissected Ac

Qm
b)

Craton
interior

Scarborough Sandstone
Quartzose
recycled

Transitional
continental

Mixed

Basement
uplift

Transitional
recycled

Dissected Ac
Lithic
continental

Lt
Undissected Ac

Figure 5-25: (a) The provenance of the Wombarra Claystone, (b) The
provenance of the Scarborough Sandstone (after Dickinson, 1985).
Qm=monocrystalline quartz, F=feldspar, Lt=rock fragment + chert.

011

Qm
a)

Bulgo Sandstone
Stanwell Park Claystone

Craton
interior

Quartzose
recycled
Transitional
continental
Mixed
Basement
uplift

Transitional
recycled

Dissected Ac
Lithic
continental

Lt
Undissected Ac

Qm

b)

Craton
interior

Newport Formation
Bald Hill Claystone
Quartzose
recycled

Transitional
continental
Mixed

Basement
uplift

Transitional
recycled

Dissected Ac
Lithic
continental

Lt
Undissected Ac

Figure 5-26: (a) The provenance of the Stanwell Park Claystone and
Bulgo Sandstone, (b) The provenance of the Bald Hill Claystone and
Newport Formation (after Dickinson, 1985). Qm=monocrystalline
quartz, F=feldspar, Lt=rock fragment + chert.

011

EAW 42(2) in the Newport Formation (Appendix 5-8 and 8-2). Quartz occurs in
the Coalcliff Sandstone at an average of 24.4%, increasing upward to an
average of 29.5% in the Scarborough Sandstone (Appendix 5-13, 5-14 and 81), due to variations in the amounts of detrital lithic fragments in these units.
Quartz percentages fluctuate between averages of 18.3% and 67.6% from the
Stanwell Park Claystone to the Newport Formation, which has the highest
average percentage (67.6%; Appendix 5-13, 5-14 and 8-1). In previous work,
Deen (1999) showed that quartz in the Newport Formation (62.25%) is higher
than in the Scarborough Sandstone (51.92%), the Coalcliff Sandstone (51.92%)
and the Bulgo Sandstone (48.37%).
In the current study, monocrystalline quartz grains increase from 19.7% in the
Coalcliff Sandstone to 23.7% in the Scarborough Sandstone and then fluctuate
between 8.2% and 63.8% from Stanwell Park Claystone to the Newport
Formation (Appendix 5-13, 5-14 and 8-1). Polycrystalline quartz grains increase
significantly from 4.7% in the Coalcliff Sandstone to 23.5% in the Bald Hill
Claystone, decreasing again to 3.8% in the Newport Formation (Appendix 5-13,
5-14 and 8-1). Monocrystalline quartz grains exceed polycrystalline quartz
grains in most samples (Appendix 5-13 and 5-14; Bai, 1991; Bai and Keene
1996). All samples include polycrystalline quartz, except sample EAW 30(9) in
the Wombarra Claystone (Appendix 5-8). The Bald Hill Claystone contains more
polycrystalline quartz (23.5%) than monocrystalline quartz (8.2%; Appendix 513, 5-14 and 8-1).
In this study, detrital quartz is present at an average of 32.6% and 19.5% in
sandstone and in siltstone, respectively (Appendix 5-11). Plutonic quartz is
observed in all samples whereas volcanic quartz is rare to absent in most
samples (Appendix 5-8).
There is no correlation between monocrystalline quartz and grain size and also
there is no correlation between polycrystalline quartz and grain size in the
Narrabeen Group. Whereas, a moderate negative correlation is recorded
between quartz and grain size in the Coalcliff Sandstone (r = 0.3; Fig. 5-27a). A
good negative correlation occurs between quartz and rock fragments in the
Narrabeen Group (r = 0.5; Fig. 5-27b). Quartz shows a good positive
011

correlation with height above base in the Narrabeen Group (r = 0.5; Fig. 5-28a).
A good negative correlation is present between quartz and rock fragments in
the Coalcliff Sandstone (r = 0.5; Fig. 5-28b) and this increases to a high
negative correlation with rock fragments in the Bulgo Sandstone (r = 0.8; Fig.
5-29).
a

1200

y = -17.788x + 826.03
R = 0.315

Grain size (micron)

1000
800
600
400
200
0
0

10

20

30
Quartz (%)

90

40

50

60

y = -0.7432x + 56.037
R = 0.4909

80

rock fragment (%)

70
60
50
40
30
20
10
0
0

20

40

60

80

100

Quartz(%)

Figure 5-27: (a) Quartz is moderately negatively correlated with


grain size in the Coalcliff Sandstone, (b) A good negative
correlation between quartz and rock fragments in the Narrabeen
Group.
011

a
y = 3.4535x + 49.758
R = 0.531

450.0

Height above base (m)

400.0
350.0
300.0
250.0
200.0
150.0
100.0
50.0
0.0
0

20

40
60
Quartz (%)

80

90

100

y = -1.3924x + 75.758
R = 0.4891

80

Rock fragment (%)

70
60
50
40
30
20
10
0
0

10

20

30
Quartz (%)

40

50

60

Figure 5-28: (a) A good positive correlation is observed between


quartz and height above base in the Narrabeen Group, (b) A good
negative correlation is present between quartz and rock fragments
in the Coalcliff Sandstone.

019

y = -0.4626x + 39.096
R = 0.7757

20
18
Rock fragment (%)

16
14
12
10
8
6
4
2
0
45

50

55

60
65
Quartz (%)

70

75

80

Figure 5-29: There is a high negative correlation between quartz


and rock fragments in the Bulgo Sandstone.

019

5-3-2-1-2 Feldspar
Bai (1991) and Bai and Keene (1996) noted that feldspar is rare and has a
percentage of less than 2% in the Narrabeen Group. The results from this study
also indicated that feldspar contents are similar to previous work with
percentages that vary between 0 and 2.9% (Appendix 5-8 and 8-4) at an
average of 0.9% in sandstone and 0.4% in siltstone (Appendix 5-11). Feldspar
grains gradually increase from 0.7% in the Coalcliff Sandstone to 2.1% in the
Stanwell Park Claystone (Appendix 5-13, 5-14 and 8-3). Feldspar grains exist in
the Bulgo Sandstone, Bald Hill Claystone and Newport Formation at averages
of 0.9%, 1% and 0.5% (Appendix 5-13, 5-14 and 8-3). The Stanwell Park
Claystone has the highest percentage of feldspar at an average of 2.1%
whereas the Newport Formation has the lowest percentage of feldspar at an
average of 0.5% (Appendix 5-13, 5-14 and 8-3).
Plagioclase grains increase from 0.4% in the Coalcliff Sandstone to 1% in the
Stanwell Park Claystone (Appendix 5-13, 5-14 and 8-3). Up-sequence, they are
0.4%, 0.6% and 0.3% in the Bulgo Sandstone, Bald Hill Claystone and the
Newport Formation, respectively (Appendix 5-13, 5-14 and 8-3). K-feldspar
grains tend to increase from 0.2% in the Coalcliff Sandstone to 1.1% in the
Stanwell Park Claystone, but decrease to 0.2% in the Newport Formation
(Appendix 5-13, 5-14 and 8-3). In general, plagioclase dominates over Kfeldspar in all units of the Narrabeen Group except the Scarborough Sandstone,
Stanwell Park Claystone and Bulgo Sandstone (Appendix 5-13, 5-14 and 8-3).
The results did not demonstrate the presence of any significant correlation
between feldspar types and grain size in the Narrabeen Group.
5-3-2-1-3 Lithic grains
Rock fragments and chert and are the dominant components in the Narrabeen
Group. Lithic grains comprise both volcanic and sedimentary rock fragments.
Sample Surface2 has the highest percentage of lithic grains at 84.4% in the
Coalcliff Sandstone, whereas lithic grains are rare in sample EAW 30(9) in the
Wombarra Claystone (0.5%; Appendix 5-8 and 8-6). Lithic grains decrease up
017

sequence from an average of 41.8% in the Coalcliff Sandstone to an average of


32.2% in the Scarborough Sandstone (Appendix 5-13, 5-14 and 8-5). Lithic
grains fluctuate between averages of 7.3% and 34.6% from the Stanwell Park
Claystone to the Newport Formation (Appendix 5-13, 5-14 and 8-5). Sandstone
and siltstone samples contain lithic grains at averages of 33% and 3.6% in the
Narrabeen Group (Appendix 5-11). A moderate positive correlation is indicated
between lithic grains and grain size in both the Coalcliff Sandstone (r = 0.4; Fig.
5-30) and the Scarborough Sandstone (r = 0.4; Fig. 5-31).

y = 10.038x - 27.397
R = 0.3977

1200

Grain size (micron)

1000
800
600
400
200
0
0

20

40

60

80

100

Lithic grains(%)

Figure 5-30: A moderate positive correlation is observed between


lithic grains and grain size in the Coalcliff Sandstone.

011

y = 12.652x + 43.645
R = 0.3675

1200

Grain size (micron)

1000
800
600
400
200
0
0

10

20

30

40

50

60

Lithic grains(%)

Figure 5-31: Lithic grains show moderate positive correlation with


grain size in the Scarborough Sandstone

5-3-2-1-3-1 Volcanic Rock Fragment plus chert and carbonate fragments


Bai (1991) and Bai and Keene (1996) showed that igneous rock fragments
constitute between 0.4 and 55.8%, and between 4 and 64% of the Narrabeen
Group, respectively. This study found that volcanic rock fragments plus chert
and carbonate-replaced fragments form an average of 29.8% in sandstone and
3.6% in siltstone (Appendix 5-11). Up-sequence, they decrease from 38.8% in
the Coalcliff Sandstone to 28.9% in the Scarborough Sandstone (Appendix 5-13
and 5-14). They fluctuate between 5.9% and 28.9% from the Stanwell Park
Claystone to the Newport Formation (Appendix 5-13 and 5-14). The highest
percentage of volcanic rock fragments plus chert and carbonate-replaced
fragments is recorded at the base of the Narrabeen Group in the Coalcliff
Sandstone (38.8%) whereas the lowest percentage is recorded at the top of the
Narrabeen Group in the Newport Formation (5.9%; Appendix 5-13 and 5-14).
5-3-2-1-3-2 Sedimentary Rock Fragments
Bai (1991) and Bai and Keene (1996) indicated that sedimentary rock fragments
form less than 2% of the Narrabeen Group. This study showed that sedimentary
010

rock fragments occur in the succession at average of 3.1% in sandstone


whereas they were not recorded in the siltstone (Appendix 5-11). A few
sandstone samples did not include sedimentary rock fragments in this study
(Appendix 5-8). The percentages of sedimentary rock fragments fluctuated
between 1% and 6% from the Coalcliff Sandstone to the Newport Formation
(Appendix 5-13 and 5-14).
5-3-2-1-3-3 Chert
Bai and Keene (1996) described chert as a sedimentary rock fragment, which
has a percentage of less than 4%. Bai (1991) showed that chert forms less than
1% in the upper Narrabeen Group, whereas in the lower Narrabeen Group it
forms more than 2%. In this study, chert is present as a major component
ranging from 1% to 82% in the Narrabeen Group and it is observed in all
studied samples (Appendix 5-8 and 8-6). Stratigraphic horizon plays an
important role in the determination of chert content. Chert is present in all
samples and decreases in amount up sequence from 34.7% in the Coalcliff
Sandstone to 2.9% in the Bulgo Sandstone (Appendix 5-13, 5-14 and 8-5). It is
identified in the Bald Hill Claystone and Newport Formation at averages of
15.2% and 1.4% (Appendix 5-13, 5-14 and 8-5). It is the main component in the
Coalcliff Sandstone (34.7%), which is the highest percentage in this unit
(Appendix 5-13, 5-14 and 8-5). The lowest percentage is present in the Newport
Formation at average of 1.4% (Appendix 5-13, 5-14 and 8-5). In the whole
section, it is less common in siltstone (2.8%) than in sandstone (24.5%;
Appendix 5-11).
There is a moderate positive correlation between chert and grain size in both
the Coalcliff Sandstone (r = 0.4; Fig. 5-32a) and the Scarborough Sandstone (r
= 0.4; Fig. 5-32b).
In general, the lithic fragments are more common than quartz grains in the
lower units of the Narrabeen Group (Bai and Keene, 1996).

011

5-3-2-1-4 Muscovite
Up to 10% muscovite is recorded in sample EAW 30(21) in the Wombarra
Claystone (Appendix 5-8). Siltstone samples (5.3%) have a higher percentage
of muscovite than sandstone samples (0.4%; Appendix 5-11). Muscovite occurs
in the Coalcliff Sandstone at an average of 0.5% (Appendix 5-13, 5-14 and 8-4).
It increases to 0.7% in the Wombarra Claystone but then decreases up
sequence disappearing in the Bulgo Sandstone (Appendix 5-13, 5-14 and 8-4).
In the Gosford Subgroup, chert occurs at average of 0.4% and 0.9% in the Bald
Hill Claystone and Newport Formation, respectively (Appendix 5-13, 5-14 and 84).
a

y = 9.4783x + 63.014
R = 0.3528

1200

Grain size (micron)

1000
800
600
400
200
0
0

20

40

60

80

100

Chert (%)

y = 13.051x + 177.81
R = 0.3909

1200

Grain size (micron)

1000
800
600
400
200
0
0

10

20

30

40

50

Chert (%)

Figure 5-32: A moderate positive correlation occurs between chert


and grain size in the Coalcliff Sandstone and Scarborough
Sandstone.
011

5-3-2-1-5 Heavy minerals


Heavy minerals are found in trace amounts in the succession, making up
between 0.3% and 3.3% (Appendix 5-8 and 8-8). They occur at average of
1.2% and 0.8% in the sandstone and siltstone, respectively (Appendix 5-11).
They increase up sequence from 1% in the Coalcliff Sandstone to 1.5% in the
Stanwell Park Claystone but then decrease to 1.3% in the Bulgo Sandstone
(Appendix 5-13, 5-14 and 8-7). They exist at averages of 2.2% and 1.6% in
Bald Hill Claystone and Newport Formation, respectively (Appendix 5-13, 5-14
and 8-7). The Bald Hill Claystone is characterised by the highest percentage
(2.2%) whereas the Coalcliff Sandstone is characterised by the lowest
percentage (1%; Appendix 5-13, 5-14 and 8-7).
Dissolution may have reduced the percentage of heavy minerals in the
Narrabeen Group. Hornblende dominates the heavy minerals in all units of the
Narrabeen Group (Appendix 5-13 and 5-14). A moderate positive correlation is
present in Figure 5-33 between heavy minerals and height above base in the
Coalcliff Sandstone (r = 0.4).
y = 5.4468x + 98.707
R = 0.3961

120

Height above base (m)

115
110
105
100
95
90
0

0.5

1.5

2.5

3.5

Heavy minerals (%)

Figure 5-33: There is a moderate positive correlation between


heavy minerals and height above base in the Coalcliff Sandstone.

011

5-3-2-1-6 Chlorite
Chlorite is uncommon in most samples and varies between 0% and 0.6%
(Appendix 5-8 and 8-8). It is constant in the Coalcliff Sandstone, Wombarra
Claystone and Scarborough Sandstone at average of 0.1% but then increases
to 0.2% in the Stanwell Park Claystone (Appendix 5-13, 5-14 and 8-7). Upsequence, it occurs in the Bulgo Sandstone, Bald Hill Claystone and Newport
Formation at averages of 0.03%, 0.6% and 0.03%, respectively (Appendix 5-13,
5-14 and 8-7). The Bald Hill Claystone includes the highest percentage 0.6%
(Appendix 5-13, 5-14 and 8-7).
5-3-2-1-7 Matrix
The Narrabeen Group is rich in matrix which varies between 0% to 59%
(Appendix 5-8 and 8-12) at an average of 13.9% in sandstone and 44.2% in
siltstone (Appendix 5-11). The percentages of matrix fluctuate between 5% and
17.4% from the Coalcliff Sandstone to the Newport Formation (Appendix 5-13,
5-14 and 8-11). The highest percentage and the lowest percentage are
determined in the Wombarra Claystone and the Bulgo Sandstone at averages
of 17.4% and 5%, respectively (Appendix 5-13, 5-14 and 8-11). Matrix shows a
good negative correlation with grain size in both the Coalcliff Sandstone and the
Scarborough Sandstone (r = 0.5; r = 0.5; Figs 5-34 and 5-35).
y = -24.412x + 759.02
R = 0.5382

1200

Grain size (micron)

1000
800
600
400
200
0
0

10

20

30

40

50

Matrix (%)

Figure 5-34: There is a good negative correlation between matrix


and grain size in the Coalcliff Sandstone.
011

1200

y = -17.209x + 713.06
R = 0.5241

Grain size (micron)

1000
800
600
400
200
0
0

10

20

30

40

50

Matrix (%)

Figure 5-35: There is a good negative correlation between matrix


and grain size in the Scarborough Sandstone.

5-3-3 X-ray diffraction (XRD)


X-ray diffraction was used to study fine-grained samples including fine- to
medium-grained sandstone, siltstone, shale, mudstone and claystone. Finegrained lithologies are dominant in the Wombarra Claystone, Stanwell Park
Claystone, Bald Hill Claystone, Garie Formation and Newport Formation.
5-3-3-1 Quartz
In the fine-grained samples, X-ray diffraction showed that the percentages of
quartz in the Narrabeen Group was not less than 0.4%, but did not exceed
83.2% (Appendix 5-16 and 8-16). Quartz occurs at an average of 59.2% in
sandstone whereas it gradually decreases in siltstone and shale samples at
averages of 36% and 21.9% (Appendix 5-19). The Coalcliff Sandstone (48.4%),
Wombarra Claystone (44.7%), Scarborough Sandstone (44.3%) and Stanwell
Park Claystone (45.5%) have similar percentages of quartz (Appendix 5-21, 522 and 8-15). Up-sequence, percentages of quartz increase to 57.1% in the
Bulgo Sandstone as the highest percentage but then decreases to 0.6% in the
Garie Formation as the lowest percentage (Appendix 5-21, 5-22 and 8-15). In

011

the Newport Formation, quartz increases to 54.7% (Appendix 5-21, 5-22 and 815).
In the Narrabeen Group, a low negative correlation is recorded between quartz
and feldspar (r = 0.2; Fig. 5-36). Also, a strong negative correlation is present
between quartz and total clay minerals (r = 0.8; Fig. 5-37a) whereas a
moderate negative correlation exists between quartz and kaolinite in the
Narrabeen Group (r = 0.4; Fig. 5-37b). There is a moderate negative
correlation between quartz and clay minerals in the Coalcliff Sandstone (r =
0.4; Fig. 5-38a). Quartz is highly negatively correlated with clay minerals in both
the Scarborough Sandstone (r = 0.9; Fig. 5-38b) and Bulgo Sandstone (r =
0.9; Fig. 5-39a) and it has a good negative correlation with feldspar in the Bulgo
Sandstone (r = 0.5; Fig. 5-39b). A high negative correlation is also recorded
between quartz and kaolinite in the Bulgo Sandstone (r = 0.8; Fig. 5-40).

25

y = -0.0587x + 7.8762
R = 0.1768

Feldspar (%)

20

15

10

0
0

20

40

60

80

100

Quartz (%)

Figure 5-36: Quartz is low negative correlation with feldspar in the


Narrabeen Group (XRD-results).

019

a
y = -0.5822x + 57.351
R = 0.7916

80
70
Clay minerals (%)

60
50
40
30
20
10
0
0

20

40

60

80

100

Quartz (%)

b
70

y = -0.4305x + 29.545
R = 0.4278

60

Kaolinite (%)

50
40
30
20
10
0
0

20

40

60

80

100

Quartz (%)

Figure 5-37: (a) Quartz shows high negative correlation with clay
minerals in the Narrabeen Group (XRD-results), (b) A moderate
negative correlation exists between quartz and kaolinite in the
Narrabeen Group (XRD-results)

019

a
45

y = -0.3709x + 47.547
R = 0.4218

40

Clay minerals (%)

35
30
25
20
15
10
5
0
0

20

40
Quartz (%)

60

80

b
y = -0.6872x + 61.796
R = 0.9493

50
45

Clay minerals (%)

40
35
30
25
20
15
10
5
0
0

10

20

30

40

50

60

70

Quartz (%)

Figure 5-38: (a) There is moderate negative correlation between


quartz and clay minerals in the Coalcliff Sandstone (XRD-results),
(b) Quartz is highly negatively correlated with clay minerals in the
Scarborough Sandstone (XRD-results).

017

a
50

y = -0.4209x + 44.996
R = 0.9295

45

Clay minerals (%)

40
35
30
25
20
15
10
5
0
0

20

40

60

80

100

Quartz (%)

b
5

y = -0.0265x + 4.4211
R = 0.4619

4.5
4

Feldspar (%)

3.5
3
2.5
2
1.5
1
0.5
0
0

20

40

60

80

100

Quartz (%)

Figure 5-39: (a) Quartz is highly negatively correlated with clay


minerals in the Bulgo Sandstone (XRD-results), (b) Good negative
correlation is present between quartz and feldspar in the Bulgo
Sandstone (XRD results).

011

16
y = -0.1483x + 14.045
R = 0.8463

14

Kaolinite (%)

12
10
8
6
4
2
0
0

20

40

60

80

100

Quartz (%)

Figure 5-40: A high negative correlation is present between


quartz and kaolinite in the Bulgo Sandstone (XRD-results).

010

5-3-3-2 Feldspar
In this study, feldspar in the Narrabeen Group occurs at an average of 4.9% in
sandstone, 4.7% in siltstone and 6.3% in shale (Appendix 5-19). It varies
between 0.9% and 23.8% (Appendix 5-16 and 8-18). In the XRD samples, the
Coalcliff Sandstone has an average feldspar content of 5.9% that decreases up
sequence to 2.9% in the Bulgo Sandstone, and then increases to 6.8% in the
Bald Hill Claystone (Appendix 5-21, 5-22 and 8-17). Up-sequence, feldspar
decreases to 2.2% in the Newport Formation (Appendix 5-21, 5-22 and 8-17).
Plagioclase content decreases from 3.7% in the Coalcliff Sandstone (as the
highest percentage) to 2.2% in the Scarborough Sandstone (Appendix 5-21, 522 and 8-17). In the Stanwell Park Claystone and the Bulgo Sandstone, it is
observed at 2.4% and 1% but then increases to 3.5% in the Bald Hill Claystone
(Appendix 5-21, 5-22 and 8-17). Up-sequence, it decreases steadily to 0.9% in
the Newport Formation (Appendix 5-21, 5-22 and 8-17). K-feldspar content
occurs in the Coalcliff Sandstone at 2.2% and then increases to 3.3% in the
Scarborough Sandstone. Up-sequence it decreases to 1.9% in the Bulgo
Sandstone and then increases to 3.3% in the Bald Hill Claystone (Appendix 521, 5-22 and 8-17). Up-sequence, it decreases to 1.3% in the Newport
Formation which is the lowest percentage (Appendix 5-21, 5-22 and 8-17).
5-3-3-3 Clay minerals
The percentages of clay minerals contained in the fine-grained samples ranges
between 6.4% and 71.7% (Appendix 5-16 and 8-16). Clay minerals are present
at averages of 22.9%, 35.1% and 45.3% in sandstone, siltstone and shale,
respectively (Appendix 5-20). Clay minerals in the Scarborough Sandstone
(31.4%) are more common than in the underlying Wombarra Claystone (30.4%)
and Coalcliff Sandstone (29.6%) but then decrease to 20.9% in the Bulgo
Sandstone as the lowest percentage (Appendix 5-21, 5-22 and 8-15). Upsequence, they increase to 69.8% in the Garie Formation as the highest
percentage (Appendix 5-21, 5-22 and 8-15). In the Newport Formation, clay
minerals are determined at an average of 26.7% (Appendix 5-21, 5-22 and 815).
011

Kaolinite is the most widespread clay mineral in shale samples whereas


sandstone and siltstone samples are characterized by mixed-layer illite/smectite
(Appendix 5-19). Clay minerals including illite, kaolinite, dickite, chlorite and
mixed-layer illite/smectite are reported in all units of the Narrabeen Group
(Appendix 5-21 and 5-22).
5-3-3-4 Muscovite/illite
In the fine-grained samples, muscovite does not exceed 44.1% (Appendix 5-16
and 8-20) and occurs at an average of 1.4% in sandstone, 13.8% in siltstone,
and 11.3% in shale (Appendix 5-19). Muscovite decreases from 6.7% in the
Coalcliff Sandstone to 5.2% in the Stanwell Park Claystone (Appendix 5-21, 522 and 8-19). Up-sequence, it increases to its highest percentage (11.2%) in
the Bulgo Sandstone but then decreases to its lowest percentage (2.5%) in the
Garie Formation (Appendix 5-21, 5-22 and 8-19). Muscovite is observed in the
Newport Formation at average of 7.8% (Appendix 5-21, 5-22 and 8-19).
5-3-3-5 Heavy minerals
Minor amounts (1.3%-32.3%; Appendix 5-16 and 8-20) of heavy minerals are
reported in the fine-grained samples from the Narrabeen Group. They are found
at average of 5% in sandstone, 4.1% in siltstone, and 9.3% in shale (Appendix
5-19). Six types of heavy minerals are identified: hematite, hornblende, rutile,
zircon, tourmaline and pyrite.
The heavy minerals are less common in the Coalcliff Sandstone at an average
of 3.6%, but then increase up sequence to 7.3% in the Stanwell Park Claystone
(Appendix 5-21, 5-22 and 8-19). Up-sequence, percentages of heavy minerals
fluctuate between 4.2% and 19.1% from the Bulgo Sandstone to the Newport
Formation (Appendix 5-21, 5-22 and 8-19).
5-3-3-6 Carbonate Minerals
Carbonate minerals occur in percentages varying between 0.3% and 43.6% in
the fine-grained samples at an average of 6.6% in sandstone, 6.2% in siltstone
and 5.9% in shale (Appendix 5-16, 5-19 and 8-20). They consist of siderite,

011

ankerite, calcite and dolomite. Siderite is the main carbonate minerals in the
fine-grained samples.
The average content of carbonate minerals in the fine-grained samples from the
Coalcliff Sandstone (5.8%), Wombarra Claystone (7.6%), Scarborough
Sandstone (6.5%) and the Stanwell Park Claystone (8.3%) is somewhat similar
but it then decreases to 3.3% in the Bulgo Sandstone (Appendix 5-21, 5-22 and
8-19). It increases up-sequence to a maximum of 11.8% in the Garie Formation
(Appendix 5-21, 5-22 and 8-19). At the top of the Narrabeen Group, 4.5%
carbonate occurs in the Newport Formation (Appendix 5-21, 5-22 and 8-19).
Three samples of altered basalt were analysed by X-ray diffraction, including
4.7% quartz, 44.7% feldspar, 20.5% clay minerals, 13.5%illite, 0.5% hematite,
0.6% pyrite, 0.1% hornblende, 2.1% rutile and 13.2 carbonate (Appendix 5-19).

Hawkesbury Sandstone
5-3-1 Introduction
Thirty seven samples of the Hawkesbury Sandstone were chosen from outcrops
and two wells (EAW 18 a, and EDEN 115). These samples comprised twenty
nine samples of sandstone, four samples of siltstone, three samples of shale
and one sample of siderite. Thirty two samples were examined under a
polarizing microscope. Twenty two of the samples, consist of fine-grained
sandstone, siltstone, shale and siderite, were analysed by X-ray diffraction
(XRD). Eight samples were studied by scanning electron microscope (SEM).
Previous research has documented the petrology of the Hawkesbury
Sandstone, such as Standard (1961, 1969), Griffith (1986), Johnson (2006) and
Gentz (2006). Standard (1969) concluded that the composition of the
Hawkesbury Sandstone consisted of detrital grains, heavy minerals and clay
minerals. Furthermore, he described the grain size, sorting and roundness of
the sandstone. He noted that most grains are medium to coarse in the
sandstone, and are moderately to poorly sorted. Also, they occur as subangular to sub-rounded grains (Standard, 1969).
011

5-3-2 Thin Section


5-3-2-1 Detrital Composition and Texture
In this study, grain size, sorting and roundness of sandstone and siltstone
samples were interpreted from thin section, using point count grain size analysis
with a polarizing microscope. The results indicated that grain size varied from
siltstone (52 m) to coarse-grained sand (585 m), with an average size of
359 m (Appendix 5-9, 8-13 and 8-14). The sorting varies between very well
sorted (0.4) and poorly sorted (1.3; Appendix 5-3).
Petrographically, the triangular QFL and QmFLt diagrams were used to
determine the rock type and provenance of this unit. Petrographic data
indicated that the sandstones are quartzarenite to sublitharenite containing
abundant quartz, low feldspar and rock fragments (Q90.1%, F0.3%, R9.7%; Fig. 541). The QmFLt diagram from Dickinson (1985) indicates that the Hawkesbury
Sandstone falls into the craton interior to quartzose recycled provenance
classes (Qm89.7%, F0.3%, Lt10%; Fig. 5-42). The classification data for samples in
the Hawkesbury Sandstone according to Folk (1968) and Dickinson (1985) is
recorded in Appendix 5-6.
Q

Quartz
arenite

Hawkesbury Sandstone
Sublitharenite

Subarkose

Litharenite

Arkose

Lithic
arkose

Feldspathic
litharenite

Figure 5-41: Classification of the Hawkesbury Sandstone (after


Folk, 1968. Q=quartz, F=feldspar, R=rock fragments.
011

Qm
Hawkesbury Sandstone
Craton
interior
Quartzose
recycled
Transitional
continental
Mixed

Transitional
recycled

Dissected Ac
Basement
uplift

Lithic
continental

Lt

F
Undissected Ac

Figure 5-42: The provenance of the Hawkesbury Sandstone (after


Dickinson, 1985). Qm = monocrystalline quartz, F = feldspar, Lt =
rock fragment + chert.

5-3-2-1-1 Quartz
In examined samples, the Hawkesbury Sandstone is a quartz-rich sandstone.
Quartz grain abundance varies between 24.4% and 83.3% with an average of
64.5% in the sandstone, and 31.5% in the siltstone (Appendix 5-9, 5-12 and 82). These results are not very different from the results obtained by a number of
authors (Standard, 1964; Bayliss et al., 1965; Branagan, 2000; Franklin, 2000;
Dragovich, 2000; Freed, 2005; Johnson, 2006). Branagan (2000) showed that
quartz is widespread in the Hawkesbury Sandstone and varies between 60 and
70%. Franklin (2000) noted that unweathered sandstone has 78% quartz
whereas weathered sandstone contains 64% quartz. Gentz (2006) interpreted
that fractured quartz grains constitute 64.4% and are more abundant than nonfractured quartz grains. Also, Johnson (2006) indicated that simple unfractured
quartz comprises of 42.4%, while unfractured composite quartz constitutes
5.9%.

011

Quartz percentages are similar in all samples, except for two siltstone samples
where quartz content is lower (Appendix 5-9). The average abundance of
monocrystalline quartz grains is 61.6% in sandstone and 30.9% in siltstone
(Appendix 5-12). Monocrystalline quartz is more prevalent than polycrystalline
quartz grains, which contain an average composition of 2.8% in sandstone and
0.6% in siltstone (Appendix 5-12). The percentages of plutonic quartz are higher
than volcanic quartz in both sandstone and siltstones samples (Appendix 5-12).
The results indicated no correlation between total quartz and grain size and,
also no correlation between monocrystalline quartz or polycrystalline quartz and
grain size in the Hawkesbury Sandstone.
5-3-2-1-2 Feldspar
In the Hawkesbury Sandstones, feldspar grains are mainly rare. Its abundance
ranges from 0 to 0.5% (Appendix 5-9 and 8-4) at an average of 0.1% in
sandstone and 0.4% in siltstone (Appendix 5-12). K-feldspar is slightly more
common than plagioclase but most samples do not include feldspars (Appendix
5-9).
The correlation between feldspar and grain size is very low in the Hawkesbury
Sandstone.
5-3-2-1-3 Rock Fragments
Griffith (1986) showed that the massive sands contain more sedimentary rock
fragments than in the stratified channel sands.
In this study, rock fragments consist entirely of sedimentary rock fragments with
a complete absence of volcanic and metamorphic rock fragments in both the
sandstone and siltstone samples (Appendix 5-9). Sedimentary rock fragments
are present at an average abundance of 1.5% in sandstone and 1.1% in
siltstone (Appendix 5-12).

019

5-3-2-1-4 Chert
Chert is clear and observed in most samples and varies between 0 and 17.3%
(Appendix 5-9 and 8-6), with an average of 5.4% in sandstone and 2.7% in
siltstone (Appendix 5-12). Chert is absent in two samples (Appendix 5-9).
5-3-2-1-5 Muscovite
Based on point count results, muscovite exists in the Hawkesbury Sandstone at
an average of 0.03% in sandstone and 2.4% in siltstone (Appendix 5-12). The
regionally metamorphosed rocks and granite are the source of mica. Muscovite
is moderately negatively correlated with grain size in the Hawkesbury
Sandstone (r = 0.4; Fig. 5-43).

y = -132.21x + 392.81
R = 0.3821

700

Grain size (micron)

600
500
400
300
200
100
0
0

0.5

1.5

2.5

3.5

Muscovite (%)

Figure 5-43: Muscovite is moderately negatively correlated with


grain size in the Hawkesbury Sandstone.

5-3-2-1-6 Heavy minerals


McElroy (1954; 1957; 1958; 1961) studied heavy minerals in the Sydney Basin,
indicating that rutile is more common in the Hawkesbury Sandstone than in the
Narrabeen Group. In this study, the occurrence of heavy minerals was
recognized in sandstone and siltstone in minor amounts (Appendix 5-9), at an
019

average of 0.5% in sandstone and 0.4% in siltstone (Appendix 5-12). Heavy


minerals were noticeable in all samples, except three sandstone samples
(Appendix 5-9). Tourmaline is the most common heavy mineral in both the
sandstone and siltstone at an average of 0.3% (Appendix 5-12).
Pre-existing sedimentary rocks are the source of the rounded rutile grains. Also,
the rounded zircon grains are derived from earlier sedimentary rocks (Standard,
1964). Galloway (1972) showed that the variations in provenance have resulted
in differences between the heavy minerals in term of frequency, size and
morphology in the Hawkesbury Sandstone and Narrabeen Group.
5-3-2-1-7 Chlorite
Chlorite occurs in sandstone samples at average of 0.05% but disappears in
siltstone samples (Appendix 5-12).
5-3-2-1-8 Matrix
Matrix is mainly observed in the Hawkesbury Sandstone at an average
abundance of 8.6% in sandstone and 44.8% in siltstone (Appendix 5-12). The
fine-grained samples contain more matrix than the coarse-grained samples.
There is a good negative correlation between matrix and grain size in the
Hawkesbury Sandstone (r = 0.5; Fig. 5-44).
y = -8.6546x + 463.42
R = 0.5017

700

Grain size (micron)

600
500
400
300
200
100
0
0

10

20

30
Matrix (%)

40

50

60

Figure 5-44: There is a good negative correlation between matrix


and grain size in the Hawkesbury Sandstone.
017

5-3-3 X-ray diffraction analysis (XRD)


X-ray diffraction was used to determine the mineralogical composition and clay
mineral abundances of samples including fine- to medium-grained sandstone,
fine-grained sandstone, siltstone, shale and siderite.
5-3-3-1 Quartz
X-ray diffraction confirmed that the fine-grained samples in the Hawkesbury
Sandstone are characterized by quartz with percentages ranging from 27.8% to
78% (Appendix 5-17 and 8-16) with an average of 66.3% in sandstones, 36.3%
in siltstones and 36.7% in shale (Appendix 5-20). Thus, whereas sandstone
includes more quartz than siltstone and shale, the quartz content in shale and
siltstone is equivalent. A good negative correlation is present between quartz
and feldspar in the Hawkesbury Sandstone (r = 0.5; Fig. 5-45). Also, quartz
shows high negative correlation with clay minerals in this unit (r = 0.7; Fig. 546).

y = -0.0905x + 6.977
R = 0.4682

9
8

Feldspar (%)

7
6
5
4
3
2
1
0
0

20

40

60

80

100

Quartz (%)

Figure 5-45: Good negative correlation is present between quartz


and feldspar in the Hawkesbury Sandstone (XRD results).

091

60

y = -0.4655x + 59.082
R = 0.6688

Clay minerals (%)

50
40
30
20
10
0
0

20

40

60

80

100

Quartz (%)

Figure 5-46: Quartz is highly negatively correlated with clay


minerals in the Hawkesbury Sandstone (XRD-results).
5-3-3-2 Feldspar
Feldspar is present at an average of 0.7% in sandstones, whereas it is higher in
siltstone (3.1%) and shale (5.8%; Appendix 5-20). Albite and labradorite
occurred in most samples in trace amounts but were missing in other samples
(Appendix 5-17). Orthoclase is completely absent in sandstone and shale
samples, whereas it is observed in the one sample of siltstone (Appendix 5-17).
Microcline is present in all shale and siltstone samples whereas it is rare to
absent in sandstone samples (Appendix 5-17).
5-3-3-3 Clay minerals
In previous work, Gentz (2006) interpreted that sandstone contains between 5
and 15% kaolinite. A mixed-layer clay is evidenced in 14.7% of samples (Gentz,
2006). The maximum percentage of illite is 4.7% in sandstone (Gentz, 2006).
Mixed-layer clays are most prevalent in the claystones (Hutton et al., 2006;
Gentz, 2006). Kaolinite and illite are recorded in the claystones with an average
of 14.6% and 7.6%, respectively (Gentz, 2006).
090

In this study, the description of the clay minerals in the succession was largely
dependent on X-ray diffraction analysis. Clay minerals are significant
components in the sandstone (29.1%), siltstone (44.8%) and shale (34.6%;
Appendix 5-20). The siltstone contains clay minerals in greater amounts than in
the sandstone and shale. Kaolinite is the most prevalent clay mineral in
sandstone (20.3%), siltstone (24.6%) and shale (13.5%; Appendix 5-20). Other
clay minerals occurring in the sandstone, siltstone and shale are dickite, illite
and mixed-layer illite-smectite. The illite was derived from phyllites and slates in
the source area (Standard, 1964). Chlorite is present in few samples (Appendix
5-17). X-ray diffraction indicated that chlorite occurs in siltstone samples at an
average 0.7% (Appendix 5-20).
5-3-3-4 Muscovite/illite
Muscovite is absent in most fine-grained sandstone samples but it is present in
all siltstone and shale samples where it ranges from 3.4% to 21% (Appendix 517 and 8-20), with an average of 7.8% in siltstone and 15.1% in shale
(Appendix 5-20).
5-3-3-5 Heavy minerals
Heavy minerals varied between 0 and 9.5% (Appendix 5-17 and 8-20), with an
average of 1.8% in sandstone, 3.8% in siltstone and 6.7% in shale (Appendix 520). Tourmaline, hematite, pyrite, zircon and hornblende are heavy minerals
identified in sandstone, siltstone and shale, while rutile is only present in the
sandstone (Appendix 5-20). Also, X-ray diffraction results confirmed that
tourmaline and hornblende are the most widespread heavy minerals in the
siltstone and shale (Appendix 5-20).
5-3-3-6 Carbonate minerals
X-ray diffraction analysis indicated that carbonate minerals are represented by
siderite, calcite, dolomite and ankerite in various amounts (0-12.3%; Appendix
5-17 and 8-20). They exist with an average of 2.2% in sandstone, 4.3% in
siltstone and 1.1% in shale (Appendix 5-20). Siderite occurs as the most
091

common carbonate mineral at an average of 1.2% in sandstone, 3.6% in


siltstone and 0.5% in shale (Appendix 5-20).
A siderite sample was chosen from the outcrop succession, and contains a high
percentage of siderite (65.5%) with other components including quartz (6.2%),
feldspar (4%), clay minerals (11.3%), heavy minerals (3.4%), and muscovite
(8.1%; Appendix 5-20).

091

091

Chapter six
Diagenesis in the southern Sydney Basin
6-1 Introduction
In this chapter, diagenetic alteration and authigenic minerals in the southern
Sydney Basin are recorded. Also, the relationship between detrital composition
and diagenesis is discussed in this part. The diagenesis was described from
thin section and Scanning Electron Microscope analysis.
Diagenesis of the Eckersley Formation was previously studied by Cusack
(1991) who provided a diagenetic sequence. Descriptions of diagenesis in the
Narrabeen Group have been determined by Bai (1991) and Bai and Keene
(1996). Standard (1964), Griffith (1986) and Spry (2000) introduced a study of
diagenesis in the Hawkesbury Sandstone.

6-2 Illawarra Coal Measures


6-2-1 Scanning Electron Microscope (SEM)
Porosity, permeability and diagenesis of the Illawarra Coal Measures were
interpreted using Scanning Electron Microscope images. This technique
indicated the presence of clays as grain coatings (Appendix 7-1b, 7-2f and 74a) and as pore-filling cements (Appendix 7-1c, d, 7-2d and 7-3f) in all units of
the Illawarra Coal Measures. SEM showed clay mineral types including
kaolinite, illite, mixed-layer illite-smectite and chlorite (Appendix 7-1a-d, 7-2c-f,
7-3a-f and 7-4a-f). These minerals are described below.
Additionally, very tight packing of quartz and albitized feldspar grains was
shown by SEM in this sequence (Appendix 7-1e-f). However, SEM also
displayed quartz overgrowths and carbonate cementation in the Illawarra Coal
Measures (Appendix 7-1a, 7-2a-c, e and 7-3a, f). In addition, small and large
pores were determined by SEM (Appendix 7-1a-d, f, 7-2b-c, f, 7-3c-d, f, 7-4a-b
175

and d). Primary and secondary pores are described in chapter seven (Appendix
7-1a-d, f, 7-2b, c, 7-3c-d, f, 7-4a, b and d).

6-2-2 Diagenesis
Thin section and SEM analysis were used to describe the diagenesis of all units
in the Illawarra Coal Measures. Authigenic minerals comprise quartz
overgrowths, authigenic clay minerals, carbonate cement, authigenic feldspar,
authigenic pyrite and iron oxide cement. Also, the types of dissolution recorded
in the Illawarra Coal Measures include silica dissolution, dissolution of unstable
grains and dissolution of carbonate cement.

6-2-2-1 Diagenetic Minerals


6-2-2-1-1 Quartz cementation
Quartz cement is seen using both thin section and SEM techniques (Appendix
6-1a, 6-11a, 7-1a and 7-2a). It is uncommon in the Illawarra Coal Measures
which includes more detrital lithic grains than detrital quartz. Thus, quartz
cement is volumetrically negligible. Quartz cement percentages range from 0 to
2% (Appendix 5-7). Average quartz cement content fluctuates between 0.2%
and 0.4% from the Wilton Formation to the Wongawilli Coal (Appendix 5-13, 514 and 8-5), then it increases steadily up sequence from 0.3% in the Unnamed
Member Three to 0.6% in the Loddon Sandstone, which contains the highest
percentages (Appendix 5-13, 5-14 and 8-5).
Quartz cement exists as syntaxial overgrowths around detrital quartz grains or
is arranged as crystals oriented perpendicular to the quartz grains (Appendix 611a and 6-19b). It also occurs as a pore-filling cement (Appendix 7-1a, 7-2a and
7-3a) in most units of the Illawarra Coal Measures. Quartz overgrowths
decrease in abundance in sandstones that include high percentages of
carbonate cement (Umar et al., 2011). The presence of authigenic clays also
affects the overgrowth habit of authigenic quartz (Rossi et al., 2002; Umar et al.,
2011). Thus, precipitation of quartz overgrowths is difficult in samples with
greater contents of clay minerals. This means that grain-coating clays are
important in preventing the development of quartz overgrowths, and in some
samples they enhance reservoir quality. Double quartz overgrowths were
176

detected in some samples (Appendix 6-11b), whereas fluid inclusions occur


rarely in the succession. Fluid inclusions, clay coatings and iron oxides may be
observed separating quartz cement from detrital quartz grains (Appendix 6-11a
and 6-15a) whereas in some samples, the distinction between quartz
overgrowths and detrital grains is not clear.
A good positive correlation occurs between detrital quartz and quartz
overgrowths in the Laurence Sandstone and the Kembla Sandstone (r = 0.6
and 0.7; Figs 6-1 and 6-2a). Thus, quartz overgrowths are related to both
detrital quartz grains and porosity in these units. This confirms that quartz
overgrowths are generally more common in sandstone samples rich in detrital
quartz (Lee and Lim, 2008). In the Kembla Sandstone, quartz overgrowths have
a moderate negative correlation with grain size since the coarser samples
contain a higher proportion of lithic grains (r = 0.5; Fig. 6-2b).

y = 0.022x + 0.092
R = 0.578

1.4

Quartz overgrowth (%)

1.2
1
0.8
0.6
0.4
0.2
0
0

10

20

30
Quartz (%)

40

50

60

Figure 6-1: A good positive correlation is present between


detrital quartz and quartz overgrowths in the Lawrence
Sandstone.

177

y = 0.034x - 0.044
R = 0.666

0.9

Quartz overgrowth (%)

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0

10

15

20

25

Quartz (%)

y = -572.3x + 561.5
R = 0.535

900.0
800.0

Grain size (micron)

700.0
600.0
500.0
400.0
300.0
200.0
100.0
0.0
0

0.2

0.4
0.6
Quartz overgrowth (%)

0.8

Figure 6-2: a) Strong positive correlation between quartz and


quartz overgrowths in the Kembla Sandstone, b) Quartz
overgrowths are moderately negatively correlated with grain size
in the Kembla Sandstone.

178

6-2-2-1-2 Authigenic clay minerals


Authigenic clay minerals are described from SEM data and include kaolinite,
illite, mixed-layer illite/smectite and chlorite. Authigenic clay minerals represent
the second most abundant cement in the Illawarra Coal Measures and are
common as a pore-filling cement (Appendix 7-1c, d, 7-2d and 7-3f) and as
grain-coatings around detrital grains (Appendix 7-1b, 7-2f and 7-4a). Most
samples include two or more types of authigenic clay minerals.
Kaolinite (0.2-68.7%) is distributed in all units in the Illawarra Coal Measures. It
fills pores between grains, including primary and secondary pore spaces, and
exists as booklets and vermicular aggregates (Appendix 7-2d and 7-3b-c).
Kaolinite is coated by chlorite in some samples (Appendix 7-3b) and is
intergrown with some detrital clay (Appendix 7-3c). Illite is also associated with
kaolinite in some samples (e.g. Appendix 7-3c). Kaolinite is mostly authigenic,
indicated by its high level of crystallinity (Weaver, 1989; Cusack, 1991). The
conversion of kaolinite into dickite is observed by scanning electron microscope
(Appendix 7-2d). This technique showed that dickite is characterised as blocky
crystals, with thick smooth surfaces, thus it is different from kaolinite which is in
booklets and vermicular aggregates of thin crystals with etched surfaces.
Mixed-layer illite/smectite (0.2-48.5%) is usually the most common clay
mineral in the Illawarra Coal Measures. It was not identified in thin sections of
samples but SEM showed that mixed-layer illite/smectite coats most detrital and
authigenic grains, such as detrital quartz grains, carbonate cement, quartz
overgrowths and altered feldspar (Appendix 7-2c,e,f, 7-3d-e). Mixed-layer
illite/smectite is also recorded as a pore-filling cement (Appendix. 7-1d).
Illite (0-17.9%) occurs as fibrous crystals that are mainly oriented perpendicular
to detrital grain surfaces (Appendix 7-3c-f and 7-4a-b). It exists as a pore-filling
cement in some samples (Appendix 7-3f) where it is partly incorporated into
quartz overgrowths (Umar et al., 2011). In some samples, it is observed as high
birefringent patches. Some detrital grains and pores are bordered or coated
with illite (Appendix 7-4a-b). Grain-coating illite is present in two forms as ultrathin layers and thin mat-like crystals (Appendix 7-4a-b).

179

Chlorite (0-2.9%) is rarely present in the Illawarra Coal Measures. It is


identified as scattered platelets in pseudomatrix oriented perpendicular to grain
surfaces (Appendix 7-1a) and as rims around overgrowth grains (Appendix 7-1a
and 7-3a-b). Also, it exists as a replacive mineral in detrital grains and encloses
some quartz overgrowths (Appendix 7-4c-f). Appendix 7-1c shows chlorite as a
pore-filling cement.
6-2-2-1-3 Carbonate cement
Carbonate cement occurs in the Illawarra Coal Measures and varies between 0
and 60.4% with an average of 14.8% in sandstone and 13.5% in siltstone
(Appendix 5-7 and 5-10). The Darkes Forest Sandstone has the highest
percentage of carbonate cement at an average of 39.4% whereas the lowest
percentage occurs in the Wongawilli Coal at an average of 3.4% (Appendix 513, 5-14 and 8-5). Carbonate cement is recorded in the Wilton Formation at
20%. It then increases up the sequence from 6.2% in the Tongarra Coal to
39.4% in the Darkes Forest Sandstone (Appendix 5-13, 5-14 and 8-5). It
fluctuates between 3.4% and 20% from the Allans Creek Formation to the
Unnamed Member Three (Appendix 5-13, 5-14 and 8-5). It then decreases from
24.7% in the Unnamed Member Two to 6.3% in the Loddon Sandstone
(Appendix 5-13, 5-14 and 8-5). It is generally the first cement in the Illawarra
Coal Measures and includes ankerite, siderite, calcite and dolomite. Carbonate
minerals were determined by thin section and SEM. Thin sections showed two
or more types of carbonate cement are present in most samples. They are
recorded as pore-filling cement, as grain coatings and as replacement.
Carbonate replacement of detrital grains, such as rock fragment and quartz
grains, occurs in two forms: from the centre to the periphery or from the
periphery towards the centre of the grains (Appendix 6-15b-c).
Siderite (0-51.1%) is present as large euhedral, subhedral and anhedral
crystals (20 m) in most samples (Appendix 6-11c and 6-12a). Large crystals
are blocky and poikilotopic and exist as pore-filling cement (Appendix 6-11c and
6-12a) and as partial grain-coatings on detrital grains such as quartz and rock
fragments (Appendix 6-11c, 6-12a, 6-13a-b and 6-19b). Small crystals of porefilling siderite are observed in some samples (e.g. Appendix 6-13a-b). Also,
180

siderite cement is present as a replacive mineral in quartz and feldspar in some


samples (Appendix 6-12b-c). Siderite is usually now stained by iron oxide which
may be derived from alteration of ilmenite or from iron hydroxides (Karim et al.,
2010). More siderite is found in fine-grained samples than in the coarser
grained samples, thus it decreases observably with increasing grain size. Fluid
inclusions are present in the siderite cement but are uncommon. Spry (2000)
indicated that the siderite has commonly changed because of moist, oxidizing
conditions.
Ankerite (0-46.7%) is observed as a cement and replacement mineral. It
appears as small crystals (10-20 m) and large crystals (20 m; Appendix 614a-c). In both cases, it occurs as a pore-filling cement and coats some detrital
and authigenic grains, such as quartz, feldspar, lithic grains and quartz
overgrowths (Appendix 6-14a-c, 6-15a and 6-19c). In some samples, ankerite
cement is noted as a replacement mineral in detrital grains (Appendix 6-15b-c).
SEM displays euhedral ankerite with a rhomohedral shape occurring as a porefilling cement (Appendix 7-2b-c).
Calcite (0-44.7%) exists in most samples as large euhedral to anhedral crystals
(20 m). Also, it fills large and small pores in some samples (Appendix 6-16ab) and the margins of some grains are coated by calcite cement (Appendix 616a-b and 6-17b). Detrital grains such as quartz may be replaced partially or
totally by calcite cement in some samples (Appendix 6-16c and 6-17a). The
presence of fluid inclusions is rare in the calcite cement.
Dolomite (0-11.1%) is less widespread as a carbonate cement, although it
occurs in all units of the Illawarra Coal Measures (Appendix 5-13 and 5-14). It is
usually present as small crystals which are anhedral to euhedral (Appendix 618a and 6-13c). However, it was observed as replacive rhombs in some
framework grains and as a pore-filling cement (Appendix 6-13c, 6-17c and 618a). Some detrital grains are coated by dolomite cement as shown by thin
section (Appendix 6-13c, 6-17c and 6-18a).

181

6-2-2-1-4 Authigenic feldspar


In the Illawarra Coal Measures, authigenic feldspar is uncommon but it was
observed in thin section and SEM analysis in few samples (e.g. Appendix 7-1f,
7-3d and 6-20c).
Authigenic feldspars occur as overgrowths on detrital feldspar grains (Appendix
6-20c). In sandstones rich in volcanic rock fragments, albitization is most
common (Sur et al., 2002) whereas authigenic feldspar is rare in sandstone rich
in early calcite cement (Kim et al., 2007). The presence of detrital feldspar
grains together with supersaturation and circulation of pore water controls the
development of feldspar overgrowths (Wolela and Gierlowski-Kordesch, 2007).
6-2-2-1-5 Authigenic pyrite
Authigenic pyrite is absent in most samples and, where present, varies between
0 and 0.6% in the Illawarra Coal Measures. It occurs as a diagenetic mineral in
several forms including framboidal aggregates, scattered euhedral crystals and
as a pore-filling cement (Appendix 6-18b). Framboidal aggregates include
closely packed tiny crystals. Authigenic pyrite appears as a replacement within
some mudclasts.
6-2-2-1-6 Iron oxide cement
Iron oxide cement ranges from 0 to 15.1% in the Illawarra Coal Measures. Most
samples did not contain iron oxide cements. They coat the margins of detrital
grains and were observed as a pore-filling cement (Appendix 6-18c). Their
colour is mainly brownish.

6-2-3 Diagenetic Sequence


In the Illawarra Coal Measures, the diagenetic sequence consists of early and
late stages (Fig. 6-3). The early stage consists of authigenic clay minerals and
carbonate cement. The late stage includes dissolution/alteration of unstable
detrital grains, quartz overgrowths, late authigenic clay minerals, late carbonate
cementation, ferruginous cement and authigenic feldspar.

182

6-2-3-1 Compaction
Mechanical and chemical compaction are shown in both sandstone and
siltstone samples in the Illawarra Coal Measures. Chemical compaction is
determined through the increased grain contacts in the sequence. Three types
of contacts, including concavo-convex, sutured contacts and long contacts,
provide evidence for chemical compaction in the Illawarra Coal Measures
(Appendix 6-19a-b) whereas pressure solution is rare in the succession,
indicated by point contacts and long contacts (Wolela, 2009). Where mica is
present it is bent because of physical compaction (Appendix 6-9b). The
presence of pseudoplastic deformation of mud intraclasts, ductile grains and
grain arrangement provides evidence of mechanical compaction (Appendix 65a). Compaction is uncommon in some samples due to the presence of
widespread carbonate cement. Compaction is included in both the early and
late stages of diagenesis.

6-2-3-2 Cementation
6-2-3-2-1 Authigenic clay minerals
Authigenic clay minerals are characteristically oriented perpendicular to the
surface of detrital grains when they occur as pore-linings and pore-fillings. In the
Illawarra Coal Measures, the disaggregation and chemical breakdown of lithic
constituents is the source of clays (Bamberry, 1992).
Fluvial deposits include predominantly grain-coating clays (Morad et al., 2000;
Ketzer et al., 2003b; Worden and Morad, 2003; El-Ghali et al., 2006a). Graincoating clays may be recorded as mechanical infiltration (cf. Matlack et al.,
1989; Morases and De Ros, 1990, 1992; El-Ghali et al., 2006b), occurring as
coatings on detrital and overgrowth grains. Kaolinite, mixed-layer illite/smectite,
illite and chlorite are present as early diagenetic minerals in the Illawarra Coal
Measures.
Authigenic clay minerals, such as kaolinite, illite, chlorite and mixed-layer
illite/smectite commonly fill intergranular pore spaces (Appendix 7-1c, d, 7-2d
and 7-3c, f). The occurrence of kaolinite as grain-coatings on detrital grains or
as a pore-filling cement indicates that kaolinite was precipitated as an early
183

diagenetic mineral. The vermicular texture of kaolinite also indicates the origin
of kaolinite at an early stage (cf. Wilkinson et al., 2004; Abouessa and Morad,
2009). The textural characteristics of illite such as fibrous crystals, ultra-thin
layers and thin mats also suggest an early diagenetic origin for the illite. The
presence of illite in the form of pore-fillings and grain-coatings are also
interpreted as early authigenic illite. Mixed-layer illite/smectite is recorded as
grain coatings on most detrital grains (Appendix 7-2f, and 7-3d) and this
demonstrates its formation during early diagenesis. Also, authigenic kaolinite is
engulfed, and thus pre-dates, authigenic chlorite (Appendix 7-3b).
Pore-water is the source of kaolinite which occurs in the form of booklets (cf.
Irwin and Hurst, 1983; Wolela, 2009). Low pH and low ionic strength waters are
recorded as supportive factors for the precipitation of kaolinite (Bjorlykke et al.,
1989, 1986; Worden and Burley, 2003; Hammer et al., 2010). Authigenic
kaolinite can be produced by dissolution of carbonates, feldspar and unstable
minerals in lithic fragments (cf. Xu et al., 2003; Zhang et al., 2009; Appendix 647a). Alteration of volcanic rock fragments, breakdown and alteration of
feldspar, illitization of matrix constituents and earlier kaolinite can be the
sources of illite and chlorite (Kim et al., 2007). Replacement of feldspar may
also contribute to the formation of illite (Umar et al., 2011). In some samples,
carbonate grains are coated by clays, indicating that precipitation of the latter
was a result of dissolution and alteration processes (Bai, 1991). The formation
of smectite clays occurs during weathering, particularly in arid to semi-arid
climatic conditions (El-Ghali et al., 2009a).
The conversion of kaolinite to dickite is indicated by the presence of vermicular
stacked booklets of kaolinite converting to the thick to blocky habit of dickite
(Ehrenberg et al., 1993; McAulay et al., 1994; Morad et al., 1994, Lanson et al.,
2002 and Abouessa and Morad, 2009). According to Morad et al. (1994) and
Abouessa and Morad (2009), dissolution and re-precipitation can play an active
role in the conversion of kaolinite into dickite, which characteristically occurs at
a temperature of >100C (Ehrenberg et al., 1993; McAulay et al., 1994; Morad
et al., 1994, Lanson et al., 2002; Abouessa and Morad, 2009).

184

Many evidences showed chlorite as a pore-filling and grain-coating cement,


indicating its precipitation during early diagenesis. The occurrence of chlorite as
grain coatings may be due to detrital mineral transformation or the formation of
eogenetic clay minerals as suggested by (Gier et al., 2008). Authigenic chlorite
grain coatings are characterised by the presence of iron (Pittman et al., 1992;
Ehrenberg, 1993; Bloch et al., 2002). Chlorite growth is enhanced when the
temperature exceeds 60-70C (Worden and Morad, 2003). Many studies have
shown that the relative timing of authigenic chlorite precipitation in sandstone
occurs during early diagenesis, soon after the period of deposition and prior to
complete compaction (Ehrenberg, 1993; Aagaard et al., 2000; Grigsby, 2001;
Bloch et al., 2002; Billault et al., 2003). Chlorite may be precipitated as result of
ions which are derived from alteration of framework grains such volcanic rock
fragments (Bai, 1991). Dissolution of volcanic rock fragments and feldspar also
play a role in the formation chlorite in secondary pores by contributing Mg, Fe
and Si ions (cf. Umar et al., 2011). The precipitation of Mg and Fe ions may be
recorded as early diagenetic (Sur et al., 2002) and this leads to the
interpretation that pore-filling chlorite and chlorite grain coatings are early
diagenetic.
6-2-3-2-2 Carbonate cementation
Carbonate cementation mainly post-dates the authigenic clays and occurs as
grain-coatings around detrital grains (Appendix 6-11a, c, 6-12a, 6-13a-c, 6-14ac, 6-15a-c, 6-16a-b, 6-17b-c, 6-18a and 6-19b-c). This supports the concept that
carbonate cement is early diagenetic. The early formation of carbonate cement
is also indicated by the large crystal size of the pore-filling carbonate cement
(Appendix 6-11c, 6-12a, 6-14b and 6-16a). These results indicate that
carbonate cementation takes place during early diagenesis. Unformed ductile
grains are observed in sandstone rich in carbonate cement, supporting the
occurrence of carbonate cement as early diagenetic according to Salem et al.
(2000). Also, the occurrence of early diagenetic carbonate cement is confirmed
by the loose grain packing of the sandstone (cf. Salem et al., 2000). Quartz
overgrowths are rare in the Illawarra Coal Measures, which is rich in carbonate
cement. This indicates that carbonate cement was precipitated during early

185

diagenesis, and pre-dates the development of quartz overgrowths according the


interpretation of Odigi and Amajor (2010).
Pore-filling siderite and grain-coating siderite indicate that siderite also formed
during early diagenesis. Lee and Lim (2008) showed that the formation of
siderite during early diagenesis is evidenced by the existence of pore spaces
which were large enough for the development of euhedral siderite crystals
(Appendix 6-11c and 6-12a). However, the precipitation temperature for siderite
was raised, supported by high Mg and low Ca which are characteristic in the
siderite cement (Abouessa and Morad, 2009). Also, the high Fe/Ca supports an
interpretation of the precipitation of siderite cement within the early stages of
diagenesis (Berner, 1981; Lee and Lim, 2008). The large crystals of calcite
cement and poikilotopic calcite cement (e.g. Appendix 6-16a) as well as the
loose grain packing seen in thin sections also indicates that calcite cementation
occurred during early diagenesis (cf. Lee and Lim, 2008). The solubility product
can be exceeded as a result of increasing the carbonate ion and this supports
the precipitation calcite as an early stage diagenetic mineral.
6-2-3-2-3 Dissolution/alterations of unstable detrital grains
Several types of dissolution are recorded in the Illawarra Coal Measures as the
products of mid to late stage diagenesis. These are quartz dissolution, feldspar
dissolution, rock fragment dissolution and carbonate dissolution (Appendix 619c, 6-20a-b and 6-21a-b, 6-40a-c, 6-41a-c and 6-48a). The dissolution of
feldspar, carbonate cement and lithic grains provides secondary porosity
(Appendix 6-20a-b, 6-21a-b, 6-40a-c, 6-41a-c and 6-48a; Chi et al., 2003).
Feldspar albitization is a mid to late authigenic stage but is rare in the
succession (Appendix 7-1f, 7-3d and 6-20c). The precipitation of feldspar
overgrowths is associated with detrital feldspar dissolution because of the
greater chemical stability of authigenic albite with respect to detrital calcic
plagioclase (De Ros et al., 1994). High silica activities and high K+/H+ ratios are
important factors to support K-feldspar precipitation (Morad et al., 2000; Rossi
et al., 2002).

186

6-2-3-2-4 Authigenic quartz


Quartz overgrowths represent a mid to late authigenic mineral in this
succession and the occurrence of quartz cement as overgrowth around detrital
quartz grains supports this interpretation. The temperature suggested for the
precipitation of quartz overgrowths is between 90-130C (Hartmann et al., 2000;
Morad et al., 2000; El-Ghali et al., 2006) whereas in other studies it was
suggested to be between 40-60C (Land and Dutton, 1978; McBride et al.,
1988; Salem et al., 2000). Quartz overgrowths post-date the carbonate cement
and authigenic clay minerals. Also, quartz overgrowth crystals do not include
corrosion and this interpretation support that quartz overgrowth occurred as a
late authigenic mineral.
Several studies have shown that quartz cementation is derived from silica
released by carbonate mineral replacement of quartz grains, feldspar
dissolution and the transformation of clays (Hower et al., 1976; Boles and
Franks, 1979; Burley and Kantorowicz, 1986; Rezaee and Tingate, 1997; Kim et
al., 2007). Also, illitization of smectite, kaolinitization and K-feldspar albitization
are recorded as sources of silica in other studies (Morad, 1986; Luo et al.,
2009). In this study, pressure dissolution is recorded as the source of quartz
cement (cf. Kordi et al., 2011). The distribution of grain-coatings, infiltrated clays
and pseudomatrix control the quartz overgrowth distribution (cf. El-Ghali et al.,
2009a).
6-2-3-2-5 Late carbonate cementation
Thin section studies have indicated that ankerite cement was precipitated again
after the quartz overgrowths since it partially coats quartz overgrowths in some
samples (Appendix 6-11a and 6-15a); thus it is also a late authigenic mineral.
Also, late stage ankerite cement has re-precipitated as new cement following
dissolution of early ankerite cement (Appendix 6-20).
Late poikilotopic calcite may be formed because of carbonate grain or cement
dissolution (cf. Al-Ramadan et al., 2004). The occurrence of euhedral quartz
overgrowths that are embedded in calcite cement indicate that the quartz
overgrowths developed in pores prior to calcite cementation (Fig. 6-17b; cf.
Umar et al., 2011). Also, late stage calcite includes minor Mg and Fe that may
187

support the precipitation of calcite during this stage (Sur et al., 2002). Also, the
solubility product can be exceeded as a result of increasing carbonate ions by
evaporation of vadose or near-surface phreatic ground water, suggesting that
the precipitation of calcite cement during late diagenesis post-dated the quartz
overgrowths.
Also, Appendix 6-20a showed new late authigenic siderite cement has reprecipitated after dissolution of early siderite cement (Appendix 6-20a). Other
studies have shown that late stage diagenetic siderite cement has high Mg and
Ca contents (Curtis and Coleman, 1986; Rossi et al., 2001).
6-2-3-2-6 Late mixed layer illite/smectite and chlorite
Some authigenic clay minerals were precipitated again after the quartz
overgrowths (Appendix 7-1a, 7-2a, e and 7-3a, e). Mixed-layer illite/smectite is
developed on quartz overgrowths and carbonate cement (Appendix 7-2c, e).
This indicates that mixed-layer illite/smectite is precipitated during late
diagenesis. Detrital inheritance from K-, Ca- and Mg-rich silicates under the
influence of alkalinity conditions suggests the precipitation of mixed layer
illite/smectite and chlorite as late stage diagenetic minerals after the
development of quartz overgrowths.
Also, quartz overgrowths may be coated by chlorite (Appendix 7-1a and 7-3a),
indicating that the latter also formed as a late diagenetic mineral.
6-2-3-2-7 Ferruginous Cement
Ferruginous cement is present as the next event in the Illawarra Coal
Measures. It formed during late diagenesis as pore-filling cement and graincoatings on detrital grains.
Bloch et al. (2002) showed that iron is derived from a meteoric-water source
that is also indicated by the presence of chlorite and siderite cement. Also,
enrichment of the pore space chemistry plays a role in the distribution of
hematite (Wolela, 2009). Wolela (2009) showed several sources of hematite
cement including dissolution or alteration of iron silicates, carbonates, ilmenite
or magnetite.

188

6-2-3-2-8 Recrystallization
Additionally, silica and chert are recrystallised to microcrystalline quartz
(Appendix 6-3a and 6-8b) and micritic cement has recrystallised to microsparite
and

sparry

calcite

(Appendix

6-16a).

These

provide

evidence

that

recrystallization has occurred in the Illawarra Coal Measures.


In conclusion, the main diagenetic stages represented by the Illawarra Coal
Measures, arranged in a chronological order, are:
1) Mild compaction
2) Early authigenic clay minerals
3) Precipitation of carbonate cement
4) Dissolution and alteration
5) Secondary silica overgrowths
6) Late carbonate cement
7) Late authigenic clay minerals
8) Ferruginous cement
9) Recrystallization.

Diagenetic minerals

Early

Late

Compaction
Quartz overgrowth
Kaolinite
Dickite
Mixed-layer illite/smectite
Illite
Chlorite
Siderite
Ankerite
Calcite
Dolomite
Dissolution/alteration
Iron oxide

Fig. 6-3: Paragenetic sequence of the diagenetic alterations in the Illawarra


Coal Measures.

189

6-3 Narrabeen Group


6-3-1 Scanning Electron Microscope (SEM)
Scanning electron microscope analysis showed that detrital and authigenic clay
minerals in both the Gosford and Clifton Subgroups of the Narrabeen Group
include kaolinite, chlorite, illite and mixed-layer illite/smectite. Clay minerals are
prevalent in the Gosford Subgroup and in the Bald Hill Claystone which is
present at the top of the Clifton Subgroup. In both subgroups, clay minerals
occur as grain coatings on most detrital grains (Appendix 7-5a, c-f, 7-6a-c, e
and 7-7a, c), and as cements filling intergranular pores (Appendix 7-5a-b, 7-6d-f
and 7-7b).
Well developed quartz overgrowths, carbonate cement, and detrital grains such
as quartz and feldspar (Appendix 7-5c, f, 7-6b-c and 7-7a, e-f) are shown by
scanning electron microscope in the Narrabeen Group. Quartz overgrowths and
carbonate cement developed in the pores (Appendix 7-5c, 7-6b-c and 7-7a).
Primary and secondary pores (Appendix 7-5a-b, f, 7-6a-b, d and 7-7a, c-f)
range from large to small in the Narrabeen Group.

6-3-2 Diagenesis
Quartz overgrowths, authigenic clay minerals, carbonate cement, authigenic
feldspar, authigenic pyrite and iron oxide cement are the main authigenic
minerals in the Narrabeen Group. Bai and Keene (1996) indicated that the
Narrabeen Group includes common authigenic carbonate minerals, clay
minerals and quartz. Also, silica dissolution, dissolution of unstable grains such
as feldspar and volcanic rock fragments and dissolution carbonate cement were
also observed in the Narrabeen Group.

190

6-3-2-1 Diagenetic Minerals


6-3-2-1-1 Quartz cement
Quartz overgrowths range from 0 to 5.5% and are found in most samples
(Appendix 5-8). They are more common in quartz-rich sandstone, particularly
the Bulgo Sandstone (3.4%), than in lithic-rich sandstone (Appendix 5-13, 14
and 8-5). The percentage of quartz overgrowths increases upwards from 1.1%
in the Coalcliff Sandstone to 3.4% in the Bulgo Sandstone (Appendix 5-13, 14
and 8-5). One sample from the Bald Hill Claystone was described from thin
section and contains a low percentage (0.2%) of quartz overgrowths (Appendix
5-13, 14 and 8-5). Near the top of the Narrabeen Group, the Newport Formation
has 1.5% of quartz overgrowths (Appendix 5-13, 14 and 8-5). Carbonate is the
most common cement and the occurrence of infiltrated clays as coatings on
quartz grains means that quartz overgrowths are rare (Salem et al., 2000).
Thus, the development of quartz overgrowths is inhibited by grain-coating clays
in the Narrabeen Group.
Three types of quartz cement are present in the Narrabeen Group microquartz, mega-quartz and quartz overgrowths (Bai, 1991). Quartz overgrowths
form the most common type of quartz cement in the Narrabeen Group. They
occur as euhedral crystal faces and partially fill pores between grains (Appendix
7-5f, 7-6c, 7-7a, f, 6-21c, 6-24a and 6-51c). Different sizes of quartz overgrowth
are observed. Different types of boundaries occur between quartz cement and
detrital quartz grains including fluid inclusions, clay coatings and iron oxides
(e.g. Fig. 6-21c). Double quartz overgrowths were observed in rare samples
(Fig. 6-22a).
A moderate positive correlation is observed between quartz overgrowths and
porosity in the Scarborough Sandstone (r = 0.4; Fig. 6-4).

191

y = 5.739x - 4.874
R = 0.442

25

Porosity (%)

20
15
10
5
0
0

0.5

1.5

2.5

3.5

Quartz overgrowth (%)

Figure 6-4: A moderate positive correlation occurs between


quartz overgrowths and porosity in the Scarborough Sandstone.

6-3-2-1-2 Authigenic clay minerals


Authigenic clay minerals include kaolinite, illite, mixed-layer illite/smectite and
chlorite. They are common and form the second most abundant cement in the
Narrabeen Group.
Kaolinite (0.2-61.3%) is distributed throughout the whole of the Narrabeen
Group and is the most common authigenic clay. Quartz-rich sandstone contains
a greater amount of authigenic kaolinite than lithic-rich sandstone because it
contained more available pore spaces. Kaolinite exists as booklets and
vermicular aggregates whether separated or in groups, and has rounded to subrounded boundaries (Appendix 7-5a-b, f and 7-6a, e). Kaolinite fills pore spaces
and occurs as coatings on the margins of pores (Appendix 7-5a-b, f and 7-6a,
e). In some samples, it is coated by illite and mixed-layer illite/smectite
(Appendix 7-5a, f and 7-6e). Bai (1988) showed that kaolinite is the most
widespread diagenetic clay mineral in the Narrabeen Group. This study shows
similar results to Bai (1991) in that lithic-rich sandstone is characterised by
authigenic kaolinite as result of porosity alteration. The conversion of kaolinite
into dickite was observed using SEM where blocky crystals of dickite are thicker
than booklets and vermicular aggregates of kaolinite (e.g. Appendix 7-11d).
192

Mixed-layer illite/Smectite (2.5-29.6%). SEM studies indicated that mixedlayer illite/smectite is common in the Narrabeen Group. In most samples, it
coats detrital and authigenic grains and exists as a pore-filling cement
(Appendix 7-5c-d, f and 7-6b-d).
Illite (0-11.9%) occurs as fibrous crystals and sheets (Appendix 7-5a, 7-6f and
7-7 d). It is observed as grain coatings on kaolinite and quartz overgrowths
(Appendix 7-6e and 7-7a). Grain-coating illite is characterised as ultra-thin
layers and thin mat-like crystals (Appendix 7-6e and 7-7a). Also, in most
samples many intergranular pore spaces are filled with authigenic illite (e.g.
Appendix 7-6f; Mckinley et al., 2011). It is typically oriented perpendicular to
grain surfaces and has a high birefringence.
Chlorite (0-7.3%) is present as authigenic rims around detrital grains (Appendix
7-7c). It is visible as a pore-filling cement and as grain-coatings in some
samples (e.g. Appendix 7-5e and 7-7b-c; Bai and Keene, 1996). Pore-filling
chlorite is thicker than grain-coating chlorite. Pore-filling chlorite cement has a
thickness ranging between 4 and 6 m, whereas grain-coating chlorite has a
thickness range of 1 to 2 m. Grain-coating chlorite is not associated with porefilling chlorite in the same sample. Moreover, chlorite is associated with mixedlayer illite/smectite and illite in the succession (Appendix 7-7c-d). Chlorite is
most common in volcanic-rich sandstone.
6-3-2-1-3 Carbonate cement
Most samples are cemented by carbonate which is the dominant cement in the
Narrabeen Group. Carbonate cement is recorded in all samples varying
between 1.5% and 59% (Appendix 5-8). It is present at an average of 10.7% in
sandstone and 25.3% in siltstone (Appendix 5-11). The Coalcliff Sandstone,
Wombarra Claystone, Scarborough Sandstone, Stanwell Park Claystone and
Bulgo Sandstone have average carbonate cement contents of 10.8%, 12.7%,
11.4%, 18.5% and 5% (Appendix 5-13, 14 and 8-5). One sample from the Bald
Hill Claystone has a higher percentage of carbonate (20.5%; Appendix 5-13, 14
and 8-5). The Newport Formation at the top of the Narrabeen Group contains
7.1% carbonate cement (Appendix 5-13, 14 and 8-5). Siderite and ankerite are
the most common carbonate cements in both the sandstone and siltstone with
193

combined average contents of 6% and 15%, respectively (Appendix 5-11). Less


common calcite and dolomite are also present (Bai and Keene, 1996). Two or
three types of carbonate cementation are observed in most samples.
Siderite (0-24.1%) is observed in all but two samples in the Coalcliff Sandstone
(Appendix 5-8). Siderite occurs as small rhombohedral crystals and as coarse
crystalline cement (Appendix 6-22b-c and 6-23a-b). Coarse crystalline siderite
is the most widespread carbonate cement in the studied samples, and it
includes blocky and poikilotopic crystals (Appendix 6-22b-c). Large and small
pore spaces are filled with siderite cement that is also present as grain-coatings
on detrital and authigenic grains in most samples (Appendix 7-5c, 7-6b, 6-22bc, 6-23a-c and 6-24a). Many detrital grains are enclosed by siderite cement
(Appendix 6-23a-c) and some detrital grains, such as feldspar and rock
fragment, are replaced by siderite cement (e.g. Appendix 6-24b-c). Fluid
inclusions are uncommon in siderite cement. Siderite is now generally stained
with iron oxide because of modern weathering. Alteration of the siderite,
ilmenite or iron hydroxides may be the source of the iron oxide coating the
siderite (Karim et al., 2010).
Ankerite (0-44.6%) is common in siltstone samples (15%; Appendix 5-11). Few
samples did not include ankerite cement. Three forms of ankerite were
determined including micritic, microcrystalline and coarsely crystalline (e.g.
Appendix 6-25a-c and 6-26a). It fills large and small pores spaces between
grains (Appendix 6-25a-c and 6-26a) where it occurs as subhedral or euhedral
crystals (Bai, 1988). Ankerite cement also occurs as coatings occasionally
enclosing the margins of quartz, feldspar, chert, rock fragments and also quartz
overgrowths (Appendix 6-25a-c and 6-26a-b). Ankerite cement also occurs as a
partial replacement of grains such as rock fragments (Appendix 6-26b).
Calcite (0-27.2%) is absent in most samples but where it is present, it occurs in
very minor amounts. It is present as microcrystalline to coarse crystalline
cement and may fill pore spaces (e.g. Appendix 6-27a-b). Grain-coating calcite
on quartz and rock fragments is also recorded in this succession (Appendix 627a-b). Minor calcite replacement of detrital grains occurs in most samples,
particularly in the lower part of the Narrabeen Group (Appendix 6-27a, c and 6194

28a-c). Fluid inclusions are uncommon in calcite cement and occur as scattered
spheres.
Dolomite (0-10.8%) exists in all units of the Narrabeen Group, although it
occurs in trace amounts as anhedral to euhedral crystals (Appendix 6-26c and
6-29b-c). Thin sections showed the presence of dolomite as a replacement
mineral in detrital grains in some samples (Appendix 6-29a and 6-30a). This
technique also showed dolomite cement covering quartz, volcanic rock
fragments and quartz overgrowths (Appendix 6-26c, 6-29a-c, 6-30a and 31b).
Dolomite is also found as a pore-filling cement (Appendix 6-26c, 6-29b-c, 6-30a
and 31b).
6-3-2-1-4 Authigenic feldspars
Alteration of feldspars is recorded in the Narrabeen Group (e.g. Appendix 6-32a
and 6-47b); occurring most commonly in volcanic rock fragments (Sur et al.,
2002). Albitization of feldspar is also observed in the Narrabeen Group (Fig. 631c).
6-3-2-1-5 Authigenic pyrite
Authigenic pyrite ranges from 0 to 0.3% and is missing in most samples. It is
present as framboidal aggregates or scattered small euhedral crystals in the
Narrabeen Group (Appendix 6-30a). Authigenic pyrite is also observed as a
pore-filling cement and as replacement mineral within mud intraclasts.
6-3-2-1-6 Iron oxide cement
In the Narrabeen Group, iron oxide cement is a brownish colour and does not
exceeding 0.8%, being absent in the majority of samples. It occurs as a porefilling cement and as a coating on detrital grains (Appendix 6-30b).

6-3-3 Diagenetic Sequence


In the Narrabeen Group, diagenetic sequences are composed of early and late
stages (Fig. 6-5).

195

6-3-3-1 Compaction
Mechanical and chemical compaction are recorded in the Narrabeen Group.
Mechanical compaction is more common than chemical compaction in the
succession. Mechanical compaction is demonstrated by deformation of ductile
grains and mica flakes between harder grains and by grain arrangement
(Appendix 6-30c). Mechanical compaction is more widespread in sandstone rich
in detrital lithic grains than in sandstone rich in detrital quartz. Early authigenic
clay minerals and carbonate cement prevent some mechanical compaction.
However, detrital lithic grains are also affected by dissolution and alteration as
well as mechanical compaction. Mechanical compaction is the main factor in the
porosity and permeability reduction in the Narrabeen Group. Chemical
compaction is evidenced by concavo-convex, sutured contacts and long grain
contacts (e.g. Appendix 6-31a) with rare pressure solution in the Narrabeen
Group (Appendix 6-46a). Compaction occurred during both the early and late
stages of diagenesis.

6-3-3-2 Cementation
6-3-3-2-1 Authigenic clay minerals
Authigenic clay minerals, including kaolinite, mixed-layer illite/smectite, illite and
chlorite are early authigenic minerals in the Narrabeen Group (Bai, 1991; Bai
and Keene, 1996). The early formation of authigenic clay minerals is evidenced
through the presence of clay minerals as grain coatings on detrital grains and
as pore-filling cement (Appendix 7-5a-b, d, f, 7-6a, d-f and 7-7b-e).
Margins of many large pores are partly coated by kaolinite (Appendix 7-6a).
Thus, grain-coating kaolinite and pore-filling kaolinite support the interpretation
that kaolinite was formed during early diagenesis. Also, the occurrence of
kaolinite in the form of vermicular texture supports this interpretation (cf.
Wilkinson et al., 2004; Abouessa and Morad, 2009). Detrital grains such
feldspar are commonly coated by mixed-layer illite/smectite (Appendix 7-5d),
indicating that illite and mixed-layer illite/smectite are early diagenetic in origin.
Mixed-layer illite/smectite is present as coatings on authigenic kaolinite
indicating that mixed-layer illite/smectite post-dates kaolinite (Appendix 7-5f).
196

Authigenic chlorite is interpreted as an early diagenetic mineral, evidenced by


pore-filling chlorite and grain-coating chlorite.
According to Bai (1991), precipitation of clay minerals, particularly kaolinite and
mixed-layer illite/smectite, occurred from pore water under oxygenated and
mildly acidic conditions. Also, early alteration of unstable detrital grains led to
the formation of kaolinite and mixed-layer illite/smectite (Bai, 1991; Appendix 647b). Precipitation of kaolinite occurs from low pH and low ionic strength waters
(Bjorlykke et al., 1989, 1986; Worden and Burley, 2003; Hammer et al., 2010).
Replacement of feldspar may also contribute to the formation of illite (Umar et
al., 2011) and, also high K+ ions support the precipitation of illite (Bai, 1991). In
the Narrabeen Group, dissolution and alteration processes lead to the
precipitation of clays which occur as grain-coating on surfaces of carbonate
grains (Figs 6-5c and 6-6b; Bai, 1991).
During burial diagenesis, the conversion of kaolinite into dickite is observed,
indicating that vermicular stacks and booklets of kaolinite are present along with
thick to blocky dickite (Ehrenberg et al., 1993; McAulay et al., 1994; Morad et
al., 1994, Lanson et al., 2002 and Abouessa and Morad, 2009). Two factors are
important in the conversion of kaolinite into dickite: dissolution and reprecipitation (Morad et al., 1994; Abouessa and Morad, 2009). The conversion
of kaolinite into dickite occurs at temperatures >100C (Ehrenberg et al., 1993;
McAulay et al., 1994; Morad et al., 1994, Lanson et al., 2002 and Abouessa and
Morad, 2009).
In the Narrabeen Group, the precipitation of grain-coating chlorite and porefilling chlorite cement occurred from pore water according to Bai (1991) who
showed that anoxic and neutral to mildly alkaline pore-waters are the source for
the precipitation of chlorite. Temperatures for growth of chlorite are about 6070C (Worden and Morad, 2003). In the Narrabeen Group, Bai (1991) indicated
that alteration of framework grains such volcanic rock fragments are significant
in the formation of chlorite. Umar et al. (2011) suggested that Mg, Fe and Si
ions are important in the formation of chlorite in secondary pores, which result
from dissolution of volcanic rock fragments. The precipitation of Mg and Fe ions

197

may be recorded in early diagenetic chlorite (Sur et al., 2002) and this supports
that pore-filling chlorite and chlorite grain coatings are early diagenetic.
6-3-3-2-2 Carbonate cement
Carbonate cement fills large pores in most samples indicating its early
diagenetic formation (Appendix 6-22b-c, 6-23a-c, 6-24a, 6-25a-c, 6-26a, c, 627b, 6-29a, 6-30a and 6-31a-b). It was precipitated after some of the authigenic
clays such as illite and kaolinite. Some detrital framework grains occur floating
within the carbonate cement which also indicates its early precipitation. The
precipitation of carbonate cement as grain-coatings on quartz, feldspar and rock
fragments also supports the early formation of carbonate cement (Appendix 622b-c, 6-23a-c, 6-24a, 6-25a-c, 6-26a-c, 6-27a-b, 6-29a-c, 6-30a and 6-31a-b).
Also, the replacement of detrital grains by carbonate cement supports this
interpretation. In the Narrabeen Group, sandstone commonly shows loose grain
packing and sandstone rich in carbonate cement is characterized by the
presence of unformed ductile grains. Both these features confirmed the early
diagenetic stage of carbonate cementation (cf. Salem et al., 2000). Also,
euhedral siderite crystals, filling large pore spaces (Appendix 6-22b) can be
interpreted as early diagenetic siderite. The occurrence of calcite in the form of
large and poikilotopic crystals (e.g. Appendix 6-27a) support the interpretation
of early diagenetic calcite (cf. Lee and Lim, 2008).
The relative concentrations of Ca2+, Fe2+, Mg2+ and Mn2+ in pore water
contribute to the determination of the type of carbonate cement. Siderite or
ankerite crystallised where high concentrations of Fe2+ are present, whereas the
occurrence of non-ferroan calcite is associated with low Fe2+ and high Ca2+ (Bai,
1991). Also, Bai (1991) showed the important factors that support siderite
formation include negative Eh, low calcium ion activity, low sulphide ion
concentration, high carbon dioxide and high ferrous ion activity. Raised
precipitation temperatures for siderite are supported by high Mg and low Ca
which are characteristic in siderite cement (Abouessa and Morad, 2009). Also,
high Fe/Ca supports the interpretation of the precipitation of siderite cement
within the early stage of diagenesis (Berner, 1981; Lee and Lim, 2008). Also,
geochemical studies such as Estupinan et al (2007) in fluvial environments
198

showed that the precipitation siderite as an early diagenetic mineral is


compatible with a low Sr content. Other authors such as Mozley (1989), Morad
(1998), Morad et al (2000) and Estupinan et al (2007) indicated the precipitation
of siderite during early diagenesis in fluvial environments is characterized by
high FeCO3 and low MnCO3 .
6-3-3-2-3 Dissolution/alterations of unstable detrital grains
Dissolution/alterations of unstable detrital grains occurred during the mid to late
stage of diagenesis. Dissolution processes are indicated by secondary pore
space, silica dissolution, feldspar dissolution, lithic fragment dissolution and
carbonate dissolution (Appendix 6-31b, 6-32b-c, 6-33a-c, 6-42c, 6-43a-c, 6-44ac, 6-45a and 6-51a). Secondary porosity is created as a result of dissolution of
feldspar, carbonate cement and lithic grains (Appendix 6-32b-c, 6-33a-c, 6-42c,
6-43a-c, 6-44a-c, 6-45a, 6-46c and 6-51a). The presence of CO2 in the
groundwater led to the formation of dissolution porosity in the Narrabeen Group.
Meteoric water and the maturation of organic matter are the source for the CO2.
Dissolution processes are followed by alteration of unstable detrital grains (Bai,
1991). In some samples, feldspar dissolution is associated with replacement by
authigenic cement (Fig. 6-37b).
Alteration of unstable detrital grains led to the formation of kaolinite+quartz, or
illite+quartz or authigenic kaolinite (Bai, 1991). This also happened under acidic
conditions.
Authigenic feldspar precipitation occurred during mid to late stage diagenesis in
the Narrabeen Group as indicated by the alteration of feldspar. Thus, the
dissolution of detrital K-feldspar is the main origin of authigenic K-feldspar (cf.
Wolela, 2009). Factors that contribute to K-feldspar precipitation include high
silica activities and high K+/H+ ratios as indicated by (Morad et al., 2000).
6-3-3-2-4 Authigenic quartz
Quartz overgrowths represent a mid to late authigenic mineral in the Narrabeen
Group, occurring as a pore-filling cement (Appendix 7-5e-f, 7-6c, 7-7a, 6-21c
and 6-24a). Some stacked kaolinite flakes are enclosed by quartz overgrowths,
which also engulf authigenic carbonate crystals. These features indicate that
199

quartz overgrowths post-date carbonate cementation and the authigenic clay


minerals. Bai (1991) confirmed that the late formation of quartz overgrowths is
indicated by aqueous fluid inclusions in the quartz overgrowth which have an
average homogenisation temperature of 90-120C.
Detrital quartz grains are the source of silica in the Narrabeen Group, thus lithicrich sandstones only contain rare quartz overgrowths (Bai and Keene, 1996).
Also, early compaction is the reason for the absence of pore spaces, preventing
circulation of pore fluids (Sur et al., 2002). The silica for the quartz overgrowths
was derived from pressure solution indicated by sutured contacts between
quartz grains. Also, the occurrence of quartz overgrowths with kaolinite
(Appendix 7-5f and 7-11d) shows that quartz cement may be sourced from
silicon dioxide through feldspar dissolution according to Zhang et al. (2010).
Umar et al. (2011) indicated that alternative sources of silica for quartz
cementation are the dissolution of feldspar and volcanic rock fragments,
chloritization, kaolinitization and chemical compaction.
6-3-3-2-5 Late carbonate cements
Carbonate cementation occurred again after quartz overgrowths during late
diagenesis. Siderite, ankerite and dolomite are all observed as pore-filling
cements that partly coat quartz overgrowths in some samples (Appendix 6-24a
and 6-26a, c). This indicates that siderite, ankerite and dolomite continued to be
precipitated during late diagenesis. Mg and Ca contents are higher in the late
siderite cement (Curtis and Coleman, 1986; Rossi et al., 2001). In other
samples, late diagenetic carbonate cements are not present.
Discontinuous carbonate cement and oversized pores are created by
dissolution of earlier carbonate cement (cf. Al-Harbi and Khan, 2008). Late
carbonate cement such as siderite, ankerite and dolomite may be formed as
new carbonate cements in some samples after dissolution of early carbonate
cement (Appendix 6-32c and 6-33b-c; cf. Chi et al., 2003).
6-3-3-2-6 Late authigenic clays
After quartz overgrowths, illite was precipitated again and coats some quartz
overgrowths (e.g. Appendix 7-7a). Mixed-layer illite/smectite was also
200

precipitated as grain coatings on quartz overgrowths and carbonate cement


(Appendix 7-5c-d, f and 7-6b-c). Chlorite and kaolinite also occur later and coat
quartz overgrowth (Appendix 7-5e and 7-11d). This evidence indicates that
some mixed-layer illite/smectite, illite and kaolinite post-date the quartz
overgrowths and carbonate cement. These precipitations indicate the late
formation of some authigenic clay minerals. Detrital inheritance from K-, Caand Mg-rich silicates under alkaline conditions suggests the precipitation of
mixed layer illite/smectite and chlorite as late stage diagenetic minerals after the
development of quartz overgrowths.
6-3-3-2-7 Ferruginous Cement
Ferruginous cement also developed during late diagenesis in the Narrabeen
Group. It fills intergranular porosity. Detrital grains such quartz are partially
coated by ferruginous cement. This evidence indicates the late formation of
ferruginous cement. Hematite cement may be derived from dissolution or
alteration of iron silicates, carbonates, ilmenite or magnetite (Wolela, 2009).
In conclusion, the chronological order of the diagenetic phases is as follows:
1) Mild compaction
2) The formation of authigenic clays
3) Precipitation of carbonate cement
4) Silica dissolution, feldspar dissolution, rock fragment dissolution and
carbonate cementation dissolution
5) Secondary silica overgrowths
6) Late carbonate cement
7) Late authigenic clay minerals
8) Ferruginous cement

201

Diagenetic minerals

Late

Early

Compaction
Quartz overgrowth
Kaolinite
Dickite
Mixed-layer illite/smectite
Illite
Chlorite
Siderite
Ankerite
Calcite
Dolomite
Dissolution/alteration
Iron oxide

Fig. 6-5: Paragenetic sequence of the diagenetic alterations in the Narrabeen


Group.

6 -4 Hawkesbury Sandstone
6-4-1 Scanning Electron Microscope (SEM)
Eight samples were examined using SEM to describe diagenetic alteration,
porosity and permeability in the Hawkesbury Sandstone.
SEM showed clay minerals represented by kaolinite, illite and mixed-layer
chlorite-smectite. Clay minerals are both detrital and authigenic. Kaolinite is
present as booklets and vermicular aggregates (Appendix 7-8a-b, f, 7-9a-b, d,
7-10a-c and 7-11c), while illite occurs as fibrous crystals (Appendix 7-8c). Also,
early and late stage quartz overgrowths, altered detrital feldspar, and detrital
muscovite are shown by SEM (Appendix 7-8e and 7-10a-b, d-e and 7-11a, d-f).
SEM indicated both small and large pores often associated with authigenic
kaolinite (Appendix 7-8b, d, and f). Primary and secondary pores were observed
and are discussed in chapter seven.

202

6-4-2 Diagenesis
The diagenetic alteration in the Hawkesbury Sandstone was described from thin
section and SEM data. Authigenic minerals include quartz overgrowths,
authigenic clay minerals, carbonate cement, authigenic feldspar, authigenic
pyrite and iron oxide. Also, three dissolution stages have been determined.
They are carbonate cement dissolution, silica dissolution and feldspar
dissolution.

6-4-2-1 Diagenetic Minerals


6-4-2-1-1 Quartz cementation
In the Hawkesbury Sandstone, most samples include authigenic quartz which is
identified as syntaxial overgrowths on quartz grains and as euhedral crystals
(Standard, 1969). It is also present as a pore-filling cement and is widespread
close to places with long intergranular contacts (e.g. Appendix 6-34a).
Authigenic quartz content varies between 0.5 % and 9.1% (Appendix 5-9) with
an average of 6.8% in sandstone and 1.8% in siltstone (Appendix 5-12). Quartz
overgrowths are usually the first cement in this unit. Sandstone contains more
quartz overgrowths than siltstone because of the higher percentage of quartz
grains in sandstone. Thus, the presence of quartz overgrowths is controlled by
detrital quartz grains and large pores, as indicated by the high positive
correlation between detrital quartz and quartz overgrowth (r = 0.8; Fig. 6-6).
Also, quartz overgrowth abundance shows a moderate positive correlation with
grain size (r = 0.4; Fig. 6-7). Coarse-grained sandstones contain common
quartz overgrowths and have high porosity. Thus, with increasing grain size and
porosity, quartz cement increases.

203

y = 0.135x - 1.968
R = 0.829
10

Quartz overgrowth (%)

9
8
7
6
5
4
3
2
1
0
0

20

40
60
Quartz (%)

80

100

Figure 6-6: Strong positive correlation between quartz and


quartz overgrowths in the Hawkesbury Sandstone.

204

y = 48.05x + 54.33
R = 0.357

700

Grain size (microns)

600
500
400
300
200
100
0
0

4
6
Quartz overgrowth (%)

10

Figure 6-7: Moderate positive correlation between quartz overgrowths


and grain size in the Hawkesbury Sandstone.

Quartz overgrowths occur along the detrital quartz grain edge, perpendicular to
the quartz grains (Appendix 6-34a). Some authigenic quartz contains local
anhedral terminations, or is intergrown with a clay phase (Appendix 7-8e, 710a-b, d and 7-11a). The clay coatings and fluid inclusions determine the
boundaries between detrital quartz and quartz overgrowths (Appendix 6-34a).
The overgrowth features that characterise quartz may be changed by authigenic
clay minerals (Rossi et al., 2002; Umar et al., 2011). Fluid inclusions have sizes
exceeding 3 m but are uncommon. Appendix 6-34a shows three stages of
quartz overgrowths. Quartz overgrowth of first stage is overlain by quartz
overgrowth of a second stage, which is overlain by quartz overgrowth of the
third stage. This indicates the presence of reworked quartz from a previous
sandstone. The first stage occurred by silica released from decomposition of
less stable minerals whereas the migration of silica-rich fluids from distant sites
contributed to the formation of the second stage. Recrystallization of quartz
supported the formation of the third stage of overgrowths.
6-4-2-1-2 Authigenic clay minerals
In the Hawkesbury Sandstone, authigenic clay minerals principally comprise
kaolinite, illite, mixed-layer illite/smectite and chlorite. The authigenic clay
205

minerals form the second most abundant cement in the Hawkesbury Sandstone
with kaolinite being the most widespread authigenic clay mineral. They fill pores
and replace detrital grains and earlier clays (Spry, 2000).
Kaolinite (1.1-39%) is ubiquitous in the Hawkesbury Sandstone. It is present as
booklets and vermicular aggregates and occurs in the mud matrix,
pseudomatrix and detrital grains (Appendix 7-8a-b), as a pore-filling cement
(Appendix 7-8b-c, d, f and 7-9a-b), and as grain coatings (Appendix 7-8e and 710a-b). Primary and secondary pore spaces are often filled by authigenic
kaolinite (Appendix 6-8b-c, d and f). Kaolinite is present with illite and quartz
overgrowths in most samples (Appendix 7-8a, c-e, 7-9a-b, d and 7-10a, b, d).
Quartz overgrowths are enclosed by authigenic kaolinite and this indicates that
the latter was precipitated after the quartz overgrowths (Appendix 7-8c and 710a-b). Kaolinite formation depends on the presence of porosity and
permeability, which leads to migration of interstitial pore waters. The authigenic
kaolinite tends to increase in areas which include poorly developed authigenic
feldspar (Wolela, 2009). The conversion of kaolinite into dickite was observed
using SEM (Appendix 7-9f). The SEM images showed that dickite is
characterised as thick blocky crystals with smooth surfaces, thus it is different
from kaolinite which occurs as booklets and vermicular aggregates with thin and
etched surfaces.
Mixed-layer illite/smectite (0.1-22.3%) is the second most abundant clay
mineral in the Hawkesbury Sandstone (Appendix 7-10c-d). Mixed-layer
illite/smectite has a honeycomb-like texture. Mixed-layer illite/smectite is
observed as pore-lining to pore-filling clay with a ragged-platy morphology
(Appendix 7-10c-d). It occurs within quartz overgrowths in some samples (e.g.
Appendix 7-10d). Mixed layer illite/smectite is present as grain coatings on
authigenic kaolinite and is associated with illite (Appendix 7-10c-d).
Illite (0-6.5%) is visible as hair like wisps and as fibrous illite within kaolinite, the
mud matrix and mud intraclasts (Appendix 7-8a, c, d, 7-9a-d and 7-10d-e). It is
oriented perpendicular to grain surfaces and has a high birefringence (Appendix
7-8c-d). In most samples, illite is associated with kaolinite (Appendix 7-8a, c, d
and 7-9a-b, d). It fills pores spaces and is observed as grain-coatings (Appendix
206

7-8c-d, 7-9a-d and 7-10d-e). Grain-coating illite thickness ranges from thin to
thick with ultra-thin layers and thin mat like crystals (Appendix 7-8d and 7-9a-b).
The coatings are identified as continuous to discontinuous layers.
Chlorite (0-2.8%) exists in the succession as a pore-filling cement and as graincoatings on authigenic minerals (Appendix 7-10f and 7-11a). It occurs as
scattered crystals, as rims on quartz overgrowths and is intergrown with
authigenic kaolinite (Appendix 7-11a, c). In some samples, chlorite is also
present as grain-coatings on kaolinite (Appendix 7-9e).
6-4-2-1-3 Carbonate cement
Carbonate cement is the third most abundant cement in the Hawkesbury
Sandstone succession, ranging from 0 to 20.1% (Appendix 5-9) with an average
of 2.8% in sandstone and 14.6% in siltstone (Appendix 5-12). The carbonate
cements consist of siderite, ankerite, calcite and dolomite. The presence of
carbonate cement is more common in fine-grained samples than in coarsegrained samples. Moderate negative correlation is recorded between carbonate
cement and grain size (r = 0.4; Fig. 6-8). Thus, carbonate cement mostly
decreases with increasing grain size and porosity.
y = -16.62x + 424.9
R = 0.387

700

Grain size (micron)

600
500
400
300
200
100
0
0

10
15
Carbonate cement (%)

20

25

Figure 6-8: Moderate negative correlation between carbonate


cement and grain size in the Hawkesbury Sandstone.

207

Siderite (0-15.7%) is the dominant carbonate cement in most samples, at an


average of 2% in sandstone and 12.4% in siltstone (Appendix 5-9 and 5-12).
Siderite exists as fine- and medium-grained rhombs, as microcrystalline siderite
( 50 m) and as coarse crystalline siderite (up to 200 m; Appendix 6-34b-c
and 6-35a-b). Siderite is usually stained by iron oxide and contains rare fluid
inclusions. Also, thin sections showed that siderite has developed in the pores
between quartz grains and also occurs as grain coatings (Appendix 6-34b-c, 635a-b, c and 6-38b). Siderite cement is observed with quartz grains in most
samples where it fills small pores between tightly packed quartz grains
(Appendix 6-34b-c and 6-35a-b, c).
Ankerite (0-13.2%) is missing in most samples. One sample contained 13.2%,
while a few samples have trace amounts (Appendix 5-9). Ankerite cement also
fills pore between quartz grains (Appendix 6-36a-b) and is present as
poikilotopic ankerite cement (Appendix 6-36b).
Dolomite (0-5.4%) is also only recorded in a few samples, while trace amounts
of calcite are restricted to four samples (Appendix 5-9). Thus, calcite cement is
less prevalent than dolomite cement.
Ferroan calcite (0-0.8%) is uncommon in the Hawkesbury Sandstone and is
associated with detrital carbonate components. Low permeability for the
circulation of Mg-rich fluids is a reason to allow ferroan calcite to survive
dolomitization.
6-4-2-1-4 Authigenic Feldspars
Authigenic feldspars are rarely present in the Hawkesbury Sandstone as Kfeldspar and plagioclase alteration (Appendix 6-5c and 7-11e). They are
observed as overgrowths on detrital feldspar grains.
In the Hawkesbury Sandstone, feldspar is present as incipient overgrowths
(Appendix 7-11e), in some cases, intergrown with an authigenic clay phase.
Moderately high temperatures, plentiful Na+ and/or K+ ions, and a source of
silica from hydrolysing silicates are important conditions for the formation of
feldspar overgrowths. It may be that some of the detrital feldspars, which tend
208

to be intermediate in both the plagioclase or alkali feldspar series and are less
stable than the pure end members at low temperatures, thus they first partially
dissolve and then reprecipitate. This would be a form of distillation of an
impure to pure feldspar, and would explain the general dependence of
authigenic feldspar on the presence of detrital feldspars in most of the studied
samples.
6-4-2-1-5 Iron oxide cement
Iron oxide cement is observed in the ten surface samples in the Hawkesbury
Sandstone (Appendix 6-36c) at an average of 0.5% in sandstone and 0.1% in
siltstone (Appendix 5-12). It is a dark reddish colour and fills pores between
grains. Also, iron oxides are observed as amorphous coatings on grains.

6-4-3 Diagenetic Sequence


The diagenetic sequence in the Hawkesbury Sandstone shows the relative
timing of diagenetic changes during both early and late stages (Fig. 6-9). The
full paragenetic sequence is not present in all the individual samples but it is
dependent on all the samples in the Hawkesbury Sandstone.

6-4-3-1 Compaction
Two types of compaction are recognised in the Hawkesbury Sandstonemechanical and chemical compaction. Mechanical compaction is represented in
the Hawkesbury Sandstone by bent flexible grains such as mica (Appendix 637c), and pseudoplastic deformation of mud intraclasts forming pseudomatrix
(Appendix 6-36a and 6-48c). Compaction is recognizably more abundant in
sandstone than in siltstone samples. Mechanical compaction post-dates some
of the main cementing minerals. If calcite cement is uncommon, compaction is
more abundant (Umar et al., 2011). Compaction occurred during both early and
late stages of diagenesis. Chemical compaction is indicated by pressure
dissolution, long and concavo-convex grain contacts and sutured contacts
(Appendix 6-37a-b). The presence of pressure dissolution is identified along
intergranular contacts (Appendix 6-37a). Widespread pressure dissolution
occurs when the quartz grains are covered with thin illite coatings or when mica
occurs at the interface along the quartz grain contacts (Appendix 6-37c).
209

Some ductile components are mainly compressed. Additionally, chemical


dissolution of detrital grains has occurred because of chemical compaction.

6-4-3-2 Cementation
6-4-3-2-1 Authigenic clay minerals
SEM studies indicated that kaolinite pore-filling and kaolinite grain-coating
occurs as an early authigenic mineral. Edges of large pores are coated by
booklets and vermicular aggregates of kaolinite (Appendix 7-8f) indicating that
the kaolinite was formed during early diagenesis. The characteristic vermicular
texture of kaolinite indicates an origin during early diagenesis (cf. Wilkinson et
al., 2004; Abouessa and Morad, 2009). Chlorite grain-coatings are also
observed to form during early to mid diagenesis (Appendix 7-9e, 7-10f and 711b, c). Kaolinite is coated by chlorite in some samples (e.g. Appendix 7-9e).
This indicates that kaolinite pre-dates chlorite in the Hawkesbury Sandstone.
Chlorite and illite formation can be derived from earlier kaolinite (cf. Kim et al.,
2007).
Dissolution of feldspar can control kaolinite precipitation (cf. Rossi et al., 2001).
Ions may become available for kaolinite authigenesis by dissolution and
overgrowths of feldspar (Wolela and Gierlowski-Kordesch, 2007). Meisler et al.
(1984), Morad et al. (2000) and Ketzer et al. (2003a, b) and El-Ghali et al.,
2006b attributed the formation of kaolinite and the associated unstable
framework grain dissolution to near surface and meteoric water diagenesis.
Wolela and Gierlowski-Kordesch (2007) indicated that the reactions that occur
between the unstable grains, carbonate cements and acidic pore water play a
role in increasing kaolinite precipitation as alkalinity increases. Low pH and low
ionic strength waters are recorded as supportive factors to the precipitation of
kaolinite (Bjorlykke et al., 1989, 1986; Worden and Burley, 2003; Hammer et al.,
2010).
The conversion of kaolinite into dickite is indicated by the presence of
vermicular stacks and booklets with the thick to blocky habits of dickite
(Ehrenberg et al., 1993; McAulay et al., 1994; Morad et al., 1994, Lanson et al.,
2002 and Abouessa and Morad, 2009). According to Morad et al. (1994) and
210

Abouessa and Morad (2009), dissolution and re-precipitation can play an active
role in the conversion of kaolinite into dickite which characteristically occurs at a
temperature >100C (Ehrenberg et al., 1993; McAulay et al., 1994; Morad et al.,
1994, Lanson et al., 2002 and Abouessa and Morad, 2009).
Illite grain-coatings are also characterised as early to mid authigenic minerals.
Illite coats earlier authigenic kaolinite and quartz grain margins (Appendix 7-8c,
7-9a-b, d and 7-10e). High activities of K+ and H4SiO4 support the precipitation
of grain-coating illite (Ali and Turner, 1982). Identification of the relative timing of
authigenic illite is difficult because the textural relationships between illite and
other authigenic minerals were not observed (cf. Sur et al., 2002). The
authigenic mineral assemblage and original sandstone composition indicate the
timing of authigenic illite to be during early diagenesis.
6-4-3-2-2 Carbonate cement
Carbonate cements, represented by siderite, ankerite, calcite and dolomite,
formed during early diagenesis in the succession. The presence of graincoating and pore-filling siderite supports this interpretation. The margins of
some detrital quartz grains are coated by carbonate cement as seen in thin
section (e.g. Appendix 6-34b-c, 6-35a-c, 6-36a-b and 6-38b). Large porespaces which are available for the development of euhedral siderite crystals
provide enough evidence for the deposition of siderite during early diagenesis
(Appendix 6-35b; cf. Lee and Lim, 2008). Minor ferric oxide is formed by
oxidation of the siderite (Standard, 1969).
Abouessa and Morad (2009) showed that siderite precipitated at high
temperatures with high Fe and low Ca contents. Also, high Fe/Ca supports the
interpretation of the precipitation of siderite cement within the early stage of
diagenesis (Berner, 1981; Lee and Lim, 2008).
6-4-3-2-3 Dissolution/alteration of detrital grains
Dissolution of detrital grains and authigenic minerals is recorded in the
Hawkesbury Sandstone. The precipitation of secondary silica was accompanied
by, and most probably succeeded, partial or complete dissolution of carbonate
cement (Appendix 6-39a). The presence of silica overgrowths maintained the
211

framework, so that the oversized pores left after the dissolution of the carbonate
fragments were preserved (Appendix 6-39b). In general, secondary porosity
was formed by dissolution in the Hawkesbury Sandstone (Appendix 6-38c and
6-39a).
Some detrital quartz grains were dissolved in the Hawkesbury Sandstone (e.g.
Appendix 6-34c). Dissolution post-dates compaction but pre-dates quartz
cementation. Leaching of the carbonate cement is observed because of fresh
water influx (Appendix 6-34c).
Authigenic feldspar is recorded during mid to late diagenesis in the Hawkesbury
Sandstone. Precipitation of feldspar is supported by the activities of K+, Al2+ and
Si4+ (Wolela, 2009). Feldspar dissolution was also determined in the
Hawkesbury Sandstone (Appendix 7-11e).
6-4-3-2-4 Authigenic quartz
Quartz overgrowths occur as mid to late diagenetic minerals. The SEM studies
indicated mid to late stage secondary quartz overgrowths (Appendix 7-8e, 710a-b, d-e and 7-11a, d). The occurrence of quartz cement in the form of
overgrowths supports this interpretation. Some authigenic clay grew before the
quartz overgrowths and this indicates the late stage of quartz cementation
(Appendix

6-38a).

Quartz

overgrowth

deposition

continued

after

the

precipitation of clays and this is indicated by the syntaxial growth of quartz


cement on grains with thin chlorite coatings. Chlorite-clay formation pre-dates
quartz overgrowths in all places (cf. Smosna and Sager, 2008), except where
chlorite coatings occur on quartz overgrowths.
Late diagenetic quartz overgrowth were described by Al-Gailani (1981) who
showed that quartz overgrowth is attributed to acidic waters percolating into the
subsurface and containing dissolved silica and alumina. The mixing of the fresh
acidic waters with the more alkaline indigenous waters would decrease the
acidity of the percolating waters, leading to a decrease in the solubility of silica
and alumina and hence precipitation of quartz overgrowths.
In some samples, secondary silica is precipitated around detrital quartz grains
(Appendix 6-34a and 7-8e). The pore waters required for the precipitation of
212

these overgrowths would have been acidic, in an environment containing


sufficient dissolved silica to allow quartz overgrowth formation. Compaction is
prevented because of quartz overgrowths, thus some primary intergranular
porosity is preserved.
The widespread quartz cementation indicates an initial clay-poor quartz-rich
sediment with a high permeability that allowed silica-rich solution circulation to
the necessary sites for the nucleation of quartz overgrowths (cf. Franklin and
Tieh, 1989; De Ros et al., 1994; Wolela, 2009). Quartz overgrowths are
common in coarse-grained clean sandstone because its high permeability
allowed easy percolation of water and silica-rich fluids (Al-Harbi and Khan,
2008).
Al-Harbi and Khan (2008) interpreted that dissolution of quartz grains,
replacement of feldspar and quartz grains by carbonate, alteration of smectite to
illite, pressure dissolution, alteration and dissolution of volcanic rock fragments
are the sources of quartz cement. The grain contact types, such as concavoconvex contacts and microstylolites, show that pressure dissolution of quartz
grains is a major source of the silica in the Hawkesbury Sandstone (cf. AlRamadan et al., 2004). The feature of syntaxial overgrowths of quartz cement
and the association of quartz cement with sites of intergranular dissolution
indicate a mesogenetic origin (Worden and Morad, 2000; Morad et al., 2000; ElGhali et al., 2006b). In the Hawkesbury Sandstone, pressure solution of quartz
at grain contacts forming stylolites, solution of fine quartz particles and
decomposition of silicate minerals may be the sources of quartz overgrowths
(Standard, 1964). Obsorne (1948) suggested that regional cementation by
secondary silica in the Hawkesbury Sandstone is because of groundwater
action.
6-4-3-2-5 Late authigenic kaolinite, illite and chlorite
Authigenic kaolinite occurs as a late stage diagenetic mineral that was
precipitated after the quartz overgrowths, as indicated by SEM studies. In most
samples, the late phase of kaolinite cement was precipitated on the quartz
overgrowth (Appendix 7-8e and 7-10a-b). This indicates that kaolinite postdates the quartz overgrowths. Significant intercrystalline microporosity is
213

preserved in the authigenic kaolinite. Also, illite and chlorite were precipitated
on the quartz overgrowth in other samples, interpreting that they are late
diagenetic stage (Appendix 7-8d, 7-10d and 7-11a).
6-4-3-2-6 Late siderite and ankerite cement
Carbonate cementation, particularly siderite and ankerite, also occurred during
the late stages of diagenesis. Siderite and ankerite cement exist as graincoatings on quartz overgrowths. This indicates that they post-date the quartz
overgrowths (Appendix 6-36b and 6-38b). Euhedral crystals of siderite are
uncommon in some samples but confirm a late phase origin. Also, later
compaction is prevented by carbonate cementation.
Mg and Ca are more abundant in late diagenetic siderite (Curtis and Coleman,
1986; Rossi et al., 2001). New siderite cement may be re-precipitated in some
samples after dissolution of early carbonate cement (Appendix 6-39a). This may
be the interpretation of the occurrence of late diagenetic siderite cement.
6-4-3-2-7 Ferruginous Cement
Ferruginous cement is recorded as a late diagenetic mineral in the Hawkesbury
Sandstone. The iron-oxide does not form an oxidized rim on detrital grains, but
its late diagenetic origin is indicated by the lack coatings at grain contacts (cf.
Sur et al., 2002).
Iron-rich materials are the source for the ferruginous cement in the Hawkesbury
Sandstone. Iron may develop as hematite in the oxidizing diagenetic
environment. The iron oxide cement may be derived from dissolution or
alteration of iron silicates, oxidation of ilmenite and magnetite, and iron released
by the replacement of siderite cement (cf. Hubert et al., 1976; Wolela, 2009).
Bloch et al. (2002) showed that iron is derived from a meteoric-water source,
indicated by the presence of chlorite and siderite cement. Also, Umar et al
(2011) indicated that alteration of Ti-Fe oxides and volcanic fragments are
important in the liberation of iron and this process represents a source of iron
oxide cements. Also, according to Wolela (2009), enrichment of the pore space
chemistry was important in the distribution of hematite (Wolela, 2009).
In conclusion, the chronological order of the diagenetic phases is as follows;
214

1) Mild compaction
2) Authigenic clay minerals
3) Carbonate cementation
4) Dissolution/alterations of detrital grains
5) Secondary silica overgrowths
6) Precipitation of authigenic kaolinite followed by siderite and ankerite
cement
7) Ferruginous cement.

Diagenetic minerals

Early

Late

Compaction
Quartz overgrowth
Kaolinite
Dickite
Mixed-layer illite/smectite
Illite
Chlorite
Siderite
Ankerite
Calcite
Dolomite
Dissolution/alteration
Iron oxide

Fig. 6-9: Paragenetic sequence of the diagenetic alterations in the Hawkesbury


Sandstone.

215

216

Chapter Seven
Porosity and permeability
7-1 Introduction
Sedimentary rocks are characterised by porosity and permeability. The two
types of porosity are primary and secondary porosity. Primary porosity consists
of inter- and intraparticle porosity and is observed between detrital grains
whereas secondary porosity occurs during diagenesis as result of dissolution of
unstable grains (Worden and Burley, 2003). Intergranular porosity exists
between grains (Choquette and Pray, 1970; Worden and Burley, 2003) whereas
intragranular porosity is observed mostly within grains (Worden and Burley,
2003). Environments, depositional energy and grain size can control primary
porosity in sandstones.
Many studies have showed the relationship between reservoir quality and
diagenetic

alteration

and

between

reservoir

quality

and

depositional

environments. These previous studies have been discussed in chapter two. For
example, El-ghali et al. (2006b) studied diagenetic alteration and its influence
on reservoir quality in the Murzuq Basin, SW Libya. Also, Zhang (2008)
described the influence of depositional facies and diagenesis on the porosity
and permeability in Silurian sandstones from the Tazhong area in the central
Tarim Basin, western China. Also, Behrenbruch and Biniwale (2005) studied
hydraulic unit analysis of clastic depositional environments and rock types in
Australian sedimentary basins.
Previous studies of porosity and permeability in the Illawarra Coal Measures are
rare to absent. Herbert, (1987), Bai (1991), Heritage (2005) and Ryan (2005)
introduced a study of porosity and permeability in the Narrabeen Group.
Porosity and permeability in the Hawkesbury Sandstone have been analysed by
Griffith (1986), Liu et al., (1996), Lee (2000), Emerson (2000), Franklin (2000),
McKibbin and Smith (2000), Freed (2005) and Gentz (2006).

712

In this chapter, sandstone porosity is described by thin section, well log and
scanning electron microscope (SEM) analysis in the southern Sydney Basin.
Helium porosity and air permeability were determined by measurements.

7-2 Porosity and permeability data


7-2-1 Illawarra Coal Measures
Thin section and scanning electron microscope (SEM) showed that porosity is
comprised of both primary and secondary porosity in the southern Sydney
Basin. The determination of microporosity by microscope is not easy in most of
the studied samples (El-ghali et al., 2009a).
7-2-1-1 Thin section porosity
In the Illawarra Coal Measures, total thin section porosity varies between a
trace and 15% at average of 1.3% in the sandstone and 0.04% in the siltstone
(Appendix 5-7 and 5-10). Thus, it is very low in most samples, with the Kembla
Sandstone, Lawrence Sandstone and Loddon Sandstone having averages of
1.1%, 1.4% and 2.6% (Appendix 5-13, 5-14 and 8-6), respectively. Secondary
porosity exceeds primary porosity in these units. In general, no consistent
correlation is observed between thin section porosity and detrital or authigenic
grains such as quartz, feldspar, lithic grains, quartz overgrowth and carbonate
cement in the Illawarra Coal Measures (Figs 7-1a-b, 7-2a-b and 7-3a). In the
Kembla Sandstone, low negative correlation is recorded between thin section
porosity and quartz grains (r2=0.2; Fig. 7-3b).
7-2-1-2 Well log density porosity
In the Illawarra Coal Measures, density porosity results from well logs were
varied and ranged between 0.1% and 22.2% at average of 5.7% in the
sandstone and 8% in the siltstone (Table 7-1 and 7-2). Density porosity
decreases from 15.7% in the Wilton Formation to 5.3% in the Darkes Forest
Sandstone. Up-sequence, it increases to its highest value at 22.2% in the Allans
Creek Formation but then decreases to 0.6% in the Unnamed Two Member
(Table 7-3). In the Lawrence Sandstone and Loddon Sandstone, it is 5% and
4.3%, respectively (Table 7-3).
712

a)
16

y = -0.0681x + 2.2343
R = 0.0343

14

Porosity (%)

12
10
8
6
4
2
0
0

b)

10

20

30
Quartz (%)

40

50

60

16
y = 0.0711x + 1.1408
R = 0.0004

14

Porosity (%)

12
10
8
6
4
2
0
0

0.5

1.5
2
Feldspar (%)

2.5

3.5

Figure 7-1: a) and b) No significant correlation between


quartz and porosity and between feldspar and porosity in the
Illawarra Coal Measures.

711

a)
16

y = 0.0274x + 0.0947
R = 0.046

14

Porosity (%)

12
10
8
6
4
2
0
0

20

40

60

80

100

Lithic grains (%)

b)
16
y = -0.0582x + 1.2785
R = 6E-05

14

Porosity (%)

12
10
8
6
4
2
0
0

0.5

1
1.5
Quartz overgrowth (%)

2.5

Figure 7-2: a) and b) No significant correlation between lithic


grains and porosity and between quartz overgrowth and
porosity in the Illawarra Coal Measures.

772

a)
16

y = -0.0228x + 1.5869
R = 0.0151

14

Porosity (%)

12
10
8
6
4
2
0
0

b)

10

20

30
40
50
Carbonate cement (%)

60

70

y = -0.1085x + 2.3391
R = 0.1762

4.5
4
Porosity (%)

3.5
3
2.5
2
1.5
1
0.5
0
0

10

15

20

25

Quartz (%)

Figure 7-3: a) No significant correlation between carbonate cement


and porosity in the Illawarra Coal Measures, b) Low negative
correlation is recorded between quartz and porosity in the Kembla
Sandstone.

771

Table 7-1: Values of thin section porosity, helium porosity, density porosity and
permeability to air for twenty one samples from the Illawarra Coal Measures,
Narrabeen Group and Hawkesbury Sandstone.
Thin
Helium Density Permeability
section
porosity porosity to air (Md)
porosity
13
16.1
16.6
1.1
9.5
11.2
17.1
0.2
4.2
15
12.4
22.3
2.5
14.9
5.9
0.2
1.5
3.2
9.8
0.03
11.9
18.1
0
79.4
16.9
14.7
0
2.1
10.2
15.3
0.7
16.1
3
9.9
1.3
0.03
4
7.5
1.4
0.04
5.9
6.8
0
0.05
6
11.7
9.2
0.1
9.2
13.5
8
0.1
0
5.1
3.5
0.01
0
5.2
0
0.01
2
6.9
4.4
0.01
3.7
7.5
2.3
0.01
0
4
4
0.01
3.7
8.1
7.4
0.01
4.7
8
7.2
0.01
0
3.5
7.4
0.01

Samples
HBSS - EAW 18 a (5)
HBSS - EDEN 115 (4)
HBSS - EDEN 115 (13)
NPFM - EAW 42 (2)
BACS - EAW 42 (14)
BGSS - EAW 42 (18)
BGSS - EAW 42 (25)
BGSS - EAW 42 (32)
SPCS - EAW 42 (34)
SBSS - EAW 42 (42)
SBSS - EAW 42 (45)
WBCS - EDEN 124 (5)
WBCS - EDEN 124 (6)
CCSS - EDEN 124 (9)
CCSS - EDEN 124 (10)
LDSS - EDEN 124 (14)
LDSS - EDEN 124 (15)
LRSS - EDEN 124 (19)
KBSS - EDEN 124 (24)
KBSS - EDEN 124 (25)
DFSS - EDEN 124 (30)

Table 7-2: Average density porosity in sandstone and siltstone in the


Illawarra Coal Measures, Narrabeen Group and Hawkesbury Sandstone.
Samples

ICM

NBG

HBSS

Sandstone

5.7

5.1

14.5

Siltstone

3.9

777

Table 7-3: Average density porosity, and average primary, secondary and
total porosity plus equivalent helium porosity in the Illawarra Coal
Measures, Narrabeen Group and Hawkesbury Sandstone.
FM
HBSS
NPFM
BACS
BGSS
SPCS
SBSS
WBCS
CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS
WTFM

N
20
2

DNPO
13.5
1.8

10
1
14
38
12
19
13
1
3

7.8
0.9
4
5.4
3.7
4.3
5
0.6
4.6

13
1
7
1
2

6.7
22.2
5.3
11.2
15.7

N
32
3
1
12
2
19
43
20
19
18
1
3
1
19
4
7
1
3

P1
6.7
0.2
0
1.9
1.3
0.7
0.5
0.5
0.3
0.3
0
0
0
0.1
0
0
0
0

P2
1.5
0.7
1.5
10
5.5
5.8
3
4
2.2
1.1
0
0
0
1
0
0
0
0

Tpo
8.2
0.8
1.5
12
6.8
6.5
3.5
4.5
2.6
1.4
0
0
0
1.1
0
0
0
0

EHP
12.8
3.6
4.4
17.4
11
10.6
7
8.2
5.8
4.3
2.6
2.6
2.6
3.9
2.6
2.6
2.6
2.6

7-2-1-3 Helium porosity and permeability to air


Helium porosity and permeability to air were measured for 6 samples from the
Darkes Forest Sandstone, Kembla Sandstone, Lawrence Sandstone and
Loddon Sandstone. One sample from the Darkes Forest Sandstone is fine
sandstone and is moderately well sorted. Two samples were selected from the
Kembla Sandstone and are coarse sandstone. They are both moderately well
sorted and are rich in lithic grains. One sample was selected from the Lawrence
Sandstone and is fine sandstone, moderately well sorted. Two samples were
selected from the Loddon Sandstone and are coarse sandstone, moderately
well sorted and rich in lithic grains. Helium porosity varies between 3.5% and
8.1% in these units (Table 7-1). Also, the results showed that permeability to air
is less than 0.01 md in all these units (Table 7-1).
7-2-1-4 Comparison between thin section porosity and density porosity
The comparison between thin section porosity and density porosity is similar in
the range where it does not exceed 15% in thin section and 22.2% in well logs
772

(Appendix 5-24). Density porosity is more common in siltstone than in


sandstone (Table 7-2) whereas thin section porosity is more common in
sandstone than in siltstone (Appendix 5-10). Density porosity is recorded in
most samples, whereas thin section porosity varies in some samples. Also, no
significant correlation is observed between thin section porosity and density
porosity (Fig. 7-4). This relationship indicates that density porosity tends to
increase weakly whereas thin section porosity is stable. This may be because
the conventional logs are controlled by clays, thus density porosity values are
higher in comparison with thin section porosity.

y = -0.1252x + 6.0158
R = 0.0072

25.0

Density porosity (%)

20.0

15.0

10.0

5.0

0.0
0

10
Thin section porosity (%)

15

20

Figure 7-4: No significant correlation between thin section porosity


and density porosity in the Illawarra Coal Measures.
7-2-1-5 Comparison between thin section porosity and helium porosity
A comparison between thin section porosity and helium porosity could only be
carried out for six samples from the Darkes Forest Sandstone, Kembla
Sandstone, Lawrence Sandstone and Loddon Sandstone. The results indicate
that helium porosity is more common than thin section porosity in these
samples (Table 7-1). This may result from the inclusion of microporosity in the
helium porosity. In contrast, thin section porosity did not include microporosity
during this study. For example, sample EDEN 124 (14) from the Loddon
772

Sandstone has 2% porosity in the thin section but shows 6.9% helium porosity
(Table 7-1).
7-2-1-6 Comparison between density porosity and helium porosity
The comparison between density porosity and helium porosity was done for the
same six samples from the Dark Forest Sandstone, Kembla Sandstone,
Lawrence Sandstone and Loddon Sandstone. The results showed that helium
porosity is similar with density porosity in samples (Table 7-1).

7-2-2 Narrabeen Group


In previous studies, Herbert (1987) showed that the Coalcliff Sandstone and
Scarborough Sandstone are characterised by low permeability, which varies
between 1 and 10 md. In contrast, the basal Bulgo Sandstone and basal
Newport Formation include higher permeability sandstone. Bai (1991) recorded
that the primary and secondary porosity does not exceed 17% and 11%,
respectively, in the Narrabeen Group. In another study, porosity was recorded
at 5% in the Gosford Sandstone (Ord et al., 1991). Ryan (2005) showed that
porosity and permeability is the least common in the Coalcliff Sandstone and
Wombarra Claystone whereas the highest porosity and permeability were
recorded in the Scarborough Sandstone. He used the helium injection method
and a Hassler Holder at 250 psi to measure porosity and permeability,
respectively. Heritage (2005) also determined porosity in the Narrabeen Group,
describing both thin section porosity and density porosity and compared them
both.
7-2-2-1 Thin section porosity
In the present study, porosity ranges from 0 to 19.5% at average of 5.4% in the
sandstone whereas it is absent in the siltstone samples (Appendix 5-8 and 511). Porosity was recorded in all units of the Narrabeen Group. It is present in
the Coalcliff Sandstone (4.5%) and Wombarra Claystone (3.5%) and then
increases to 12% in the Bulgo Sandstone but then decreases to 0.8% in the
Newport Formation (Appendix 5-13, 5-14 and 8-6). Secondary porosity is more
common than primary porosity in all units of the Narrabeen Group (Appendix 5772

13 and 5-14). There is no correlation between thin section porosity and quartz,
feldspar, lithic grains, quartz overgrowth and carbonate cement in the
Narrabeen Group (Figs 7-5a-b, 7-6a-b and 7-7a). In the Scarborough
Sandstone, quartz overgrowths are moderately positively correlated with thin
section porosity (r2=0.4; Fig. 7-7b). However, no correlation is observed
between quartz overgrowths and porosity in the Bulgo Sandstone (Fig. 7-8a).
Also, carbonate cement shows a moderately negative correlation with thin
section porosity in the Scarborough Sandstone (r2=0.3; Fig. 7-8b) whereas it is
weakly negatively correlated with thin section porosity in the Bulgo Sandstone
(r2=0.2; Fig. 7-9).
a)

25

y = 0.0731x + 2.8997
R = 0.0514

Porosity (%)

20
15
10
5
0
0

20

40

60

80

100

Quartz (%)

b)

y = 1.1074x + 4.3101
R = 0.0201

25

Porosity (%)

20
15
10
5
0
0

Feldspar (%)

Figure 7-5: a) and b) No significant correlation between quartz


and porosity and between feldspar and porosity in the
Narrabeen Group.
772

a)

25
y = -0.0211x + 5.9346
R = 0.0048

Porosity (%)

20

15

10

0
0

20

40
60
Lithic grains (%)

80

100

b)
y = 1.31x + 3.0494
R = 0.1036
25

Porosity (%)

20
15
10
5
0
0

Quartz overgrowth (%)

Figure 7-6: a) and b) No significant correlation between lithic


grains and porosity and between quartz overgrowths and
porosity in the Narrabeen Group.

772

a)

y = -0.1779x + 7.2437
R = 0.0822

25

Porosity (%)

20

15

10

0
0

b)

10

20

30
40
50
Carbonate cement (%)

60

70

y = 5.7391x - 4.8746
R = 0.4427

25

Porosity (%)

20

15

10

0
0

2
Quartz overgrowth (%)

Figure 7-7: a) No significant correlation between carbonate


cement and porosity in the Narrabeen Group. b) A moderate
positive correlation between quartz overgrowths and porosity in
the Scarborough Sandstone.

772

a)

y = 0.0155x + 11.906
R = 5E-05

20
18
16

Porosity (%)

14
12
10
8
6
4
2
0
0

4
Quartz overgrowth (%)

b)

y = -0.5074x + 12.259
R = 0.3308

25

poorosity (%)

20

15

10

0
0

10

15

20

25

30

Carbonate cement (%)

Figure 7-8: a) No correlation between quartz overgrowth and


porosity in the Bulgo Sandstone, b) Carbonate cement is
moderately negatively correlated with thin section porosity in
the Scarborough Sandstone.

771

y = -0.5943x + 14.905
R = 0.2065

20
18
16
Porosity (%)

14
12
10
8
6
4
2
0
0

6
8
Carbonate cement (%)

10

12

14

Figure 7-9: Carbonate cement is weakly negatively correlated


with thin section porosity in the Bulgo Sandstone.

722

7-2-2-2 Well log density porosity


Density porosity ranges from 0.1% to 15.8% at average of 5.1% in the
sandstone and 4% in the siltstone (Tables 7-1 and 7-2). It is present at 3.7%
and 5.4% in the Coalcliff Sandstone and Wombarra Claystone, respectively
(Tables 7-3). Up-sequence, it decreases to 0.9% in the Stanwell Park
Claystone. In the Bulgo Sandstone and Newport Formation it reaches to 7.8%
and 1.8%, respectively (Tables 7-3).
7-2-2-3 Helium porosity and permeability to air
Helium porosity and permeability to air were measured for 12 samples from the
Coalcliff Sandstone, Wombarra Claystone, Scarborough Sandstone, Stanwell
Park Claystone, Bulgo Sandstone, Bald Hill Claystone and Newport Formation.
Two samples were selected from the Coalcliff Sandstone and are fine- and
medium-grained sandstone, respectively. They are moderately well sorted.
Also, two coarse-grained sandstone samples were selected from the Wombarra
Claystone and have common lithic grains. They are moderately well sorted and
moderately sorted, respectively. Two samples were selected from the
Scarborough Sandstone and are rich in lithic grains. They are medium- and
coarse-grained sandstone, respectively and are moderately well sorted and
moderately sorted, respectively. One sample was selected from the Stanwell
Park Claystone and is medium-grained sandstone, moderately well sorted.
Three samples were selected from the Bulgo Sandstone and are coarsegrained sandstone. They are moderately well sorted and have more detrital
quartz than lithic grains. One sample was selected from the Bald Hill Claystone
and is coarse-grained sandstone, moderately well sorted. A medium-grained
sandstone sample was selected from the Newport Formation and it has a high
percentage of detrital quartz and is moderately well sorted. Helium porosity was
recorded between 3.2% and 18.1% in these units whereas permeability to air
varies in its values between less than 0.01 and 79.4 md (Table 7-3).
7-2-2-4 Comparison between thin section porosity and density porosity
The comparison between thin section porosity and density porosity indicates
that the highest values of porosity are 19.5% and 15.8% in the thin section
porosity and density porosity, respectively (Appendix 5-25). The thin section
721

porosity and density porosity are more common in sandstone than in siltstone at
similar percentages (Table 7-2 and Appendix 5-11). Most samples have thin
section porosity and density porosity. There is very low positive correlation
between thin section porosity and density porosity (r2 = 0.05; Fig. 7-10). In this
case, density porosity tends to increase weakly whereas thin section porosity is
stable. This shows the influence of clays on the conventional logs, and thus
density porosity values are higher than thin section porosity.

y = 0.1648x + 4.0847
R = 0.0541

18
16

Density porosity

14
12
10
8
6
4
2
0
0

10

15

20

25

Thin section porosity (%)

Figure 7-10: No significant correlation between thin section porosity


and density porosity in the Narrabeen Group.
7-2-2-5 Comparison between thin section porosity and helium porosity
Also, the comparison between thin section porosity and helium porosity for only
twelve

samples

from

the

Coalcliff

Sandstone,

Wombarra

Claystone,

Scarborough Sandstone, Stanwell Park Claystone, Bulgo Sandstone, Bald Hill


Claystone and Newport Formation showed a greater abundance of helium
porosity compared with thin section porosity in all these samples, except one
sample (Table 7-1). This is may be because microporosity was included in the
helium porosity whereas it could not be distinguished in thin section.

727

7-2-2-6 The comparison between density porosity and helium porosity


Helium porosity was measured in twelve samples from the Coalcliff Sandstone,
Wombarra Claystone, Scarborough Sandstone, Stanwell Park Claystone, Bulgo
Sandstone, Bald Hill Claystone and Newport Formation. The values were
compared with density porosity values for these samples. The results showed
that helium porosity values are higher than density porosity values in these
samples, except one sample in the Bald Hill Claystone (Table 7-1). This may be
because most sandstone samples include heavy minerals, thus their effective
porosity is not determined accurately by density porosity according Grain (2000)
since heavy minerals are recorded as clay volume. This is why density porosity
is lower than helium porosity.

7-2-3 Hawkesbury Sandstone


In previous studies, Mayne et al. (1974) calculated the percentage of porosity in
the Hawkesbury Sandstone to be between 8 and 18%, whereas Griffith (1986)
calculated 5 to 10% porosity in the massive channel sands. Liu et al. (1996)
indicated that porosity and permeability in sandstone do not exceed 20% and
1200 md respectively. Franklin (2000) measured porosity in both the weathered
and unweathered sandstone. It is 5% in the former whereas it increases to10%
in the latter. Freed (2005) indicated that porosity in sandstone does not exceed
9.1% whereas Gentz (2006) determined porosity which occurs at average of
6.6% in the Hawkesbury Sandstone.
7-2-3-1 Thin section porosity
In the present study, porosity is mostly primary and is observed between 0 and
19.3% at average of 9% in the sandstone whereas it disappears in siltstone
samples (Appendix 5-9 and 5-12). During burial, the removal of flow conduits
through compaction was partly effective in this unit which is characterized by
primary porosity (Reed et al., 2005). No correlation was observed between
detrital grains such quartz and feldspar with thin section porosity (Fig. 7-11a-b).
Also, there is no correlation between quartz overgrowths and thin section
porosity (Fig. 7-12a) whereas carbonate cement has a low negative correlation
with thin section porosity in the Hawkesbury Sandstone (r2=0.2; Fig. 7-12b).
722

Also, there is no correlation between quartz overgrowths and secondary


porosity in this unit (Fig. 7-13).
7-2-3-2 Well log density porosity
In the Hawkesbury Sandstone, density porosity is up to 18.3% and has a higher
value in sandstone (14.5%) than in siltstone (3.9%).
a)

y = 0.084x + 3.0266
R = 0.0224

25

Porosity (%)

20

15

10

0
0

b)

20

40

Quartz (%)

60

80

100

y = 7.4284x + 7.1924
R = 0.0389

25

Porosity (%)

20

15

10

0
0

0.1

0.2

0.3
Feldspar (%)

0.4

0.5

0.6

Figure 7-11: a) and b) No significant correlation between


quartz and porosity and between feldspar and porosity in the
Hawkesbury Sandstone.

722

a)

y = 0.8314x + 2.9141
R = 0.0483

25

Porosity (%)

20

15

10

0
0

10

Quartz Overgrowth (%)

b)

y = -0.5722x + 10.451
R = 0.2073

25

Porosity (%)

20

15

10

0
0

10

15

20

25

Carbonate cement (%)

Figure 7-12: a) No significant correlation between quartz


overgrowth and porosity in the Hawkesbury Sandstone, b)
There is a low negative correlation between carbonate cement
and porosity in the Hawkesbury Sandstone.

722

y = 0.1133x + 0.7936
R = 0.0225

5
4.5
Secondary porosity (%)

4
3.5
3
2.5
2
1.5
1
0.5
0
0

4
6
Quartz overgrowth (%)

10

Figure 7-13: No significant correlation between quartz


overgrowth and secondary porosity in the Hawkesbury
Sandstone.

722

7-2-3-3 Helium porosity and permeability to air


Helium porosity and permeability to air were measured for three samples from
the Hawkesbury Sandstone. One sample is medium- to coarse-grained
sandstone whereas two samples are medium-grained sandstone. These
samples are rich in detrital quartz and are moderately well sorted. Helium
porosity was recorded at percentages of 11.2%, 15% and 16.1% whereas
permeability to air was measured at values of 0.2, 22.3 and 1.1 md (Table 7-1).
7-2-3-4 Comparison between thin section porosity and density porosity
Most samples show both thin section and density porosity and the porosity is
greater in the sandstone than in the siltstone samples (Table 7-2; Appendix 512). Both porosity types show a similar range which does not exceed 19.3%
and 18.3% in thin section porosity and density porosity, respectively (Appendix
5-26). A low positive correlation is recorded between thin section porosity and
density porosity (r2 = 0.2; Fig. 7-14).

y = 0.2706x + 11.156
R = 0.1726

20.0
18.0

Density porosity

16.0
14.0
12.0
10.0
8.0
6.0
4.0
2.0
0.0
0

10
Thin section porosity (%)

15

20

Figure 7-14: Low positive correlation occurs between thin section


porosity and density porosity in the Hawkesbury Sandstone.

7-2-3-5 Comparison between thin section porosity and helium porosity


Above we showed that helium porosity includes values of 16.1%, 11.2% and
15% in three samples. These values are not different from thin section porosity
722

values in first two samples which are 13% and 9.5% whereas the third sample
has thin section porosity of 4.2% which is largely different from helium porosity
value (Table 7-1). The third sample may include high microporosity which was
included in the helium porosity but was not included in the thin section porosity.
7-2-3-6 Comparison between density porosity and helium porosity
No large difference is observed between density porosity and helium porosity in
three samples. The three samples have 16.6%, 12.4% and 17.1% of density
porosity and 16.1%, 15% and 11.2% of helium porosity, respectively (Table 71).

7-3 The interpretation of the difference between thin


section porosity, helium porosity and density porosity
There are many reasons support that helium porosity is more abundant than
thin section porosity in most samples. In the study of helium porosity,
microporosity was measured during core analysis whereas it was not measured
in thin section analysis since the image resolution is not accurate or because
the micropores have a size which is less than the thickness of the thin section.
Also, sorting, grain size distribution, orientation of grains and clasts, the
occurrence of horizontal laminae and drapes of glauconite and mudstone are
recorded as three-dimensional effects which all affect point count results (Anon,
1997). Also, helium was used to measure porosity by gas pressure, thus it
usually gives higher values because the gas under pressure can penetrate into
the pores which are often are very fine. Other fluids such as water and oil
cannot transit these pores. Thus, these pores are not seen by microscope
analysis and are not included in the calculation of porosity. This is because
helium gas has a very small size. The gas expansion method is more accurate
than petrographic methods.
In most samples, thin section porosity values are less than density porosity
values which are affected by clay content. This clay has an influence on
resistivity and porosity logs, thus the former is too low whereas the latter is too
high. This indicates that conventional logging tools are affected by clay. Also,
722

density porosity is affected by various factors, including: the chosen density


which may change during the reading; i.e. they use 2.65 g/cm3 in pure
sandstone. The actual density may be less because of the presence of minerals
which are less dense than quartz, such as clay or feldspar. This means that log
porosity reading is greater than the actual value, but if the sandstone was
associated with limestone, dolomite or even biotite, they will give values less
than the actual porosity. It also depends on the existing fluid and free gas (if
any).
Most sandstone samples in the studied succession are characterized by heavy
minerals. These minerals are converted to clay volume and hence reduce the
density porosity according to (Grain, 2000). This supports that thin section
porosity values are higher than density log porosity in some samples.

7-4 Correlation of helium porosity and air permeability


with detrital and authigenic grains
Twenty one samples were selected from the Southern Sydney Basin to study
helium porosity and air permeability. These results were correlated with detrital
and authigenic grains. There is a high positive correlation between detrital
quartz grains and helium porosity (r2=0.6; Fig. 7-15a) whereas matrix exhibits a
highly negatively correlated with helium porosity (r2=0.6; Fig. 7-15b). However,
feldspar (r2=0.3), lithic grains (r2=0.4), muscovite (r2=0.4) and carbonate cement
(r2=0.4; Figs 7-16a-b and 7-17a-b) are relatively moderately negatively
correlated with helium porosity whereas quartz overgrowths are moderately
positively correlated with helium porosity (r2=0.3; Fig. 7-18a) in the Southern
Sydney Basin.
Also, there is a low positive correlation between quartz grains and air
permeability (r2=0.2; Fig. 7-18b). Lithic grains have low negatively correlated
with air permeability (r2=0.2; Fig. 7-19). No any correlation is clear between
cementation such as quartz and carbonate with air permeability.

721

a)

y = 0.1349x + 5.2981
R = 0.5984

20
18

Helium porosity

16
14
12
10
8
6
4
2
0
0

20

40

60

80

100

Quartz (%)

b)
y = -0.445x + 15.805
R = 0.5882

20
18

Helium porosity

16
14
12
10
8
6
4
2
0
0

10

15

20

25

30

35

Matrix (%)

Figure 7-15: a) Good positive correlation is observed between


quartz and helium porosity, b) Good negative correlation is observed
between matrix and helium porosity.

722

a)

y = -3.08x + 12.943
R = 0.2701

20
18

Helium porosity

16
14
12
10
8
6
4
2
0
0

0.5

1.5

2.5

Feldspar (%)

b)

y = -0.1292x + 14.024
R = 0.3693
20
18

Helium porosity

16
14
12
10
8
6
4
2
0
0

10

20

30
40
Lithic grains (%)

50

60

70

Figure 7-16: a) and b) Feldspar and lithic grains are moderately


negatively correlated with helium porosity.

721

a)
20

y = -5.0132x + 11.299
R = 0.3131

18

Helium porosity

16
14
12
10
8
6
4
2
0
0

0.5

1.5

Muscovite (%)

b)

y = -0.3376x + 13.275
R = 0.4008

20
18

Helium porosity

16
14
12
10
8
6
4
2
0
0

10

20

30

40

50

Carbonate cement (%)

Figure 7-17: a) and b) Muscovite and carbonate cement are


moderately negatively correlated with helium porosity.

727

a)

y = 1.0216x + 7.3915
R = 0.3287

20
18

Helium porosity

16
14
12
10
8
6
4
2
0
0

Quartz overgrowth (%)

b)
y = 0.3341x - 5.4013
R = 0.2479

90

Permeability to air (md)

80
70
60
50
40
30
20
10
0
0

20

40

60

80

100

Quartz (%)

Figure 7-18: a) There is a moderate positive correlation


between quartz overgrowths and helium porosity, b) Quartz is
weakly positively correlated with permeability to air.

722

y = -0.3483x + 17.131
R = 0.1811

90

Permeability to air (md)

80
70
60
50
40
30
20
10
0
0

10

20

30

40

50

60

70

Lithic grains (%)

Figure 7-19: There is a low negative correlation between lithic


grains and permeability to air.
Helium porosity is highly positively correlated with thin section porosity (r2=0.6;
Fig. 7-20) and shows a low positive correlation with air permeability (r2=0.3; Fig.
7-21a). Also, air permeability has a low positive correlation with thin section
porosity (r2=0.1; Fig. 7-21b). There are no significant correlations for well log
density porosity with helium porosity or with thin section porosity.
y = 0.8039x - 2.5653
R = 0.6109

18
Thin section porosity (%)

16
14
12
10
8
6
4
2
0
0

10

15

20

Helium porosity

Figure 7-20: There is a high positive correlation between


helium porosity and thin section porosity.
722

a)
y = 2.0477x - 14.308
R = 0.2831

90

Permeability to air (md)

80
70
60
50
40
30
20
10
0
0

10

15

20

Helium porosity

b)

y = 0.0956x + 4.7744
R = 0.1279

18
Thin section porosity (%)

16
14
12
10
8
6
4
2
0
0

20

40

60

80

100

Permeability to air (md)

Figure 7-21: a) A moderate positive correlation is observed


between helium porosity and permeability to air, b) There is a
low positive correlation between permeability to air and thin
section porosity.

722

This study provided an equivalent helium porosity for the Illawarra Coal
Measures, Narrabeen Group and Hawkesbury Sandstone as result of the
correlation between helium porosity and thin section porosity (r2=0.6; Fig. 7-20).

7-5 Equivalent helium porosity


The equivalent helium porosity value for Southern Syney Basin samples can be
determined from the correlation between thin section porosity and helium
porosity (Fig. 7-20). This correlation suggested that X = 2.5653+Y/0.8039
X = equivalent helium porosity
Y = any point from thin section porosity
In the Illawarra Coal Measures, the results showed that equivalent helium
porosity ranges from 2.6% to 21.2% (Appendix 5-24). It is constant from the
Wilton Formation to the Allans Creek Formation at average 2.6% and then it
increases to 3.9% in the Kembla Sandstone (Table 7-3). Up-sequence, it
decreases to 2.6% in the Wongawilli Coal, Unnamed Member Three and
Unnamed Member Two and then it increases to 4.3% and 5.8% in the
Lawrence Sandstone and Loddon Sandstone respectively (Table 7-3).
In the Narrabeen Group, equivalent helium porosity varies between 2.6% and
26.8% (Appendix 5-25). It is recorded at averages of 8.2% and 7% in the
Coalcliff Sandstone and Wombarra Claystone, respectively (Table 7-3). Upsequence, it increases to 17.4% in the Bulgo Sandstone and then it decreases
to 3.6% in the Newport Formation (Table 7-3).
In the Hawkesbury Sandstone, equivalent helium porosity ranges from 2.6% to
26.6% and is present at average of 12.8% (Table 7-3; Appendix 5-26).
In the Illawarra Coal Measures, no significant correlation is observed between
thin section porosity (Po) and equivalent helium porosity (EHP) with height
above base (Fig. 7-22). This indicates that thin section porosity (Po) and
equivalent helium porosity (EHP) decrease weakly with height above base.
Also, no significant correlation is observed between density porosity (DNPO)
722

and height above base (Fig. 7-22). This indicates that density porosity (DNPO)
increases weakly with height above base (Fig. 7-22). In the Narrabeen Group,
thin section porosity, equivalent helium porosity and density porosity decrease
weakly with increasing height above base (Fig. 7-23a). Also, in the Hawkesbury
Sandstone, thin section porosity, equivalent helium porosity and density
porosity decrease weakly with increasing height above base (Fig. 7-23b).
This suggests that compaction and cementation affect thin section porosity and
equivalent helium porosity. In the burial history, the early influence of
compaction on porosity is most apparent before burial below 2 km. With
increasing burial below 2 km, the influence of cementation on porosity is
stronger. Total porosity includes 50% of intergranular macroporosity at depths
that do not exceed 3.5 km. Intergranular macroporosity decreases to 16% at
greater depths. In the Narrabeen Group and Hawkesbury Sandstone, density
porosity decreases with height above base because heavy minerals are
included in the clay volume (Grain, 2000).

100

Height above base (m)

90
80
70
60
50

Po

40

DNPO

30

EHP

20
10
0
0

10

15

20

25

Porosity (%)

Figure 7-22: The correlation between thin section porosity (Po),


density porosity (DNPO) and equivalent helium porosity (EHP)
with height above base in the Illawarra Coal Measures.

722

a)
450

Height above base (m)

400
350
300
250

Po

200

DNPO

150

EHP

100
50
0
0

10

15

20

25

30

Porosity (%)

b)
470

Height above base (m)

450
430
Po

410

DNPO
390

EHP

370
350
0

10

15

20

25

30

Porosity (%)

Figure 7-23: a) and b) The correlation between thin section


porosity (Po), density porosity (DNPO) and equivalent helium
porosity (EHP) with height above base in the Narrabeen Group
and Hawkesbury Sandstone, respectively.

722

7-6 Distribution of density porosity in wells


7-6-1 Illawarra Coal Measures
In this part, the study showed the maximum, minimum and average density
porosity values in the nine wells used in this study (Appendix 5-27). The results
indicated that the average maximum density porosity varies between 15% in the
Darkes Forest Sandstone and 24.7% in the Unnamed Two Member (Appendix
5-28). The average minimum density porosity ranges from 0.01% in the Bargo
Claystone and Unnamed Two Member to 0.5% in the Unnamed Three Member
(Appendix 5-28). The average density porosity for each unit fluctuates between
5% and 10.1% from the Wilton Formation to Loddon Sandstone (Appendix 528).

7-6-2 Narrabeen Group


In the Narrabeen Group, the average maximum density porosity ranges from
7.8% in the Garie Formation to 22.6% in the Coalcliff Sandstone (Appendix 528). The average minimum density porosity is zero in two samples from the
Wombarra Claystone and Bulgo Sandstone whereas in other units, it does
exceed 0.2% (Appendix 5-28). The average density porosity increases from
3.8% in the Coalcliff Sandstone to 6.2% in the Scarborough Sandstone,
decreasing to 5.1% in the Stanwell Park Claystone (Appendix 5-28). Upsequence, it fluctuates between 2.9% and 8.3% from the Bulgo Sandstone to
Newport Formation (Appendix 5-28).

7-6-3 Hawkesbury Sandstone


The average maximum, minimum and average of density porosity in the
Hawkesbury Sandstone are 24.3%, 0.003% and 11.9% respectively (Appendix
5-28).

7-7 Pore types


7-7-1 Illawarra Coal Measures
7-7-1-1 Primary porosity
In this unit, primary porosity varies between 42 and 330 m in size and consists
of intergranular porosity (Appendix 7-1a-d). It ranges from 0% to 4% with
721

average of 0.2% in the sandstone samples (Appendix 5-24 and Table 7-4). In
the Kembla Sandstone, Lawrence Sandstone and Loddon Sandstone, it is
observed at averages of 1%, 0.3% and 0.3%, respectively (Table 7-3).
Table 7-4: Average percentage of primary and secondary porosity in the
Illawarra Coal Measures, Narrabeen Group and Hawkesbury Sandstone.
Samples

Coarse Sandstone

20

Medium Sandstone
Fine Sandstone

ICM
P1

P2

0.3

2.4

27

0.3

24

0.02

Siltstone

Total Sandstone

71

NBG
P1

P2

43

1.0

6.1

1.2

39

0.7

0.1

15

0.04

0.2

1.1

97

0.7

HBSS
P1

P2

15

9.7

1.9

4.9

12

5.7

1.7

0.05

4.7

29

7.4

1.7

7-7-1-2 Secondary porosity


The size of secondary porosity is not very different from dissolved grain size
and varies between 20 and 220 m. This means that pore size and grain size
are often similar.
As indicated by Smosna and Sager (2008), the original grain may be inferred by
two characters of secondary porosity in sandstone, including size and shape.
The dissolution of unstable grains, particularly feldspar and volcanic rock
fragments, as well as carbonate cement increases the percentage of secondary
pores in sandstones (Appendix 6-21a-b, 6-40b-c and 6-41a-c). The secondary
porosity varies between 0% and 11% and is more common in sandstone (1.1%)
than in siltstone (0.04%) samples (Appendix 5-25 and Table 7-4). It is recorded
in the Kembla Sandstone at 1%, increasing to 1.1% and 2.2% in the Lawrence
Sandstone and Loddon Sandstone, respectively (Table 7-3).
Four types of secondary porosity were formed in the Illawarra Coal Measures
including intergranular porosity, intragranular porosity, mouldic pores and
oversized pores.
Secondary intergranular porosity is seen in sandstone, occurring between
grains due to dissolution of pore-filling carbonate cement and it is difficult to
distinguish from primary intergranular porosity Appendix (6-20a-b and 6-40a).
722

The dissolved detrital grains, such as feldspar and volcanic rock fragments,
contribute to the formation of intragranular porosity (Appendix 6-21a-b, 6-40b-c
and 6-41a-c; Maraschin et al., 2004). Intergranular porosity ranges from 33 m
to 210 m whereas intragranular porosity ranges from 25 m to 214 m in size.
Also, partially dissolved feldspar is clearly present associated with some
secondary pores and results in mouldic pores in sandstone (Appendix 6-21a
and 6-41b; cf. Lima and De Ros, 2002; Luo et al., 2009). Oversized pores occur
in thin section and are also affected by total dissolution of several detrital grains
in sandstone (Appendix 6-41c; Zhang et al., 2007). Mouldic pores are typically
between 37 m to 204 m in size whereas oversized pores vary in size
between 56 m to 220 m. Corroded traces of feldspar occur in both mouldic
pores and oversized pores. Also, packing inhomogeneity and honeycombed
grains are observed associated with secondary pores.

7-7-2 Narrabeen Group


7-7-2-1 Primary porosity
In the Narrabeen Group, primary porosity reaches 3.7% in sandstone samples
but it is absent in siltstone samples (Appendix 5-25, 6-42a-b and 7-5a-b, f). It
occurs in sandstone at an average of 0.7% (Table 7-4). Primary porosity occurs
in the Coalcliff Sandstone and Wombarra Claystone at an average of 0.5% and
then increases to 1.9% in the Bulgo Sandstone (Table 7-3). Up-sequence, it
disappears in the Bald Hill Claystone but then occurs in the Newport Formation
in trace amounts (Table 7-3). In this unit, the size of primary pores is recorded
between 34 and 344 m.
7-7-2-2 Secondary porosity
Secondary porosity does not exceed 18% in the Narrabeen Group (Appendix 525) at an average of 4.7% in sandstone (Table 7-4). It fluctuates between 0.7%
and 10% from the Coalcliff Sandstone to Newport Formation (Table 7-3). The
highest secondary porosity is recorded in the Bulgo Sandstone (10%) whereas
the lowest secondary porosity is recorded in the Newport Formation (0.7%;
Table 7-3). The size of secondary porosity was determined to range from 24 to
246 m, depending on grain size. This means that pore size and grain size are
often similar.
721

In the Narrabeen Group, intergranular porosity, intragranular porosity, mouldic


pores and oversized pores are also observed. Intergranular porosity is present
as result of dissolution of pore-filling carbonate cement and ranges from 34 m
to 211 m in size (Appendix 6-32c, 6-33b-c, 6-42c, 6-43a-b and 6-46c).
However, intragranular porosity exists mostly as result of partial or complete
dissolution of feldspar and volcanic rock fragments (Appendix 6-32b, 6-43c, 644a, c and 6-51a; Mansurbeg et al., 2008) and varies in size between 45 m
and 199 m. Mouldic pores are observed in both coarse- and fine-grained
sandstone but are more common in coarse-grained beds. In the Narrabeen
Group, it mainly occurs due to partial dissolution of plagioclase with sizes
between 31 m and 234 m (Appendix 6-33a, 6-44b and 6-48a). Also,
dissolution of detrital grains such feldspar can result oversized pores in
sandstones with size ranges from 65 m to 244 m (Appendix 6-44c and 6-45a;
Lima and De Ros, 2002). Connectivity is shown in this unit by mouldic and
oversized pores. Corroded traces of feldspar occur in both mouldic pores and
oversized pores. Also, packing inhomogeneity and honeycombed grains are
observed in secondary pores.

7-7-3 Hawkesbury Sandstone


7-7-3-1 Primary porosity
Porosity is mostly primary (Appendix 6-45b-c and 7-8a, d, f, 7-9a, c, d), varying
between 0% and 16.5% (Appendix 5-26) at an average of 7.4% in sandstone
samples whereas, it is absent in siltstone samples (Table 7-4). and ranges from
30 to 356 m in size.
7-7-3-2 Secondary porosity
Secondary porosity is less abundant than primary porosity in the Hawkesbury
Sandstone (Appendix 6-38c, 6-39a, 7-8c and 7-10c). It occurs up to a maximum
of 4.3% with an average of 1.7% in sandstone samples (Appendix 5-26 and
Table 7-4). Secondary pores range in size between 20 and 220 m in the
Hawkesbury Sandstone.

727

7-8 Distribution of thin section porosity with facies


Blevin et al. (2007) introduced a study of the Sydney Basin gas reservoirs. They
showed that original grain size affects the primary porosity in sandstones. Thus,
primary porosity is high in coarse- to medium-grained sandstones whereas it is
less in fine-grained sandstone which are characterised by clays.

7-8-1 Illawarra Coal Measures


In the Illawarra Coal Measures, three facies are recorded including fine-grained,
medium-grained and coarse-grained. Fine-grained facies show very low thin
section porosity varying between 0 and 2.7% (Appendix 5-7) at average of 0.1%
in fine-grained sandstone and 0.04% in siltstone samples (Appendix 5-10). The
medium- and coarse-grained facies have thin section porosities at an average
of 1.4% and 2.7%, respectively (Appendix 5-10).
The relationship between grain size and thin section porosity shows a positive
low correlation (r2 = 0.2; Fig. 7-24). This relationship indicates that the influence
of grain size on thin section porosity is low in the Illawarra Coal Measures. This
indicates that the influence of diagenesis has masked the impact of texture on
porosity according to interpretation of Baker et al. (2000).
y = 29.592x + 231.72
R = 0.2057

900
800
Grain size (micron)

700
600
500
400
300
200
100
0
0

10
Porosity (%)

15

20

Figure 7-24: A low positive correlation is recorded between


thin section porosity and grain size in the Illawarra Coal
Measures.
722

A very low negative correlation is recorded between sorting and thin section
porosity in the Illawarra Coal Measures (Fig. 7-25a). In the Lawrence
Sandstone, there is a moderate negative correlation between sorting and thin
section porosity (r2 = 0.2; Fig. 7-25b).
a)
y = -5.4907x + 4.7898
R = 0.0655

16
14

Porosity (%)

12
10
8
6
4
2
0
0.0

0.2

0.4

0.6

0.8

1.0

1.2

Sorting

b)
y = -9.8858x + 8.1444
R = 0.1704

16
14

Porosity (%)

12
10
8
6
4
2
0
0.0

0.2

0.4

0.6

0.8

1.0

1.2

Sorting

Figure 7-25: a) Sorting is weakly negatively correlated with


thin section porosity in the Illawarra Coal Measures, b) Sorting
is moderately negatively correlated with thin section porosity in
the Lawrence Sandstone.

722

7-8-2 Narrabeen Group


In this unit, fine-grained, medium-grained and coarse-grained facies are
present. Thin section porosity is very low in the fine-grained facies and ranges
from 0 to 0.5% (Appendix 5-8) with an average of 0.05% in fine-grained
sandstone whereas is absent in siltstone samples (Appendix 5-11). It tends to
increase with increasing grain size and is observed at an average of 5.6% in the
medium-grained facies and at an average of 7.2% in the coarse-grained facies
(Appendix 5-11). Thin section porosity is weakly positively correlated with grain
size in the Narrabeen Group (Fig. 7-26). This indicates that diagenesis has a
greater influence than texture on porosity according to the interpretation of
Baker et al. (2000).
The sorting is weakly negatively correlated with thin section porosity in the
Narrabeen Group (Fig. 7-27).

y = 18.008x + 312.98
R = 0.13

1800
1600
Grain size (micron)

1400
1200
1000
800
600
400
200
0
0

10

15

20

25

Porosity (%)

Figure 7-26: a) There is a low positive correlation between


porosity and grain size in the Narrabeen Group.

722

25

y = -5.77x + 9.0673
R = 0.024

Porosity (%)

20

15

10

0
0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

Sorting

Figure 7-27: There is a low negative correlation between


sorting and porosity in the Narrabeen Group.

7-8-3 Hawkesbury Sandstone


The Hawkesbury Sandstone contains three facies comprising fine-grained,
medium-grained and coarse-grained sandstone. Thin section porosity is
completely absent in the fine-grained facies, whether fine-grained sandstone or
siltstone samples. It varies between 0 and 16% in the medium-grained facies at
an average of 7.4% (Appendix 5-9 and 5-12). In the coarse-grained facies thin
section porosity ranges from 1.6% to 19.3% at an average of 11.6% (Appendix
5-9 and 5-12).
The influence of grain size on thin section porosity in this unit is more apparent
in this unit than in the Illawarra Coal Measures or the Narrabeen Group. Thin
section porosity is moderately positively correlated with grain size whereas no
clear correlation is apparent between porosity and sorting in the Hawkesbury
Sandstone (Figs 7-28a, b). This indicates that the influences of diagenesis is
greater than the impact of texture on porosity according to interpretation of
Baker et al. (2000).
722

y = 11.785x + 262.82
R = 0.3074

700

Grain size (micron)

600
500
400
300
200
100
0
0

10

15

20

25

Porosity (%)
y = -5.1016x + 11.701
R = 0.0171

25

Porosity (%)

20
15
10
5
0
0.0

0.5

1.0

1.5

Sorting

Figure 7-28: a) Porosity is moderately positively correlated with


grain size in the Hawkesbury Sandstone, b) No significant
correlation is observed between sorting and porosity in the
Hawkesbury Sandstone

722

722

Chapter Eight
Discussion
8-1 Introduction
The composition of sandstone can be controlled by many factors including
diagenesis (Sur et al., 2002), transportation, climatic regime (Zheng et al.,
2006), depositional environment (Marenssi et al., 2002; Allen and Johnson,
2010), weathering (Korsch, 1984; Ghazi and Mountney, 2011) and source
composition (Hendrix, 2000; Zheng et al., 2006). Other studies have also shown
the influence of transportation distance, source rocks and post-depositional
variations on sandstone composition (Krainer and Spotl, 1989; Das and Haake,
2003; Jin et al., 2006; Akarish and El-Gohary, 2008). Provenance, sediment
transport, weathering and tectonic setting can also be determined from
sandstone composition (e.g. Jargal and Lee, 2006). Previous studies have also
considered the relationships between diagenetic minerals, depositional facies
and sequence stratigraphy in sandstone (Taylor et al., 1995; Morad et al., 2000;
Ketzer et al., 2003a, b; Al-Ramadan et al., 2005; El-Ghali et al., 2006a-c; Kordi
et al., 2011).
Many factors can affect reservoir quality including composition, diagenesis,
depositional environments and tectonic setting (e.g. Worden et al., 2000; dos
Anjos et al., 2000; Schmid et al., 2004; Buyukutku and Sahinturk, 2005; Reed et
al., 2005; Salem et al., 2005; Zhang et al., 2008; Wolela, 2009; Islam, 2009; Luo
et al., 2009; Zhang et al., 2010; Umar et al., 2011).
Factors that control compositional variation between sandstone units in the
Southern Sydney Basin, including provenance, environment of deposition and
diagenesis, are discussed in this chapter. In addition, this chapter also
considers the influence of alteration and diagenesis on reservoir quality in the
Southern Sydney Basin.

259

8-2 Linking depositional environment, provenance and


composition variation
8-2-1 Illawarra Coal Measures and Narrabeen Group
In the Illawarra Coal Measures, Bamberry (1992) showed that sandstone in the
quartz suite is mainly sublitharenite derived from a cratonic source: the Lachlan
Fold Belt. The lithic suite includes volcarenite, chert-arenite and litharenite. The
Gerringong Volcanics is the source of the volcarenite whereas the chert-arenite
and litharenite are mostly derived from the New England Orogen (Sims 1996).
Provenance also has a significant control on the detrital composition of the
Narrabeen Group according to Bai (1991) and he indicated that the influence of
transportation and depositional environment are not as important because the
deposition of the whole Narrabeen Group occurred in fluvial/lacustrine
depositional systems. The Lachlan Fold Belt, New England Fold Belt and
eastern volcanic source all contribute to Narrabeen Group sedimentation. The
Lachlan Fold Belt and New England Fold Belt source rocks were discussed by
Conaghan et al. (1982). Ward (1971b; 1972) indicated that the middle
Narrabeen Group was derived from the Gerringong Volcanic facies which
initially emerged during the sedimentation of the Shoalhaven Group (Jones,
1990). Bai (1991) showed that the New England Fold Belt includes high
percentages of chert, which is uncommon in the Lachlan Fold Belt. In this study,
the Lachlan Fold Belt and New England Fold Belt are indicated as sources for
detrital grains in the Narrabeen Group, whereas the Gerringong Volcanic facies
was not indicated. The Narrabeen Group includes volcanic detritus representing
a felsic rather than a mafic source. The felsic volcanic detritus was also sourced
from the New England Fold Belt according to Dehghani (1994).
In this study, the Illawarra Coal Measures were determined to consist of
litharenite and rarely sublitharenite (based on Folk, 1968; Fig. 5-3a; Table 5-1).
The samples from the Illawarra Coal Measures plot in the lithic recycled,
transitional recycled and quartz recycled provenance fields according to
Dickinson (1985; Fig. 5-7a; Table 5-1). This unit was deposited in fluvial and
floodplain facies. The abundance of medium- to coarse-grained sandstone and
fine-grained mudrocks illustrate the widespread occurrence of lithic grains with
260

only a minor presence of quartz grains in the Illawarra Coal Measures. In the
study area, quartz and lithic grains in the Illawarra Coal Measures were derived
from the Lachlan Fold Belt and the New England Orogen and deposited in
localized fluvial and floodplain facies. Thus, provenance is related to
depositional environment. There is no observed large difference in the textural
characteristics between the various units of the Illawarra Coal Measures. This
may indicate that this unit was characterized by similar weathering conditions
(Wani and Mondal, 2009). More lithic grains are observed in the Illawarra Coal
Measures than in the Narrabeen Group, while lithic grains are rare in the
Hawkesbury Sandstone. The percentages of quartz grains in the Illawarra Coal
Measures are not much different from the Narrabeen Group whereas they are
much less than in the Hawkesbury Sandstone.
In this study, the Narrabeen Group cannot be differentiated from the Illawarra
Coal Measures according to the triangular QFL and QmFLt mineralogical
diagrams. The QFL plot shows that the Narrabeen Group consists of litharenite,
sublitharenite and quartzarenite (Folk, 1968; Fig. 5-21a, Table 5-1) while the
QmFLt plot shows that the Narrabeen Group is sourced from lithic recycled,
transitional recycled, quartz recycled and craton interior sources (Dickinson,
1985; Fig. 5-24a, Table 5-1). This similarity may be coupled with the similarity in
the depositional environment. The Narrabeen Group contains coarse-grained
and fine-grained sandstone that was deposited in fluvial and floodplain
environments. These environments are similar to those in the Illawarra Coal
Measures, thus there is not much difference in provenance or depositional
setting between the Narrabeen Group and Illawarra Coal Measures.
Composition variations occur between the various units in the Narrabeen
Group. The lower Narrabeen Group contains more abundant and widespread
lithic fragments than the upper Narrabeen Group, which includes more common
detrital quartz. Thus, the difference in composition between the upper and lower
Narrabeen Group indicates changing source terranes. This is reflected by the
change from litharenite to sublitharenite and quartzarenite. Up-sequence, the
increasing detrital quartz content prevents the occurrence of the unusually lithicrich interval from occupying the entire sequence. The difference between the
261

two parts of the sequence is also related to depositional environment and grain
size. The lower part of the Narrabeen Group consists mostly of fluvial deposits,
including more lithic grains, whereas the upper part is mostly floodplain deposits
and includes more quartz grains. Also, the grain size in the lower part of the
group is larger than in the upper part, supporting the abundance of lithic grains
in the lower part.
In the Illawarra Coal Measures and Narrabeen Group, quartzarenite and
sublitharenite samples include large monocrystalline quartz grains suggesting
that granite is the source of some of these grains (Appendix 6-1a-b). The sharp
extinction of the monocrystalline quartz grains supports this interpretation. In
contrast, evidence suggests that monocrystalline quartz grains in the lithic-rich
litharenite samples are sourced from a volcanic origin. A number of these grains
contain euhedral crystals of monocrystalline quartz. Based on grain size, two
sizes of monocrystalline quartz were determined: medium- to coarse grained
and fine-grained. The former is sourced from the weathering of granites (Basu
et al., 1975; Pettijohn et al., 1987; Datta, 2005; Dutta, 2007; Ghazi and
Mountney, 2011) whereas the latter is mainly sourced from an igneous
provenance (Dickinson, 1970; Ghazi and Mountney, 2011). Also, some of this
quartz could have been reworked from earlier sedimentary units in the source
areas. Also, in the mineralogy chapter, it was shown that some quartz grains
are characterized by undulose extinction and may include heavy minerals and
fluid inclusions. These characteristics confirm a plutonic origin as the source for
such quartz grains (Suzuki et al., 2000). Grain size and source areas also
control the abundance of polycrystalline quartz in the sandstone. Polycrystalline
quartz grains having more than three crystals were sourced from metamorphic
rocks (Appendix 6-3a-c; Blatt, 1967; Blatt et al., 1980; Asiedu et al., 2000; Ghazi
and Mountney, 2011). Polycrystalline quartz grains are present in fluvial,
floodplain, lagoonal and shallow marine deposits. More polycrystalline quartz
grains are present in the Narrabeen Group than in the Illawarra Coal Measures
but they both have lower total quartz contents than in the Hawkesbury
Sandstone. In the fine-grained deposits, X-ray diffraction showed that quartz
shows a strong negative correlation with clay minerals in both the Illawarra Coal
Measures and Narrabeen Group (r2 = 0.5; r2 = 0.8; Figs 5-18 and 5-35a). These
262

relationships indicate that quartz is more abundant in cleaner sandstone facies


with low clay mineral contents in these two groups. In coarse-grained
sandstones, particularly the Kembla Sandstone, there is a strong negative
correlation between quartz content and grain size (r2 = 0.7, Fig. 5-11a), whereas
in the Coalcliff Sandstone, this reduces to a moderate negative correlation
between quartz content and grain size (r2 = 0.3; Fig. 5-27a). This provides
evidence that the fine-grained deposits are richer in monocrystalline quartz than
coarser grained lithic-rich sandstones (cf. Espejo and Lopez-Gamundi, 1994).
Plagioclase exceeds K-feldspar in most units of the Illawarra Coal Measures
and Narrabeen Group (Appendix 5-13 and 5-14). A low negative correlation
occurs between feldspar and quartz grains in the Narrabeen Group (r2= 0.2; Fig.
5-36), indicating that feldspar decreases with increasing quartz content based
on X-ray diffraction. Volcanic and plutonic rocks are recorded as sources of
plagioclase grains whereas plutonic rocks are represented as sources for
orthoclase grains (Kwon and Boggs, 2002). The influence of weathering on
plagioclase is larger than on K-feldspar as indicated by the degree of
weathering and diagenetic alteration. In the Illawarra Coal Measures and
Narrabeen Group, the percentages of feldspar are similar in all units (Appendix
5-13 and 5-14). Detrital feldspar occurs in fluvial environments more than in
shallow marine environments where it may be destroyed because grain-to-grain
collisions lead to breakage (Mack, 1978). Less feldspar is recorded in the
Narrabeen Group than in the underlying Illawarra Coal Measures, whereas it is
more common in both these groups than it is in the overlying Hawkesbury
Sandstone. Thus, feldspars are more common in the Illawarra Coal Measures
than in the Narrabeen Group or Hawkesbury Sandstone. Also, the Illawarra
Coal Measures are more directly related to the volcanic source with probably
shorter transport distances.
Volcanic rock fragments combined with chert and carbonate-replaced fragments
are most common in the Illawarra Coal Measures and Narrabeen Group, easily
exceeding the abundance of sedimentary rock fragments (Appendix 5-13 and 514). Thus volcanic rocks are classified as the main sediment source (cf. Kwon
and Boggs, 2002). Felsic grains are recorded as volcanic detritus in both the
263

Illawarra Coal Measures and Narrabeen Group, sourced mainly from New
England. Grains consisting of both quartz and feldspar are included with the
felsic grains and are probably sourced from silicic volcanic rocks, whereas
composite clay, feldspar and quartz grains are included with the vitric grains
which are sourced from glassy flows and tuffs. Both the Illawarra Coal
Measures and Narrabeen Group may reflect short transport distances due to
the presence of mafic volcanic clasts in most their units (cf. Willner et al., 2001).
Also, the short-distance of transport of detritus in the Illawarra Coal Measures
and Narrabeen Group is supported by the abundance of total lithic fragments
(Appendix 6-6c, 6-7a, 6-7b-c, 6-8a, 6-8b-c and 6-9a) as well as the presence of
only minor quantities of rounded quartz grains as shown by Asiedu et al. (2000).
Sedimentary rock fragments are uncommon throughout the studied section,
which may be a result of chemical weathering (Caracciolo et al., 2011) or a
scarcity of suitable source rocks especially if a lot of the detritus is being
derived from a volcanic arc. Inter-channel floodplain deposits are the source for
some of the shale fragments which consist of intrabasinal mudclasts. Shale
fragment sizes exceed the size of adjacent detrital grains and also partially
engulf some hard detrital grains as a result of mechanical compaction. The
presence of a sedimentary provenance can be determined through the
occurrence of sedimentary rock fragments in the Illawarra Coal Measures and
Narrabeen Group. A minor clastic sedimentary source is suggested by the
presence of sedimentary lithic grains, quartz grains and recycled quartz
overgrowth-bearing grains in the Illawarra Coal Measures and Narrabeen Group
(Mackey, 2009). Chert grains are widespread in sandstone samples rich in
detrital volcanic lithic fragments. In this study, the New England Fold Belt is
suggested as the major source of lithic grains in the Illawarra Coal Measures
and Narrabeen Group. Thus, the New England Fold Belt was a more important
and active source during the deposition of both these units than the Lachlan
Fold Belt. This was probably associated with a decrease in the importance of
the mafic volcanic arc responsible for the Gerringong Volcanics. Lithic grains
are distributed in all facies but are more common in fluvial channel facies,
represented by medium- and coarse-grained sandstone. These facies are rich
in lithic grains derived from the New England Fold Belt. Lithic grains are
264

uncommon in shallow marine (Bargo Claystone) and lagoonal deposits


(Newport Formation). Lithic grains are common in the Illawarra Coal Measures
and Narrabeen Group but are rare in the Hawkesbury Sandstone. Several
factors may have caused the source area to have a weak weathering intensity
when lithic grains were widespread in the Illawarra Coal Measures and
Narrabeen Group. These factors are short transport distance and arid climatic
conditions (Jargal and Lee, 2006) but they are probably also related to the
presence of a local elevated volcanic source with abundant loose volcanic ash.
Tourmaline and rare zircon grains (Appendix. 6-10b) are present in the Illawarra
Coal Measures and Narrabeen Group and may be important in the
determination of provenance (Morton, 1991; Henry and Dutrow, 1992; Gotze
and Blankenburg, 1994; von Eynatten and Gaupp, 1999; Li et al., 2004; Morton
et al., 2005; Nascimento et al., 2007). However, many heavy mineral grains,
such as tourmaline, zircon and rutile, are present as rounded grains that are
associated with sedimentary lithic fragments, indicating derivation from older
(pre-existing) sedimentary rocks according to Osae et al. (2006). Some heavy
minerals are recorded in the form of angular grains. Tourmaline grains are
observed in fluvial channel, floodplain, lagoonal and shallow marine deposits as
rounded and angular grains. There is generally no influence of chemical and
mechanical abrasion on tourmaline grains (Nascimento et al., 2007). The
occurrence of tourmaline grains in most samples supports that these units may
be sourced from meta-sediments and granites (Anani, 1999; Nascimento et al.,
2007). Zircon grains are also recorded in fluvial channel, floodplain, lagoonal
facies and shallow marine deposits and may indicate the source area based on
their morphologies. A recycled and metasedimentary source is indicated by the
rounded to well rounded zircon grains in these units. In contrast, first cycle
sediments are interpreted from the presence of euhedral and sub-euhedral
forms which are typical of many zircon and tourmaline grains in the Illawarra
Coal Measures and Narrabeen Group (cf. Nascimento et al., 2007). Prolonged
transport and reworking of zircon grains can be a contributing factor to the
difference of zircon morphologies (Nascimento et al., 2007). Also, Suzuki et al.
(2000) showed that older sedimentary rocks can be indicated by round zircon
grains. Zircon grains can be transported for long distances and can also be
265

subject to strong reworking influences. This is demonstrated by sub-rounded to


sub-angular zircon grains as shown by Grades (2003). Minor rutile grains also
occur in the fluvial channel, floodplain, lagoonal facies and shallow marine
deposits.
At the base of the Sydney Subgroup, the Wilton Formation is mostly litharenite
including an abundance of lithic grains, indicating that the New England Belt
Fold is very important in this unit (Fig. 5-3b and Table 5-1). Also, a local source
almost certainly contributed as shown by northwest palaeocurrents in the lower
Wilton Formation. According to Dickinsons (1985) classification, this unit plots
in lithic recycled and transitional recycled to quartz recycled provenance areas
(Fig. 5-7b and Table 5-1). Also, this unit was deposited in floodplain facies and
includes common siltstone and shale with minor sandstone. Clay minerals are
common in this unit based on X-ray diffraction analysis which may be due to the
abundance of fine-grained deposits and matrix in this unit. The Wilton
Formation is derived from volcanic sources.
Up-sequence, the Bargo Claystone is different from the underlying Wilton
Formation as result of the difference in provenance and depositional
environment. One sample from the Bargo Claystone was described since this
unit includes mostly siltstone and shale. This unit plots as sublitharenite
according to Folk (1968; Fig. 5-3b and Table 5-1) and is sourced from a quartz
recycled provenance according to Dickinson (1985; Fig. 5-7b and Table 5-1).
These variations reflect the change in the source from lithic recycled,
transitional recycled and quartz recycled provenance for the Wilton Formation to
a quartz recycled provenance for the Bargo Claystone. Although parts of the
upper Wilton Formation were deposited in a marginal marine environment,
floodplain deposits are more common. The transitional from the floodplain
deposits in the Wilton Formation to the shallow marine deposits in the Bargo
Claystone affected composition but this is more related to grain size variation
and hence depositional environment. The detrital quartz amount increases from
the Wilton Formation to the Bargo Claystone, which is associated with the
decreasing content of feldspar and lithic grains due to the change to a lower
energy environment. Sorting is better in the sandy portions of the Bargo
266

Claystone than in the underlying unit and may contribute to the preservation of
higher percentages of quartz (cf. Marenssi et al., 2002). The Bargo Claystone
has a small average grain size, including siltstone and shale, thus lithic grains
are rare. Detrital quartz is more common than lithic grains in the shallow marine
deposits of the Bargo Claystone (Appendix 5-13 and 5-14) and its abundance
was initially derived from a cratonic source in the Lachlan Fold Belt. Eastward
thinning of the Bargo Claystone was in response to the increased distance from
the source and the slower rate of subsidence near the margin with cratonic
Lachlan Fold Belt (Roche, 1997). Non-undulose monocrystalline quartz grains
are observed more commonly in shallow marine environments compared with
fluvial

environments

whereas

undulose

monocrystalline

quartz

and

polycrystalline quartz grains are less mechanically stable, thus they decrease in
abundance in shallow marine environments (cf. Mack, 1978). Detrital feldspar is
rare in this unit according to thin section analysis (Appendix 5-13 and 5-14) and
may be obliterated because grain-to-grain collisions lead to breakage of
feldspar grains as shown by Mack (1978). This suggests that unstable detrital
feldspar grains are less mechanically stable in shallow marine environment (cf.
Mack, 1978).
Total quartz decreases from the Bargo Claystone upwards to the Darkes Forest
Sandstone and Allans Creek Formation due to the increase in the amount of
lithic and feldspar grains as the grain size increases (Appendix 5-13 and 5-14).
This variation in composition is coupled with and controlled by both provenance
and depositional environment. The composition suggests that the source varies
from quartz recycled for the Bargo Claystone to lithic recycled and transitional
recycled for the Darkes Forest Sandstone and to lithic recycled and quartz
recycled for the Allans Creek Formation (according to plots from Dickinson,
1985; Fig. 5-8a and Table 5-1). Provenance determinations were affected by
the environmental change from shallow marine deposits for the Bargo
Claystone to fluvial and floodplain deposits for the Darkes Forest Sandstone
and Allans Creek Formation. This change in environment and increase in grain
size supports the increase in lithic and feldspar grains with a decrease in quartz
grains in the Darkes Forest Sandstone and Allans Creek Formation. Thus the
Bargo Claystone is characterized by shale and siltstone whereas the Darkes
267

Forest Sandstone and Allans Creek Formation have common fine- to mediumgrained sandstone. The increase in feldspar grains in the Darkes Forest
Sandstone and Allans Creek Formation compared to the Bargo Claystone may
be due to rapid sedimentation which supports the presence these grains (cf.
Marenssi et al., 2002). In the Darkes Forest Sandstone and Allans Creek
Formation, lithic grains are derived from the New England Belt.
The Kembla Sandstone falls into the litharenite field and has a lithic recycled,
transitional recycled and quartz recycled provenance according to classification
diagrams in Folk (1968; Fig. 5-4b and Table 5-1) and Dickinson (1985; Fig. 5-8b
and Table 5-1), respectively. Lithic grains trend to increase whereas quartz and
feldspar grains decrease in this unit compared with the underlying Darkes
Forest Sandstone and Allans Creek Formation (Appendix 5-13 and 5-14).
Sandstone composition is related to depositional environment where fluvial
channel activity is more important in the Kembla Sandstone than in the
underlying units. Thus, the Kembla Sandstone contains mostly medium- to
coarse-grained sandstone which has more lithic grains, derived from the New
England Orogen, localized within the fluvial channel facies. This means that
grain size is larger in the Kembla Sandstone than in the underlying units
(Appendix 5-13 and 5-14) and this increased grain size contributes to the
decrease in quartz grains. In the Kembla Sandstone, a high negative correlation
is present between quartz and grain size (r2=0.7; Fig. 5-11a), supporting the
decrease in quartz grains with increasing grain size. Also, lithic grains are
moderately positively correlated with grain size (r2=0.4; Fig. 5-13b), and this
supports the increase in lithic grains associated with an increase in grain size.
This also leads to a negative relationship between quartz and lithic grains in this
unit (r2=0.3; Fig. 5-12). X-ray diffraction results show the relationship between
clay minerals and quartz in fine- and medium-grained samples to have a
moderate negative correlation (r2=0.4; Fig. 5-19a). This suggests that the
abundance of clays in some samples affects the quantities of quartz grains. In
thin section, sorting is worse in this unit than in the underlying units, thus quartz
grains continue to decrease in this unit (cf. Marenssi et al., 2002).

268

The Wongawilli Coal includes mainly coal and tuff, and only one sample was
described. This sample showed that the Wongawilli Coal has a litharenite
composition and originated from a lithic recycled provenance according to
graphs from Folk (1968; Fig. 5-5a and Table 5-1) and Dickinson (1985; Fig. 59a and Table 5-1), respectively. The facies in this unit are different from those in
the underlying Kembla Sandstone, indicating the effects of variations in
environment and grain size. Thus, the Wongawilli Coal is a floodplain and
swamp deposit and has a much smaller grain size than in the Kembla
Sandstone (Appendix 5-13 and 5-14). These variations did not affect the
provenance of this unit compared with the underlying Kembla Sandstone
possibly because only one sample was described from the Wongawilli Coal.
The sequence from the Darkes Forest Sandstone to Wongawilli Coal is rich in
lithic grains which were derived from the New England Orogen. Thus, during
deposition of the Darkes Forest Sandstone, Allans Creek Formation, Kembla
Sandstone and Wongawilli Coal the New England Belt Fold was a more active
sediment source than it was for the underlying the Bargo Claystone. There was
no observed change in abundance of lithic grains from the Darkes Forest
Sandstone to Wongawilli Coal. The increases in lithic and feldspar grains,
associated with a decrease in quartz grains, indicate that detritus higher in the
section is more immature (cf. Zheng et al., 2006).
Up-sequence, lithic grains in the Unnamed Three Member and Unnamed Two
Member decrease gradually with increase in quartz grains (Appendix 5-13 and
5-14). The QFL triangular diagram shows that these units fall into the litharenite
field according to Folks (1968; Fig. 5-5a and Table 5-1) classification. Also, the
QmFLt triangular diagram shows that the Unnamed Three Member plots in the
lithic recycled, transitional recycled and quartz recycled provenance fields but
Unnamed Two Member has a transitional recycled provenance according to
Dickinsons (1985; Fig. 5-9a and Table 5-1) classification. These units are
composed of fine-grained deposits such as fine-grained sandstone, siltstone
and shale. Thus, they are floodplain deposits with no coarser fluvial channel
deposits. Therefore, the lithic grain contents decrease in these units compared
to the underlying unit and they are mainly derived from the New England
Orogen. In these units, grains are moderately to well sorted. This character
269

supports the abundance of quartz grains in these units compared with the
underlying unit (cf. Marenssi et al., 2002).
Up-sequence, lithic grain contents increase in the Lawrence Sandstone and are
more common than quartz grains (Appendix 5-13 and 5-14), suggesting the
presence of lithic recycled, transitional recycled and quartz recycled sources for
this unit according to Dickinsons (1985; Fig. 5-9b and Table 5-1) classification.
The Lawrence Sandstone consists of litharenite and rarely sublitharenite
according to Folks (1968; Fig. 5-5b and Table 5-1) classification. This unit is
somewhat different in provenance from the Unnamed Two Member. This
difference is probably related to facies and grain size. Fluvial channel activity is
recorded in the Lawrence Sandstone through the abundance of mediumgrained sandstone. Thus, the average grain size in the Lawrence Sandstone is
larger than in the Unnamed Two Member (Appendix 5-13 and 5-14) and is
associated with an increase in lithic grains and a good positive correlation
between lithic grains and grain size (r2=0.6; Fig. 5-14a). The Lawrence
Sandstone also contains floodplain deposits recognized by the presence of finegrained facies which also include more lithic grains than quartz grains indicating
the significance of the New England Fold Belt as the dominant source in this
part of the basin. Clay minerals show a moderate negative correlation with
quartz in the fine- and medium-grained samples (r2=0.4; Fig. 5-19b) according
to X-ray diffraction. This suggests that clay minerals are a minor component in
samples which are characterized by common quartz grains.
The Lawrence Sandstone is similar to the overlying Loddon Sandstone which
includes even more lithic grains with a decrease in quartz grains (Appendix 5-13
and 5-14) and is classified as litharenite (Folk, 1968; Fig. 5-6 and Table 5-1).
Thus, the Loddon Sandstone is derived from a lithic recycled source (after
Dickinson, 1985; Fig. 5-10 and Table 5-1). This unit is characterized more than
the underlying unit by coarse sandstone, thus it is a high energy fluvial channel
deposit. Siltstone and shale are also present and represent associated
floodplain deposits. In this unit, fluvial activity is greater and grain size is larger
than in the underlying unit. Thus, provenance determination is affected and it is
somewhat different from the underlying unit since lithic grains have increased
270

whereas quartz and feldspar grains decreased. The textural characteristics of


the Lawrence Sandstone and Loddon Sandstone are similar to the Kembla
Sandstone and the lithic grains in all these units were derived from the New
England Orogen. In the Loddon Sandstone, X-ray diffraction showed that clay
minerals are strongly negatively correlated with quartz grains in fine- and
medium-grained sandstone (r2=0.9; Fig. 5-20a). Also, kaolinite is moderately
negatively correlated with quartz grains in fine- and medium-grained sandstone
(r2=0.5; Fig. 5-20b). These correlations indicate that the sandstone beds in the
Loddon Sandstone are relatively free of clay matrix.
At the base of the Narrabeen Group, the Coalcliff Sandstone is present. In
previous studies, the Coalcliff Sandstone was identified as litharenite
(Loughnan, 1963), lithic litharenite (Bai and Keene, 1996) and quartz-rich
litharenite (Deen, 1999). In this study, the Coalcliff Sandstone falls into the
litharenite field and rarely into quartzarenite according to Folk's (1968; Fig. 521b and Table 5-1) classification and plots in the lithic recycled, transitional
recycled, and rarely quartz recycled and craton interior provenance areas,
according to Dickinsons (1985; Fig. 5-24b and Table 5-1) classification. This
change in provenance is related to depositional environment with the transition
from floodplain swamp facies in the Bulli Coal to fluvial channel facies in the
Coalcliff Sandstone. Thus coarse-grained sandstone is more common in the
Coalcliff Sandstone, and the average grain size is larger and lithic grains are
higher in this unit than in the underlying unit. This variation in facies and grain
size affects the interpreted provenance. This unit was mostly deposited as a
mixed-load fluvial deposit and the percentage of lithic grains indicates the
significance of the New England Fold Belt as the source for the localized
channel facies. Quartz grains are present in the Coalcliff Sandstone and were
also sourced from the New England Fold Belt and localized in the channel and
floodplain facies. In the Coalcliff Sandstone, quartz and lithic grains are both
correlated to grain size but whereas quartz decreases with increasing grain size
(r2=0.3; Fig. 5-27a), lithic grains increase with increasing grain size (r2=0.4; Fig.
5-30). There is also a good negative correlation between quartz and lithic grains
(r2=0.5; Fig. 5-28b). These correlations reflect the influence of grain size on
composition variation in the Coalcliff Sandstone. In fine- and medium-grained
271

sandstone the abundance of quartz grains and lack of matrix results in the clay
minerals being moderately negatively correlated with quartz grains (r2=0.4; Fig.
5-36a).
Up-sequence, the Wombarra Claystone is not very different from the Coalcliff
Sandstone in terms of provenance. Quartz increases weakly as the lithic grains
decrease (Appendix 5-13 and 5-14). Folk's (1968) classification diagram
indicates that the Wombarra Claystone plots in litharenite, sublitharenite and
quartzarenite fields (Fig. 5-22a and Table 5-1). The provenance of the
Wombarra Claystone is like to be that of the underlying Coalcliff Sandstone and
was derived from lithic recycled, transitional recycled, quartz recycled and rarely
craton interior provenance areas (Fig. 5-25a and Table 5-1). The abundant
detrital lithic grains in the Coalcliff Sandstone and Wombarra Claystone reveal a
volcanic source. The Wombarra Claystone includes mostly fine-grained
deposits, thus it was deposited as floodplain facies. This unit differs from the
underlying Coalcliff Sandstone which was mostly channel facies. The low
energy of the floodplain facies may contribute to the increase in quartz grains
and decrease in lithic grains in the Wombarra Claystone compared with the
Coalcliff Sandstone. In general, this difference in environment did not affect
sandstone composition and provenance greatly between the Wombarra
Claystone and Coalcliff Sandstone. Quartz and lithic grains are observed in the
coarser floodplain facies of the Wombarra Claystone and come from the
Lachlan and New England Fold Belts whereas the finer floodplain facies
consists of illite, mixed-layer clays and muscovite identified by X-ray diffraction
analysis. In the Coalcliff Sandstone and Wombarra Claystone, during lower
energy periods, fine-grained deposits accumulated. However, syntectonic
conglomerate and coarse fluvial deposits originated during periods of uplift of
the New England Fold Belt (Ward, 1971b; 1972; Bowman, 1972; Herbert,
1980b; Hamilton et al., 1987) which also produced the southwesterly flowing
tributary streams which are main depositional elements in the Coalcliff
Sandstone and Wombarra Claystone (Hamilton et al., 1987).
According to Loughnan (1963), the Scarborough Sandstone is classified as
coarse lithic arenite and as lithic-rich arenite according to Bai and Keene (1996)
272

and Deen (1999). In this study, the proportion of quartz and feldspar grains
increases from the Coalcliff Sandstone to the Scarborough Sandstone,
accompanied by a decrease in the proportion of lithic grains in the Scarborough
Sandstone (Appendix 5-13 and 5-14). Thus, the variation in quartz content is
inversely related to the variation of detrital lithic grains. This indicates that the
Scarborough Sandstone can be classified as litharenite according to Folk (1968;
Fig. 5-22b and Table 5-1) and comes from transitional recycled, quartz recycled
and lithic recycled provenance areas according to Dickinsons (1985; Fig. 5-25b
and Table 5-1) classification. Thus, provenance of the Scarborough Sandstone
is not very different from the underlying Wombarra Claystone and Coalcliff
Sandstone. The increase in quartz and feldspar grains and the decrease in lithic
grains are associated with an increase in grain size. Also, sorting is better in the
Scarborough Sandstone than in the underlying Wombarra Claystone and
Coalcliff Sandstone, and quartz content commonly increases with better sorting
as shown by Marenssi et al. (2002). Thus, the role of grain size and sorting
indicates the importance of environment in the composition variation and
provenance determination. This means that the environmental transition from
floodplain facies in the Wombarra Claystone to channel and floodplain facies in
the

Scarborough

Sandstone

affects

the

sandstone

composition

and

provenance. Channel facies are more common in the Scarborough Sandstone


and show upward-fining. Coarse- and medium-grained sandstone in the
channel facies include common detrital lithic grains which continue throughout
the Scarborough Sandstone suggesting the importance of detrital material
sourced from the New England Fold Belt. Also, monocrystalline quartz
commonly increases in coarse- and medium-grained sandstones in the
Scarborough Sandstone, indicating increasing detrital quartz derived from the
Lachlan Fold Belt compared to the New England Fold Belt. Ward (1971b; 1972),
Herbert (1980b) and Hamilton et al. (1987) showed in their studies that the New
England Fold Belt was characterized by heightened tectonic activity which
produced the fluvial deposits of the Scarborough Sandstone. X-ray diffraction
analysis of the minor fine-grained deposits which were sourced from the New
England Fold Belt and occur in localized floodplain facies. This technique
showed that quartz decreased in the floodplain facies of the Scarborough
273

Sandstone when compared to the underlying units, associated with an increase


in clay minerals. This confirms the fact that quartz is highly negatively correlated
with clay minerals in this unit (r2= 0.9; Fig. 5-36b). Thus, the presence of clays
in the fine-grained deposits indicates a different suspended load transport
mechanism compared to the bed-load quartz grains.
Up sequence, detrital lithic fragments tend to increase in sandy interbeds in the
Stanwell Park Claystone associated with decreasing detrital quartz (Appendix 513 and 5-14). Two sandstone samples from the Stanwell Park Claystone were
classified as litharenite and are similar to the Coalcliff Sandstone, Wombarra
Claystone and Scarborough Sandstone (Fig. 5-23a and Table 5-1). The
Stanwell Park Claystone is characterized by floodplain activity and is mainly
derived from a lithic recycled source (Fig. 5-26a and Table 5-1). Provenance of
the Stanwell Park Claystone is different from provenance of the underlying
Scarborough Sandstone. This difference may be related to depositional
environment since the depositional environment in the Stanwell Park Claystone
is different from that in the underlying Scarborough Sandstone, which is
dominated by fluvial channel deposits. Sorting is worse in the Stanwell Park
Claystone than in the Scarborough Sandstone, and the quartz content
decreases as matrix increases (cf. Marenssi et al., 2002). Lithic grains from the
New England Fold Belt are present in sandy interbeds in the Stanwell Park
Claystone floodplain facies. Deposition of the Stanwell Park Claystone probably
occurred during a period of tectonic quiescence in the southern part of the
Sydney Basin. The northern part of the basin was subjected to strong
subsidence which resulted in only fine material being transported into the
southern part of the basin to form the floodplain deposits of the Stanwell Park
Claystone. This strong subsidence was related to renewed tectonic activity in
the New England Fold Belt to the north (Dehghani, 1994). Tectonic activity in
the New England Magmatic Belt had its last main stage during the deposition of
the floodplain facies of the Stanwell Park Claystone.
Up sequence, the Bulgo Sandstone was determined by Loughnan (1963), Bai
and Keene (1996) and Deen (1996) to be quartz-lithic arenite, quartz-rich
litharenite and litharenite, respectively. Cowan (1993) classified the upper Bulgo
274

Sandstone as litharenite. In this study, the Bulgo Sandstone is interpreted as


sublitharenite to rarely litharenite (Fig. 5-23a and Table 5-1) and is differentiated
from the underlying units by the presence of widespread detrital quartz
(Appendix 5-13 and 5-14). This indicates that the sandstone composition in the
Bulgo Sandstone is different from the sandstone composition in the underlying
Stanwell Park Claystone as a result of a difference in provenance and
depositional environment. The petrographic data showed the provenance of the
Bulgo Sandstone to plot in the quartzose recycled and rarely craton interior
fields and is different from the provenance of all the underlying units (Fig. 5-26a
and Table 5-1). Quartz grain content increases rapidly, associated with a large
decrease in lithic grains in this unit compared to the underlying units. Thus, a
high negative correlation exists between quartz and lithic grains in the Bulgo
Sandstone (r2= 0.8; Fig. 5-29). These changes indicate that provenance
interpretation is affected by depositional environment. Depositional style in this
unit is different from that in the underlying Stanwell Park Claystone. This means
a transitional from floodplain facies in the Stanwell Park Claystone to common
fluvial channel and rare floodplain facies in the Bulgo Sandstone, which
progressively fines upwards. This character in the Bulgo Sandstone was due to
a steady decrease in fluvial and tectonic activity from the New England Fold
Belt. Fluvial facies such as medium- to coarse-grained sandstone are common
in the Bulgo Sandstone and contain mostly detrital quartz, sourced from the
Lachlan Fold Belt, which is a more important source than the New England Fold
Belt; thus lithic grains are uncommon. Quartz grains are more common in this
unit than in the underlying units whereas lithic grains decrease in this unit
compared to the underlying unit (Appendix 5-13 and 5-14). This indicates the
impact of depositional environment on provenance determination and
sandstone composition. Sorting is better in the Bulgo Sandstone compared than
in the underlying units, which is reflected in the higher quartz content (cf.
Marenssi et al., 2002). Conaghan et al. (1982) and Cowan (1993) suggested
that the Lachlan Fold Belt is the source of the quartz grains in this unit, which is
confirmed by an anticlockwise change of palaeocurrent directions. Thus, the
Lachlan Fold Belt became more importance as a source up-section (Deen,
1999). Ward (1972) showed that the volcanic lithic fragments in the Bulgo
275

Sandstone still indicate contributions from a volcanic source. The abundance of


quartz grains in the Bulgo Sandstone indicates that the maturity of the
sediments increases upward (Korsch, 1984). Also, based on X-ray diffraction
analysis, a very strong negative correlation occurs between quartz grains and
clay minerals (r2= 0.9; Fig. 5-37a). This correlation indicates that the quartz-rich
sandstone contains relatively little clay matrix.
Up sequence, only one sandy sample from the Bald Hill Claystone was
prepared for thin section analysis because most of the samples were claystone.
The sandstone is classified as litharenite (Fig. 5-23b and Table 5-1), indicating
an increase in detrital lithic fragments with a decrease in detrital quartz in this
unit compared to the underlying unit (Appendix 5-13 and 5-14). This again
indicates the influence of provenance and depositional environment on
compositional variation. The Bald Hill Claystone is determined as coming from a
lithic recycled source (Fig. 5-26b and Table 5-1). Depositional environment has
influenced the provenance determination with the environment changing from
fluvial deposits in the Bulgo Sandstone to floodplain facies in the Bald Hill
Claystone. Thus, interpreted provenance of the Bald Hill Claystone is different
from the provenance of the Bulgo Sandstone. Floodplain facies contain finegrained deposits such as siltstone and shale which are most common. This
environmental change may be the reason for the increase in detrital lithic
fragments and decrease in detrital quartz in this unit. The Lachlan Fold Belt may
be the source of quartz grains in the floodplain facies of the Bald Hill Claystone
whereas the lithic grains were derived from the New England Fold Belt. Coarsegrained sediments disappear completely in this unit, and no fluvial activity is
recorded in this unit. Aggradation of sediments can be recorded as a factor that
contributes to deposition on the floodplain and this occurred without any major
tectonic activity. Grains are poorly sorted in the Bald Hill Claystone, thus quartz
grains have decreased in this unit compared to the underlying Bulgo Sandstone
as shown by Marenssi et al. (2002). Deen (1999) showed that hematite and
siderite are prevalent in the Bald Hill Claystone and the red-brown colour occurs
in the core samples. Loughnan (1963) indicated that the immediate source of
the detritus is the reason for difference between the Stanwell Park Claystone
and Bald Hill Claystone in terms of clay mineral contents.
276

Deen (1999) indicated that higher percentages of quartz content in the Newport
Formation are consistent throughout the entire study area. He recognized the
Newport Formation as quartzose litharenite. In this study, three samples from
the Newport Formation were interpreted as sublitharenite to quartzarenite
(Fig.5-23b and Table 5-1), and are characterized by higher quartz contents than
in other underlying units (Appendix 5-13 and 5-14). Feldspar and lithic grains
are present but in low abundance (Appendix 5-13 and 5-14). Thus, the highest
percentages of detrital quartz and the lowest percentages of detrital lithic grains
in the Narrabeen Group are noted in this unit. This is consistent with the
Newport Formation plotting in the quartzose recycled and craton interior fields;
thus it has difference provenance from underlying units. This is coupled with the
environmental change from floodplain facies in the Bald Hill Claystone to
lagoonal facies in the Newport Formation. Also, fluvial activity is absent in the
Newport Formation. This again indicates the influence of depositional
environment on provenance interpretation. Also, sorting and grain size affect
composition variation. This unit is better sorted than in the underlying Bald Hill
Claystone but grain size is smaller than in the analysed sandstone from the
underlying Bald Hill Claystone. Thus, the increase in quartz in this unit may be
partly due to the changes in sorting and grain size. This indicates that
provenance and depositional environment control composition variation. Quartz
grains in the Newport Formation are derived from the Lachlan Fold Belt which is
the main source for the Newport Formation. This unit is recorded as a
transitional stage from the Narrabeen Group to the Hawkesbury Sandstone
(Dehghani, 1994).

8-2-2 Hawkesbury Sandstone


Standard (1969) classified the source of the Hawkesbury Sandstone into two
areas: a southwestern source area and a western source area based on the
interpretation of heavy minerals and clay minerals. The southwestern source
area is the major source, and the Hawkesbury Sandstone appears to have been
derived from the southwest direction based on palaeocurrent studies (Standard,
1969; Veevers, 1984).

277

In this study, the Hawkesbury Sandstone is interpreted as sublitharenite and


quartzarenite according to Folks (1968; Fig. 5-39 and Table 5-1) classification.
According to Dickinsons (1985) classification, the Hawkesbury Sandstone plots
in a craton interior to rarely quartzose recycled provenance area because it has
a very high percentage of quartz with low feldspar and lithic grains (Fig. 5-40
and Table 5-1). This interpretation shows that the Hawkesbury Sandstone has a
provenance which is different from the Illawarra Coal Measures and Narrabeen
Group. Quartz grains are abundant whereas feldspar and lithic grains are minor
constituents as a result of changes in provenance and depositional
environment. Thus, the Hawkesbury Sandstone is more quartz-rich than the
underlying Newport Formation at the top Narrabeen Group as result of the
environmental change from lagoon environment to fluvial channel environment.
In the Hawkesbury Sandstone, channel facies are represented by medium- to
coarse-grained sandstone whereas the minor floodplain facies include finegrained sandstone, siltstone and shale. These facies have high proportions of
quartz grains which are derived from a quartz-rich source. On the other hand,
the top of the underlying Newport Formation consists of fine-grained deposits,
thus it contains less quartz grains than the Hawkesbury Sandstone. Also,
sorting is better in the Hawkesbury Sandstone than in the Narrabeen Group and
Illawarra Coal Measures. This also supports the abundance of quartz grains in
the Hawkesbury Sandstone (cf. Marenssi et al., 2002).
Quartz grains (Appendix 6-1c, 6-2c and 6-3c) are sourced from the cratonic
Lachlan Orogen and/or the New England Fold Belt. They are more common in
medium- and coarse-grained sandstone than in fine-grained facies. This
indicates that quartz grains are more abundant in the coarser fluvial channel
facies than in floodplain facies. Plutonic quartz is mainly of granitic origin,
whereas felsic lavas are the source for the volcanic quartz. Also, plutonic and
low- to high-grade metamorphic rocks are indicated as possible sources for the
Hawkesbury Sandstone due to the observation of strained and unstrained of
monocrystalline and polycrystalline quartz under the microscope, as shown by
Zheng et al. (2006). However metamorphic rocks are unlikely in the
Hawkesbury Sandstone catchment. Original polycrystalline quartz may have
been largely disaggregated since the Hawkesbury Sandstone is rich in
278

monocrystalline quartz (cf. Dabbagh and Rogers, 1983; Wanas and AbdelMaguid, 2006). Also, monocrystalline quartz grains show better sorting and
roundness in the Hawkesbury Sandstone than in the Narrabeen Group and
Illawarra Coal Measures. This is due to the larger grain size, longer transport
distance and relatively prolonged abrasion (Grades, 2003) suffered by the
quartz in the Hawkesbury Sandstone. This study showed that polycrystalline
quartz grains are more widespread in coarse-grained than in fine-grained
samples, perhaps because of the composite nature of the polycrystalline quartz
grains. Standard (1964) indicated that the parts of the basin closer to the source
area include more metamorphic quartz grains. In general, polycrystalline quartz
is less common than monocrystalline quartz in this unit (Appendix 5-13 and 514) and this may be due to the long transport distance (Dabbagh and Rogers,
1983; Ullah et al., 2006). A warm, semi-arid climate during the deposition of this
unit is demonstrated by the widespread presence of monocrystalline quartz and
the relative lack of unstable grains (Ghazi and Mountney, 2011). Also, X-ray
diffraction analysis showed a good negative correlation between quartz and
feldspar (r = 0.5; Fig. 5-43) and between quartz and clay minerals (r = 0.7; Fig.
5-44), indicating that the sandstone is rich in quartz grains with low feldspar and
clay mineral contents. In the Hawkesbury Sandstone, quartz amounts are much
higher than in the underlying Narrabeen Group and Illawarra Coal Measures.
Unstable components must be in low abundance or weathered in the source
rocks of the Hawkesbury Sandstone. Standard (1964) showed that the
Ordovician intermediate volcanic rocks in the Sofala or Wellington districts were
the source of plagioclase in some western parts of the basin. Also, he indicated
that no high feldspar-bearing facies are recorded in the southwestern part of the
basin (major source area) even though granitic rocks are important constituents
in the adjacent Lachlan Fold Belt. This indicates a long period of weathering in
the source area. In this study, less feldspar is present in the Hawkesbury
Sandstone than in the underlying Narrabeen Group and Illawarra Coal
Measures (Appendix 5-13 and 5-14). In the Hawkesbury Sandstone, feldspar
content may be destroyed by breakage and abrasion in moderately high energy
environments (Marenssi et al., 2002). K-feldspar is more common than
plagioclase in the Hawkesbury Sandstone indicating a plutonic origin (Hindrix,
279

2000; Osae et al., 2006; Akarish and El-Gohary, 2008). However, alteration of
plagioclase can occur more commonly in the source area (Grades, 2003)
accounting for the low plagioclase even from the common granodiorite plutons
in the source area. The chemical stability of alkali feldspar is higher than that of
plagioclase during transportation and weathering, and this supports the greater
abundance of alkali feldspar over plagioclase in the Hawkesbury Sandstone
(Tucker, 1991; Ullah et al., 2006). Ochoa et al. (2007) also showed that
chemical weathering played a role in the reduction of feldspar in the Iberian
Chain during the Early Cretaceous. This is consistent with the higher K-feldspar
content than plagioclase in the Hawkesbury Sandstone. Alkali feldspar-rich
granite is present in the source area, and it is also indicated by the occurrence
of detrital mica (Akarish and El-Gohary, 2008). Rock fragments are certainly
affected by weathering in the Hawkesbury Sandstone, thus their distinction from
clay pellets in the thin section is difficult. Siltstone and shale intraclasts are
included as sedimentary rock fragments even though they are derived from a
nearby floodplain source in the area of deposition. Lithic grains are very sparse
in the Hawkesbury Sandstone compared with the underlying Narrabeen Group
and Illawarra Coal Measures (Appendix 5-13 and 5-14). In the Hawkesbury
Sandstone, lithic fragments and feldspar contents are low, and K-feldspar is
higher than plagioclase. This supports the concept that this unit was subjected
to prolonged weathering, transportation or recycling as shown by Akarish and
El-Gohary (2008) and Tijani et al. (2010). Also, the source area was chemically
weathered during a warm humid climate, indicated by the rare occurrence of
feldspar and rock fragments (Pettijohn et al., 1987; Amireh, 1991; Wanas and
Abdel-Maguid, 2006).
The influence of weathering and diagenesis on heavy minerals such as rutile,
tourmaline and zircon is absent since these minerals are characterized by their
stability (Morton and Hallsworth, 1999; Fontanelli et al., 2009). For example,
Nascimento et al. (2007) showed that the influence of chemical and mechanical
abrasion on tourmaline grains is absent. These minerals can indicate plutonic
sources (Wanas and Abdel-Maguid, 2006) or they may be reworked from earlier
sediments. In the fluvial channel facies, two types of rutile grains were
determined based on their roundness: that is rounded and angular rutile grains.
280

Pre-existing sedimentary rocks are the source of the former whereas granitic
rocks are the source of the latter (Standard, 1969). Also, rounded and angular
tourmaline grains are present in the fluvial and floodplain deposits in the
Hawkesbury Sandstone. Reworked sediments and granitic rocks are the
sources for the rounded and angular tourmaline grains, respectively (Standard,
1969). The fluvial channel facies in the Hawkesbury Sandstone also contains
rounded and angular zircon grains whereas the floodplain facies mainly only
contains rounded zircon grains. The rounded and angular zircon grains are
derived from sedimentary and acid igneous sources, respectively (Standard,
1969). Also, the presence of rounded to well rounded zircon grains in the
Hawkesbury Sandstone may indicate recycling and a metasedimentary source.
Standard (1969) showed that the rounded zircon in the minor amounts comes
from the southwestern source whereas the rounded zircon in larger amounts
comes from the western sources.

8-3 Linking depositional environment, diagenetic and


composition variation
8-3-1 Illawarra Coal Measures and Narrabeen Group
In the Illawarra Coal Measures and Narrabeen Group, quartz overgrowths are
recorded in the fluvial, floodplain, lagoonal and shallow marine facies (Appendix
6-1a, 6-11a, 6-19b, 6-21c, 6-22a, 6-24a, 7-1a, 7-2a, 7-3a, 7-5f, 7-6c, 7-7a and
f). They are more common in the fluvial facies than in the floodplain, lagoonal
and shallow marine facies which are characterised by thick and continuous clay
grain-coatings (cf. El-Ghali et al., 2009a). This correlates with the greater
abundance of intergranular pressure dissolution of quartz grains in the fluvial
facies compared to the floodplain, lagoonal and shallow marine facies
(Appendix 6-46a). Quartz cement is derived by dissolution from straight and
concavo-convex contacts between quartz grains in the fluvial facies (cf. Kim et
al., 2007). The influence of grain-coatings on quartz overgrowths can be
determined by a comparison between the Illawarra Coal Measures, Narrabeen
Group and Hawkesbury Sandstone, which indicates that the presence of
common grain-coatings reduces the amount of quartz overgrowths in the first
281

two units much more than in the Hawkesbury Sandstone. In addition, the total
quantity of detrital quartz in Illawarra Coal Measures and Narrabeen Group is
lower than in the Hawkesbury Sandstone and results in a lower quartz
overgrowth content due to the sparse sites for quartz dissolution and
precipitation as well as the abundance of grain-coating clays in these units
(Appendix 7-1a, 7-2a, e-f, 7-3a, e, 7-5e, 7-6c and 7-7a). Quartz grains become
more common in the Narrabeen Group than in the Illawarra Coal Measures,
which has more grain-coating clays. Thus, the abundance of quartz grains is
related to the number of quartz overgrowths in the Illawarra Coal Measures and
Narrabeen Group, as demonstrated by the positive correlation between quartz
grains and quartz overgrowths. This relationship is strongest in the Kembla
Sandstone (r2=0.7; Fig. 6-2a) and Lawrence Sandstone (r2=0.6; Fig. 6-1). X-ray
diffraction showed a high negative correlation between quartz and clay minerals
in the Illawarra Coal Measures (r2=0.5; Fig. 5-18) and Narrabeen Group (r2=0.8;
Fig. 5-35a), indicating that the abundance of clay reduces the percentage of
quartz grains in some units. Also, pseudomatrix is present in the fluvial,
floodplain, lagoonal and shallow marine facies of the Illawarra Coal Measures
and Narrabeen Group, which contributes to the reduction in the amount of
quartz overgrowths in these facies (Appendix 6-5a and 6-46b). Therefore, the
percentage of quartz grains in the Illawarra Coal Measures and Narrabeen
Group is also less frequent than in the Hawkesbury Sandstone due to the
presence of pseudomatrix, which is most common in the Illawarra Coal
Measures. Thus, there are fewer quartz grains in the Illawarra Coal Measures
than in the Narrabeen Group due to pseudomatrix. The sedimentation rate is
lower and the compaction induced subsidence is greater in the Illawarra Coal
Measures than the Narrabeen Group due to the environmental variations which
occur between the dominant floodplain to lagoonal deposits in the former and
the more prominent fluvial deposits in the latter (Sur et al., 2002).
In the Illawarra Coal Measures and Narrabeen Group, carbonate cement occurs
in all the fluvial, floodplain, lagoonal and shallow marine facies according to thin
section results and X-ray diffraction analysis (Appendix 5-7, 5-8, 5-15 and 5-16).
Pore-filling carbonate cement, grain-coating carbonate cement and carbonate
replacement of grains and matrix are most common in the fluvial facies, but also
282

occur in the floodplain, lagoonal and shallow marine facies (Appendix 6-11c, 612a-c, 6-13a-c, 6-14a-c, 6-15a-c, 6-16a-c, 6-17a-c, 6-18a, 6-19b-c, 6-22b-c, 623a-c, 6-24a-c, 6-25a-c, 6-26a-c, 6-27a-c, 6-28a-c, 6-29a-c, 6-30a, 31b, 7-2b-c,
7-5c, 7-6b). Medium- to coarse-grained sandstone samples are characterized
by pore-filling carbonate cement and grain-coating carbonate cement.
Conditions of semiarid climate are most appropriate for the formation of early
carbonate cement in fluvial facies (Morad et al., 2010) but this would not be the
case in the coal measures where plant life was abundant associated with
swamp conditions. The early pore-filling carbonate cement shows variations in
composition, indicating that depositional environments control the early
diagenesis in the Illawarra Coal Measures and Narrabeen Group as shown by
Kim et al. (2007; Appendix 6-11c and 6-25a). In lagoonal deposits, early porefilling carbonate cement is often subject to dissolution due to the presence of
CO2-bearing pore water and then it is re-precipitated associated with higher Fe
and Mn contents (Kim et al., 2007; Appendix 6-46c). Sedimentary environments
and diagenetic changes to pore water composition during burial can affect the
chemical composition of carbonate cemented sequences (Kim et al., 2007). Iron
contents in carbonate cement are related to concentrations of Fe2+ in the pore
water, but even high iron contents in the cement occur at low concentrations of
Fe2+ in the pore water (Kim et al., 2007). In the Illawarra Coal Measures and
Narrabeen Group, authigenic clay minerals contribute to the formation of late
carbonate minerals in fluvial, floodplain and lagoonal deposits which have thick
shale beds interbedded with the sandstone, and this is confirmed by the
ferromagnesian composition of later diagenetic carbonate minerals (Kim et al.,
2007).
A relationship between carbonate cement and detrital carbonate grains is
indicated where dissolution of detrital carbonate grains is recorded as the
source of carbonate cement (cf. Gier et al., 2008). The main influence of
carbonate cement on composition occurs through carbonate replacement.
Detrital grains such as feldspar and lithic grains may be partially or completely
replaced by carbonate cement in the Illawarra Coal Measures and Narrabeen
Group (Appendix 6-12c, 6-15b, 6-17a, c, 6-24c, 6-26a, 6-27a-b, 6-28a-c and 629a). This process affects the percentages of feldspar and lithic grains under
283

the microscope. Carbonate replacement is present in the Illawarra Coal


Measures and Narrabeen Group but is more common in the latter. This means
that the influence of carbonate replacement of feldspar and lithic grains is
stronger in the Narrabeen Group. Thus, the higher feldspar and lithic grain
content in the Illawarra Coal Measures than in the Narrabeen Group may due to
this process. Plagioclase albitization may be a source of calcium for the calcite
(Land, 1984; Lee and Lim, 2008). Carbonate cement is common in the Illawarra
Coal Measures and Narrabeen Group that are rich in volcanic rock fragments,
confirming that the latter may be a source of calcium for the carbonate cement
according to Moraes and De Ros (1990), Sur et al., (2002) and Lee and Lim
(2008).
Frohlich et al. (2010) showed that calcite cement often occurs in fluvial deposits.
In this study, it is observed in both fluvial and floodplain deposits whereas it
disappears in the lagoonal and shallow marine deposits (Appendix 6-16a-c, 617a-b, 6-27a-c and 6-28a-c). Calcite precipitation was affected by meteoric pore
waters due to many factors, such as the exposure of the fluvial sediments
during deposition (Frohlich et al., 2010). Determination of the origin of the early
calcite cement can depend on trace element composition of the groundwater
(Veizer, 1983; Kim et al., 2007). Strong porosity heterogeneity may occur in
fluvial deposits in the Illawarra Coal Measures and Narrabeen Group as result
of the presence of calcite cement (Morad et al., 2010). Fe and Mg contents are
high in calcite cement in marine environments, such as the Bargo Claystone
(Kim et al., 2007). Chemistry of marine water can be indicated by early porefilling calcite in the marine environments (Kim et al., 2007). Pore-filling calcite
cement initially formed from pore water of fresh-water composition, particularly
in the fluvial and floodplain deposits in these units (Kim et al., 2007; Appendix 616a-b and 6-27a-b). This was later followed by siderite cement as result of
reducing conditions, which are related to a high Fe2+/Ca2+ (Kim et al., 2007).
Suboxic or reducing environments are appropriate conditions of the formation of
early siderite cement (Hammer et al., 2010). Mesophilic and thermophilic ironreducing bacteria also contribute to siderite formation (Zhang et al., 2001; Karim
et al., 2010). Low-Mg calcite and pure siderite are included in the carbonate
cements that occur in fluvial and floodplain deposits, respectively (Mozley,
284

1989; Morad et al., 2000; Morad et al., 2010). Low Mg and Ca contents are
present in siderite intraclasts which can be sedimentologically proximal in origin
(Karim et al., 2010). In fluvial deposits, siderite cement post-dates calcite
cement because the redox conditions range from oxic to anoxic with increasing
burial (Kim et al., 2007). Bertier et al. (2008) showed that siderite precipitates
from reducing, non-sulphidic pore waters that evolve in the sub-oxic and
microbial methanogenesis zones of fluvial depositional environments. In
general, all recognized facies are characterized by siderite cement but it is more
common in the fluvial facies (Appendix 6-11c, 6-12a-c, 6-13a-b and 6-19b, 622b-c, 6-23a-c, 6-24a-c, 7-5c and 7-6b). Siderite cement is more abundant in
the non-marine environments than in marine environments because of the
response of Ca2+ associated with bicarbonate at normal marine concentrations
(Matsumoto and Lijima, 1981; Sager, 2007). The greater abundance of ankerite
cement in floodplain facies than in fluvial facies is clear, particularly in the
Illawarra Coal Measures. Pore-filling ankerite and grain-coating ankerite occurs
predominantly in the floodplain facies (Appendix 6-14a, c and 6-25a-c)
compared with the fluvial channel facies, which is characterised by ankerite
replacement (Appendix 6-15b-c and 6-26a-b). In the channel facies, dolomite is
formed prior to ankerite cement. In general, dolomite cement is minor in all
facies although pore-filling dolomite and grain-coating dolomite are more
common in the floodplain facies than in the fluvial channel facies (Appendix 613c, 6-17c, 6-18a, 6-26c, 6-29a-c, 6-30a and 31b).
Clay mineral abundance in most sediments is influenced by depositional facies
but they tend to be widespread in fluvial facies (Pay et al., 2000). In general,
fluvial deposits are characterized by the presence of grain coating clays (Morad
et al., 2000; Ketzer et al., 2003b; Worden and Morad, 2003; El-Ghali et al.,
2006a). In the Illawarra Coal Measures and Narrabeen Group, pore-filling clays
and grain-coating clays are common in the fluvial channel, floodplain, lagoonal
and shallow marine facies (Appendix 7-1a, c-d, 7-2c-f, 7-3a-f and 7-4a-f). They
are more common in the floodplain facies than in the channel, lagoonal and
shallow marine facies. Mixed-layer illite-smectite is prominent in all facies
whereas early diagenetic kaolinite is a characteristic component in the
floodplain facies, particularly in the Garie Formation according to X-ray
285

diffraction analysis (Appendix 5-21 and 5-22). The presence of kaolinite


indicates the impact of meteoric water on this facies (Karim et al., 2010).
Kaolinite is formed in appropriate conditions by the migration of meteoric waters
through sediments deposited in fluvial environments (Keller, 1970; Gorniak,
1997; Kim et al., 2007). Well-sorted coarse- and medium-grained sandstones
are characterized by the presence of kaolinite based on scanning electron
microscope analysis (cf. Abouessa and Morad, 2009; Appendix 7-2d, 7-5a-b
and 7-6e). A relationship between diagenetic alteration and composition
variation is present. Dissolution of feldspar is the main source of kaolinite in the
Illawarra Coal Measures and Narrabeen Group (Appendix 6-47a and b; cf.
Hammer et al., 2010). Kaolinite is more common in the Hawkesbury Sandstone
than in the Illawarra Coal Measures and Narrabeen Group based on X-ray
diffraction analysis and scanning electron microscope studies, although the
abundance of original feldspar grains in the Illawarra Coal Measures and
Narrabeen Group was probably much larger than in the Hawkesbury
Sandstone. This disagrees with Rossi et al. (2002) who showed that kaolinite
increases with increasing original feldspar grains. Volcanic rock fragments,
feldspar and matrix can play a role in the formation of early illite and chlorite
because their dissolution can lead to illitization of the matrix in the Illawarra Coal
Measures and Narrabeen Group (cf. Kim et al., 2007; Appendix 6-48b). Also,
the abundance of volcanic rock fragments and the presence of mica in the
Illawarra Coal Measures and Narrabeen Group can support the formation of
early chlorite in this unit as shown by Kim et al. (2007; Appendix 7-1a, c, 7-3a-b,
7-4c-f and 7-7b-d). Chlorite can be formed by addition of Fe2+, Mg2+, K+ and Na+
derived from the break-down of volcanic detritus and detrital mica in the
Illawarra Coal Measures and Narrabeen Group (Morad, 1990; Kim et al., 2007).
Sur et al. (2002) showed that the abundance of authigenic clay minerals is
related to the abundance of volcanic rock fragment in the Shimonoseki
Subgroup and chlorite is distributed through specific depositional environments
(Higgs et al., 2010).
In the Illawarra Coal Measures and Narrabeen Group, feldspar overgrowths
occur in the fluvial and floodplain facies and could indicate dry and warm
continental environments (Morad et al., 2000; Appendix 7-1f, 7-3d, 6-20c, 6-31c
286

and 6-32a), whereas they are absent in lagoonal and shallow marine facies.
Feldspar overgrowths are more common in the floodplain facies than in the
fluvial channel facies where mechanically infiltrated clays can reduce feldspar
overgrowths (cf. Wolela and Gierlowski-Kordesch, 2007). Near-surface
eodiagenesis by meteoric waters supports the abundance of K-feldspar
overgrowths in fluvial and floodplain facies (El-Ghali et al., 2009b). The nature
of the interstitial environment, associated with K, Al and Si in solution, is an
important factor in the crystallization of authigenic K-feldspar (Bertier et al.,
2008). Microfractures, cleavage planes and grain contact margins were
determined as sites for the occurrence of albitization of feldspar. The Illawarra
Coal Measures and Narrabeen Group are rich in volcanic rock fragments and
this supports the greater abundance of albitization in these units than in the
Hawkesbury Sandstone (cf. Sur et al., 2002). This correlation indicates the
relationship between diagenetic alteration and composition variation. Thus, in
the Illawarra Coal Measures the amount of albitization exceeds that in the
Narrabeen Group due to the more widespread of occurrence of volcanic rock
fragment in the Illawarra Coal Measures compared to the Narrabeen Group.
Also, the relationship between diagenetic alteration and composition variation is
shown through the influence of dissolution and replacement of detrital feldspar
in the Illawarra Coal Measures and Narrabeen Group (Milliken, 2004).
Plagioclase and K-feldspar both decrease due to dissolution and replacement in
these units (Appendix 6-12c, 6-21a, 6-28a, 6-32a, 6-33a, 6-41b-c, 6-44b and 645a). The influence of dissolution and replacement on detrital feldspar is greater
in the Narrabeen Group because of the abundance of dissolution remnants and
carbonate replacement, although carbonate cement is more common in the
Illawarra Coal Measures. Thus, alteration of feldspar, which is recorded in some
samples, may cause the rarity of K-feldspar in these units (Cowan, 1974).
Dissolution of unstable grains occurs in the fluvial, floodplain and lagoonal
facies in the Illawarra Coal Measures and Narrabeen Group. Dissolution of lithic
grains is more common in the fluvial facies due to the abundance of lithic grains
in this facies (Appendix 6-21b, 6-32b, 6-40b-c, 6-41a, 6-43c, 6-44a and c),
whereas dissolution of feldspar is more common in the floodplain facies
(Appendix 6-21a, 6-32a, 6-33a, 6-41b-c, 6-44b and 6-45a). In lagoonal facies,
287

feldspar dissolution is rare and lithic grain dissolution is absent. Carbonate


cement can be dissolved in the fluvial, floodplain and lagoonal facies and
dissolution is most apparent in the fluvial facies (Appendix 6-20a-b, 6-32c, 633b-c, 6-39c, 6-40a, 6-42a and 6-43a-b). Acidic pore water contributes to the
dissolution of unstable grains and carbonate cement, which, in turn, can support
the formation of authigenic minerals (Wolela and Gierlowski-Kordesch, 2007).
Dissolution affects the percentages feldspar and lithic grains in the Illawarra
Coal Measures and Narrabeen Group. This influence is greater in the
Narrabeen Group, thus this unit contains less feldspar and lithic grains than the
Illawarra Coal Measures. Thus, composition variation is strongly affected by
diagenetic alteration.
In the Illawarra Coal Measures and Narrabeen Group, the fluvial facies shows
more mechanical and chemical compaction than the floodplain and lagoonal
facies (Appendix 6-5a, 6-9b, 6-19a-b, 6-30c and 6-31a). The abundance of
pseudoplastic deformation of mud intraclasts and ductile grains in the fluvial
facies supports the occurrence of mechanical compaction in this facies. Also,
medium- to coarse-grained sandstone is rich in concavo-convex, sutured and
long grain contacts, indicating the presence of chemical compaction in the
fluvial facies (Appendix 6-19a-b and 6-31a). Mechanical and chemical
compaction also occurs in the floodplain and lagoonal facies but they are less
common. In the fluvial facies compaction pre-dates cementation, whereas in
floodplain and lagoonal facies compaction post-dates the earlier phases of
cementation. Diagenetic alteration, represented by compaction, has affected the
mineralogical composition in all facies. For example, the Illawarra Coal
Measures and Narrabeen Group are rich in ductile lithic grains, thus mechanical
compaction is higher in these units than in the Hawkesbury Sandstone (e.g.
Bjolykke et al., 1989; Worden and Burley, 2003; Hammer et al., 2010). Also, the
abundance of labile lithic grains in the Illawarra Coal Measures and Narrabeen
Group has enhanced mechanical compaction, as shown by Higgs et al. (2010).
Also, the fluvial channel facies is rich in pseudomatrix, particularly in the
medium- to coarse-grained sandstone, which represents altered lithic fragments
(cf. Bertier et al., 2008). This further illustrates the relationship between
composition and diagenetic alteration. Also, during early diagenesis, the
288

disintegration and kaolinitization of mud intraclasts can be effective in the


formation of pseudomatrix in the Illawarra Coal Measures and Narrabeen Group
(cf. Bertier et al., 2008), again indicating the influence of diagenesis on
composition. Pseudomatrix content is higher in the Illawarra Coal Measures and
Narrabeen Group than in the Hawkesbury Sandstone, thus mechanical
compaction is less in latter unit, and the sandstone composition is less affected
by diagenesis. The presence of bent and deformed mica flakes is evidence of
the importance of mechanical compaction in the Illawarra Coal Measures and
Narrabeen Group also indicating the relationship of diagenesis with composition
(Appendix 6-9b and 6-30c).

8-3-2 Hawkesbury Sandstone


Sandstone environments usually have quartz overgrowths that are mesogenetic
in origin and are closely related to intergranular dissolution (Worden and Morad,
2000; Morad et al, 2000; El-Ghali et al., 2006b). In the floodplain environment,
quartz overgrowths are less common than in the fluvial channel environment
because thick and continuous grain-coating clays are more common in the
floodplain environment (El-Ghali et al., 2009a; Appendix 6-34a, 7-8c-e, 7-10a-b,
d-e and 7-11a). The presence of thick and continuous grain-coating clays often
controls how widespread quartz overgrowths are in the floodplain environment.
Also, pseudomatrix plays a role in the inhibition of quartz overgrowths in
floodplain environments. In contrast, the fluvial channel environment is
characterized by thin and discontinuous clay grain coatings. This contributes to
the development of intergranular pressure dissolution of quartz grains, and thus
quartz overgrowths are distributed throughout this environment. Reed et al.
(2005) showed a similar development of quartz cement to be the result of weak
grain-coatings in the Appalachian Carboniferous sandstones. Also, chemical
compaction can be inhibited in lower energy areas with more abundant graincoating clays and pseudomatrix in the fluvial channel environment, thus quartz
overgrowths show a widespread but patchy developed in this environment (ElGhali et al., 2009a). The abundance of quartz overgrowths in the fluvial channel
facies, which is the main environment in the Hawkesbury Sandstone, indicates
289

the presence of much more abundant quartz in this unit than in the Narrabeen
Group and Illawarra Coal Measures. This abundance of quartz in the
Hawkesbury Sandstone and the small quantity of ductile lithic grains results in
the sparse occurrence of thick and continuous grain-coating clays in the fluvial
facies. Thus, the presence of quartz overgrowths is controlled by the
abundance of detrital quartz grains and large pores, as indicated by the high
positive correlation between detrital quartz and quartz overgrowths (r = 0.8;
Fig. 6-6a). This relationship is evidence of the influence of diagenesis on
composition and indicates that silica dissolution is the source of the quartz
cement. This may indicate that dissolved silica is derived partly from detrital
grains, and it is precipitated in a nearby part of the rock unit (Blatt, 1979; Kim et
al., 2007).
In this study, grain coating clays are recorded in both the fluvial and floodplain
environments but are more common in the latter (Appendix 7-8c-e, 7-9a, 7-10ab, d-e and 7-11a). According to X-ray diffraction and scanning electron
microscope analysis, kaolinite and illite occur in greater abundance in the finegrained floodplain deposits rather than in the coarse-grained fluvial channel
sandstone (Appendix 5-17). These variations indicate the environmental
influence on sandstone composition. Kaolinite may have originated from Kfeldspar in the fluvial environment (Appendix 6-5c), as demonstrated by the
negative correlation between them (cf. Hammer et al., 2010). This confirms the
influence of diagenetic alteration on sandstone composition where kaolinite
increases in the Hawkesbury Sandstone as result of feldspar alteration. The
presence of kaolinite in the form of discrete masses also suggests that detrital
feldspar may be the source of the kaolinite (cf. Pay et al., 2000, Appendix 7-8a
and b). Rossi et al. (2002) showed also the importance of feldspar dissolution in
the formation of kaolinite in the Triassic TAGI sandstone in the Berkine Basin in
Algeria. Also, Frohlich et al. (2010) indicated that the influence of meteoric
diagenesis on sandstones can enhance the presence of kaolinite. The
abundance of quartz grains may prevent the widespread development of clay
minerals in the Hawkesbury Sandstone. This is evidenced by the good negative
correlation between quartz and clay minerals (r = 0.7; Fig. 5-44). Under humid
climate conditions, the source area may have suffered strong and prolonged
290

chemical weathering that would account for the abundance of kaolinite in the
Hawkesbury Sandstone (Ghazi and Mountney, 2011). Widespread kaolinite,
with or without illite, in the Hawkesbury Sandstone is evidence that this unit was
sourced from a sedimentary origin in continental conditions (Akarish and ElGohary, 2008; Appendix 7-8a-e, 7-9a-b, d-e, 7-10a-c and 7-11c). Pore-filling
illite and chlorite are more characteristic in the floodplain environment
(Appendix 7-9c and 7-10f).
In general, carbonate cement is rare in the Hawkesbury Sandstone but when it
is present it occurs as pore-filling siderite and ankerite cement that is more
common in the floodplain environment compared with the fluvial channel
environment (Appendix 6-35a-b, c and 6-36a-b). Carbonate replacement of
detrital grains is completely absent in both the fluvial and floodplain
environments. This means that carbonate cement has had no influence on
sandstone composition in the Hawkesbury Sandstone.
In the fluvial facies, feldspar overgrowths were observed in coarse-grained
sandstone by thin section and scanning electron microscope studies, whereas
the overgrowths disappear in floodplain samples (Appendix 6-5c and 7-11e).
Also, the presence of mechanically infiltrated clays may reduce the occurrence
of K-feldspar overgrowths in the channel facies in the Hawkesbury Sandstone
(cf. Wolela and Gierlowski-Kordesch, 2007). This may indicate that the
percentages of feldspar grains in the floodplain facies were higher than in the
channel facies and they are also fresher, indicating the influence of diagenesis
on composition.
The occurrence of pseudomatrix in fine-grained sandstone, siltstone and shale
in the floodplain facies (Appendix 6-48c) exceeds its occurrence in the fluvial
channel facies and it results from the mechanical compaction of mud intraclasts.
This confirms that diagenesis, represented by mechanical compaction, supplies
pseudomatrix and, thus, it can affect composition. Much less pseudomatrix is
observed in the Hawkesbury Sandstone than in the Narrabeen Group and
Illawarra Coal Measures, which include more mechanical compaction of mud
intraclasts and thus sandstone composition is related to diagenesis. Also,
mechanical compaction is less in the Hawkesbury Sandstone due to the
291

abundance of quartz grains (e.g. Bjolykke et al., 1989; Worden and Burley,
2003; Hammer et al., 2010). This illustrates the relationship between diagenesis
and sandstone composition. Sandstone rich in quartz has less mechanical
compaction as indicated by the presence of undeformed mica flakes that occur
in this sandstone (cf. Abouessa and Morad, 2009; Appendix 6-37c). In general,
less mechanical compaction occurs in the Hawkesbury Sandstone than in the
Illawarra Coal Measures and Narrabeen Group, which include much more
pseudomatrix and less quartz grains.

8-4 The influence of diagenetic


composition on reservoir quality

alteration

and

Diagenetic alteration that controls the reservoir quality in the southern Sydney
Basin basically includes compaction, carbonate cement, authigenic clay
minerals, dissolution and quartz overgrowths (Salem et al., 2000; Al-Ramadan
et al., 2004; Salem et al., 2005; Zhang et al., 2007; Zhang et al., 2008; Luo et
al., 2009; Islam, 2009; Jia-liang et al., 2010).

8-4-1 Illawarra Coal Measures


8-4-1-1 Compaction
Compaction is the main reason for porosity loss in the Illawarra Coal Measures.
Chemical and mechanical compaction are both moderate to high and play a
dominant role in the porosity reduction in the Illawarra Coal Measures (cf. Kilda
and Friis, 2002). Mechanical compaction is recorded between framework grains
and is represented by deformation of ductile grains (mica and rock fragments)
and mud intraclasts as the main factors that contribute to the porosity reduction
(Appendix 6-5a and 6-9b).
In sandstone rich in detrital lithic fragments, the widespread presence of ductile
grains in most samples indicates that most porosity was reduced by mechanical
compaction as shown by Buyukutku and Sahinturk (2006) and Molenaar et al.
(2007). Thus, pore space is not commonly available for quartz precipitation. So,
the influence of mechanical compaction on porosity reduction is strengthened
292

by the presence of ductile grains in the Illawarra Coal Measures (e.g. Pittman
and Larese, 1991; Buyukutku and Sahinturk, 2006; Ehrenberg et al., 2008).
Also, the common occurrence of mica and mudrock fragments in the Illawarra
Coal Measures supports the influence of compaction to reduce porosity (cf. Luo
et al., 2009; Appendix 6-9b). Pore spaces were partly or completely filled by
pseudomatrix, which is formed by mechanical compaction and has ability to
reduce porosity in the Illawarra Coal Measures (El-Ghali et al., 2009a; Appendix
6-5a). Its importance is stronger in sandstones that have few remaining lithic
grains (cf. Al-Ramadan et al., 2004). This illustrates the influence of sandstone
composition on thin section porosity measurements in the Illawarra Coal
Measures. Mansurbeg et al. (2009) showed that mechanical compaction is an
important cause of porosity loss, supported by the presence of bent mica flakes,
carbonate grains and plastic deformation of ductile sedimentary grains. Modern
alteration and extension of detrital mica, especially in outcrop samples, is
another significant factor in the partial porosity loss in the Illawarra Coal
Measures, as shown by Hlal (2008). Also, detrital mica contributes to the partial
porosity loss through the development of intergranular pressure dissolution of
quartz grains (cf. Hlal, 2008). Mechanical compaction reduces porosity during
early diagenesis and had more influence in samples with low contents of early
carbonate cement in various units of the Illawarra Coal Measures (cf. Umar et
al., 2011). This indicates that the reservoir quality was reduced due to
mechanical compaction during early burial in the studied succession. For
example, the Loddon Sandstone contains rare carbonate cement (Appendix 513 and 5-14), thus mechanical compaction had more influence on porosity
whereas the influence of compaction is less apparent with the more abundant
carbonate cement in the Darkes Forest Sandstone (Appendix 5-13 and 5-14).
Mechanical compaction strongly affects porosity, particularly in the Kembla
Sandstone and Loddon Sandstone, which are dominated by clays with very low
quartz contents (cf. Ehrenberg et al., 2008). Grain rearrangement is also
important to reduce porosity where lithic grains are absent, particularly in
quartzose samples (Worden et al., 2000). Mechanical compaction also reduces
early secondary porosity particularly before quartz cementation.

293

The occurrence of long, concavo-convex and sutured grain contacts shows the
influence of chemical compaction on porosity reduction in the Illawarra Coal
Measures (Appendix 6-19a-b). The interpenetration of grains provides important
evidence of porosity reduction by chemical dissolution in the Illawarra Coal
Measures (Appendix 6-49a). Also, chemical compaction is more common in the
absence of carbonate cement in the reduction of porosity (cf. Kim and Lee,
2004).
In general, the porosity that is reduced by mechanical and chemical compaction
cannot be recovered, as shown by Wolela (2009). In some samples, secondary
porosity is completely absent which may also be due to compaction. This
indicates that the secondary porosity in these samples was created before
compaction had ended and then it became compacted, thus it disappeared (cf.
Kim and Lee, 2004). The influence of compaction on porosity is stronger in
coarse-grained and moderate sorted sandstone. Also, a long burial period and
the abundance of lithic grains in the Kembla Sandstone and Loddon Sandstone
are important factors that enhanced the influence of compaction to reduce
porosity (Li et al., 2002).
In the Wilton Formation, Bargo Claystone and Darkes Forest Sandstone the
influence of compaction on reservoir quality is rare because of the abundance
of early diagenetic carbonate cement in these units (cf. Maraschin et al., 2004).
Thus, this confirmed that porosity is mainly reduced by compaction of poorly
cemented units. The proportion of porosity reduction under the influence
compaction during burial can be determined by the ratio of lithic to quartzose
grains (Buyukutku and Sahinturk, 2006).
8-4-1-2 Carbonate cementation
After the influence of compaction on porosity, carbonate cement is the second
main factor causing porosity loss in the Illawarra Coal Measures (Baker et al.,
2000; Ketzer et al., 2003a; Xiaomin et al., 2004). In general, carbonate cement
is the main type of porosity loss that may be recoverable as result of dissolution
of carbonate cement as shown by Wolela (2009). The reservoirs that include no
or only a little carbonate cement are the best in the Illawarra Coal Measures.
294

The relationship between thin section porosity and carbonate cement is


negative but negligible (Fig. 7-3a). This relationship indicates that as carbonate
cement increases, porosity decreases. The relationship showed that in one
sample the porosity equals 3.5% and the carbonate cement equals 25.3%
whereas in another sample when the porosity equals 9.5%, carbonate cement
equals 2.2% (Fig. 7-3a).
In the coarse-grained deposits, carbonate cement has caused a significant
decrease in the porosity and it is also significant in the secondary porosity
increase, via dissolution, in the Kembla Sandstone, Lawrence Sandstone and
Loddon Sandstone. Thin section porosity may also be destroyed by eogenetic
carbonate cement, which is observed as a pore-filling cement in the coarsegrained deposits (e.g. El-Ghali et al., 2009b; Umar et al., 2011). Ankerite
cement is distributed in both large and small pores and is the most important
carbonate cement in the porosity loss in these coarse deposits (Appendix 6-11a
and 6-14a-c). Also, porosity is lost in the coarse-grained deposits because of
the occurrence of dolomite and calcite cements (cf. Jia and Gu, 2002; El-Ghali
et al., 2009a, b; Appendix 6-13c, 6-16a and 6-18a) and eogenetic siderite (cf.
Bertier et al., 2008; Appendix 6-11c, 6-12a and 6-13a) in pores spaces. This
indicates that the abundance of carbonate cement prevents the development of
porosity in the coarse-grained deposits as shown by Gier et al. (2008). Also
early carbonate cement is a barrier to the development of quartz overgrowths
(cf. Umar et al., 2011) and intense mechanical compaction (cf. Rossi et al.,
2001) in the coarse-grained deposits. When widespread the early carbonate
cement completely fills both large and small pores and reduces porosity,
whereas if there is only minor early carbonate cement it can still support the
framework grains against compaction.
In some samples from the Kembla Sandstone, Lawrence Sandstone and
Loddon Sandstone, low porosity is preserved by early carbonate cement which
is significant in inhibiting the development of early mechanical compaction. In
later

diagenesis,

dissolution

of

carbonate

cement

contributes

to

the

development of reservoir quality when and where it creates the low quantities of
secondary porosity in these units (e.g. Zhang et al., 2008; Chi et al., 2003;
295

Zhang et al., 2007; Appendix 6-20a-b, 6-39c and 6-40a). However, where early
cementation was abundant and almost completely occluded primary porosity,
later dissolution was compensated by re-precipitation of calcite cement and no
new porosity was created. New secondary porosity was generated if the
dissolution of this early calcite cement was not compensated by new cement
formation. Similar results were observed in many studies, for example, the
study by Chi et al. (2003) in the Late Carboniferous strata in the Maritimes
Basin. In other cases, carbonate grains are effective in the development or
destruction of porosity. They may be dissolved and result in secondary porosity
or, in other cases, they are dissolved and then re-precipitation as carbonate
cement (cf. Hlal, 2008).
In some samples pore-filling carbonate cement completely reduces porosity in
the Illawarra Coal Measure. In other samples, open pore space can be
observed as a result of partial carbonate cementation or later dissolution of
carbonate cement (cf. Zhang et al., 2008; Appendix 6-20a-b, 6-39c and 6-40a).
In pores with incomplete filling, low porosity can also be observed in some
samples as shown by Al-Ramadan et al. (2004; Appendix 6-39c). Late
diagenetic pore-filling carbonate cement is minor in some samples and in this
case it has little influence on reservoir quality.
In fine-grained deposits, early carbonate cement is present as a pore-filling
cement that reduces porosity in these deposits (cf. Frohlich et al., 2010; AlRamadan et al., 2004; Appendix 6-16b). In these deposits, porosity loss is not
recoverable by dissolution of carbonate cement.
8-4-1-3 Authigenic clay minerals
Authigenic clay minerals affect thin section porosity but are of less influence
than compaction and carbonate cement in the Illawarra Coal Measures. In
coarse-grained deposits, represented by the Kembla Sandstone, Lawrence
Sandstone and Loddon Sandstone, authigenic clay minerals, including mixed
layer illite/smectite, kaolinite, illite and chlorite, have influenced the thin-section
porosity. They were observed by scanning electron microscope as pore-filling,
pore-bridging and pore coating clays, thus they decrease the porosity in these
296

units. In general, pore-filling and pore-lining clays prevent the preservation or


development of porosity in these units (cf. Wolela and Gierlowski-Kordesch,
2007). Fibrous mixed layer illite/smectite is recorded as a pore-filling cement,
reducing primary porosity in these units (Appendix 7-1d). Vermicular booklets of
kaolinite crystals are also present in primary and secondary pores and also
reduce both primary and secondary porosity (cf. Borgohain et al., 2010;
Appendix 7-2d). Primary porosity also decreases in these units as a result of
pseudomatrix, which is generated by deformation of ductile lithic grains, or the
development of late diagenetic kaolinite and compaction in some samples.
Chlorite is observed to fill primary pores and causes a porosity loss in these
units (e.g. Zhang et al., 2007; Al-Ramadan et al., 2004; Appendix 7-1c). It also
fills intergranular secondary pores leading to a porosity reduction. On the other
hand, Illite exists as a pore-filling cement (Al-Ramadan et al., 2004; Mckinley et
al., 2011) and as pore-bridges between grains; consequently it can reduce
porosity in the sandstone (Appendix 7-3f). In general, pore-filling mixed-layer
illite/smectite and pore-filling chlorite cause a greater reduction of porosity in
coarse-grained deposits than pore-filling kaolinite and pore-filling illite. This
porosity reduction by pore-filling clays is unrecoverable in the Illawarra Coal
Measures.
Many studies have showed the role of clay-coatings in the prevention of the
development of quartz overgrowths, consequently this process helps in the
preservation of porosity in sandstone samples (Heald and Larces, 1974;
Thomson, 1979; Pittman et al., 1992; Ehrenberg, 1993; Bloch et al., 2002;
Anjos et al., 2003; Thomas et al., 2004; Taylor et al., 2004; Salem et al., 2000,
2005; Zhang et al., 2008; Smosna and Sager, 2008; Taylor et al., 2010; Morad
et al., 2010). Also, grain-coating clay can help preserve the porosity developed
by intergranular quartz dissolution and chemical compaction (Oelkers et al.,
1996; Salem et al., 2000).
In coarse-grained deposits and according to scanning electron microscope
studies, chlorite, illite and mixed layer illite/smectite are present as grain
coatings and prevent the development of quartz overgrowths, thus minor
primary porosity is preserved in these deposits. In many samples, scanning
297

electron microscope analysis showed grain-coating mixed-layer illite/smectite as


a barrier to repel the growth of quartz overgrowths, thus retaining low porosity
(Appendix 7-2e and 7-3e). Also, the development of quartz overgrowths around
quartz grains is prevented by thick grain-coating chlorite, which preserves low
porosity as result of this process in coarse-grained deposits (e.g. Al-Ramadan
et al., 2004; Salem et al., 2005; Zhang et al., 2008; Armitage et al., 2010;
Appendix 7-1a, 7-3a and 7-4c). Continuous layers are created by grain-coating
chlorite on the surface of detrital grains adjacent to intergranular pore spaces,
thus this process prevents the growth of quartz overgrowths and preserves
porosity (Appendix 7-1a). Primary porosity can also be preserved by graincoating illite, which also prevents the development of quartz overgrowths in
coarse-grained deposits (Moraes and De Ros, 1990; Bloch et al., 2002;
Abouessa and Morad, 2009; Appendix 7-2a). Grain-coating mixed-layer
illite/smectite and grain-coating chlorite have more influence on quartz
overgrowths than secondary kaolinite and illite in the coarse-grained deposits.
Thus, they are more important in the preservation of porosity. Also, quartz
cement is influenced more by authigenic clay coatings than by detrital clay rims.
Authigenic clays coating are more continuous than detrital clay rims that are
present as a result of mechanical infiltration (Taylor et al., 2010). In general,
quartz overgrowths increase with small quantities of grain-coating clays and is
reduced by more abundant grain-coating clays. Grain-coating clays are more
important in the preservation of porosity, particularly in the presence of
carbonate cement and quartz overgrowths, which prevent the development of
compaction (cf. Salem et al., 2000). Microporosity is present in coarse-grained
deposits due to kaolinitization of feldspar (e.g. Bertier et al., 2008; Abouessa
and Morad, 2009; Appendix 6-47c). Dutton and Loucks (2009) showed that
kaolinite fills pores but forms common microporosity in the Wilcox Sandstone.
In fine-grained deposits, authigenic clay minerals are dominant and prevent the
preservation or growth of secondary porosity. Kaolinite, illite, chlorite and mixedlayer illite/smectite are distributed through the pores and cause a deterioration
of porosity in these deposits. The presence of grain-coating clays generally
does not preserve porosity in fine-grained deposits (Zhang et al., 2010). This is
because they are thin and discontinuous or because they are very rare, thus
298

they were not able to inhibit diagenetic alteration, such quartz overgrowths.
Thus, the influence of grain-coating clays on quartz overgrowths is weak in
these samples.
8-4-1-4 Dissolution
Dissolution of grains is recorded as a mid to late stage diagenetic process that
generates and preserves minor secondary porosity during burial of the Illawarra
Coal Measures. This process has less influence on porosity than compaction;
carbonate cement and pore-filling clay minerals, but it has more influence on
the measured porosity than grain-coating clays. In the Kembla Sandstone,
Lawrence Sandstone and Loddon Sandstone, secondary porosity is observed
where the abundance of lithic grains and the presence of feldspars allow the
development of secondary porosity. Thin sections showing partial to complete
dissolution of detrital grains, such as feldspar and rock fragments, supports the
generation and preservation of minor secondary porosity in these sandstones
(cf. Ketzer et al., 2003a; Al-Ramadan et al., 2005; Islam, 2009; Appendix 6-21b,
6-40b-c and 6-41a-c). Feldspar grains were partly dissolved in some samples
resulting in minor secondary porosity that is preserved in these samples (cf. Luo
et al., 2009; Appendix 6-41b-c). This provides evidence that the dissolution of
detrital feldspar is significant in the enhancement of secondary porosity in the
Illawarra Coal Measures (e.g. Lima and De Ros, 2002; Rossi et al., 2002;
Ketzer et al., 2003a; Dutton and Loucks, 2009; El-Ghali et al., 2009a; Wolela,
2009). Secondary porosity, which formed by partial dissolution of coarse
plagioclase grains, is higher than that produced by fine plagioclase grains in the
Illawarra Coal Measures. Also, secondary porosity, created by dissolution of
volcanic rock fragments is more than by dissolution of sedimentary rock
fragments because the former dominate the latter. Also the glassy volcanic
detritus is much more unstable than the quartz and clays forming the
sedimentary clasts. Here, the role of composition has a major influence on
porosity generation. The presence of feldspar and lithic grains can enhance
porosity through their dissolution. Also, dissolution of carbonate cement is a
mechanism to increase porosity in the studied succession (cf. Chi et al., 2003)
and can be formed after quartz overgrowth precipitation in the Illawarra Coal
299

Measures (Rossi et al., 2001; Appendix 6-20a-b, 6-39c and 6-40a). In some
samples, large sections of open, well-connected secondary intergranular pores
can be produced by the dissolution of calcite cement (cf. Mansurbeg et al.,
2008).
Wolela and Gierlowski-Kordesch (2007) showed that feldspar and carbonate
cements were dissolved in the Hartford Basin, Newark Supergroup, USA, and
this generated most of the secondary porosity in this area. Zhang et al. (2010)
showed that the Shahejie Formation in the Huimin depression in eastern China
has secondary porosity as result of dissolution of unstable feldspar and lithic
grains as well as dissolution of carbonate cement.
In the Illawarra Coal Measures, dissolution of detrital grains, such as feldspar
and lithic grains, is more important in the formation of secondary porosity than
dissolution of cement (Giles, 1987; dos Anjos et al., 2000). This indicates that
the formation of secondary porosity by dissolution of detrital grains is
volumetrically subordinate in the Illawarra Coal Measures. However, the
determination of secondary porosity generated by framework grain dissolution is
easier than recognising secondary porosity resulting from pore-filling cement
dissolution (Taylor et al., 2010). In some samples, dissolution of feldspar did not
result in secondary porosity due to the precipitation of kaolinite in the resultant
pores (Shanmugam, 1988; Appendix 6-47a). This means that kaolinite may be
formed to fill secondary pores created by dissolution of feldspar (cf. Bertier et
al., 2008). Also, dissolution of feldspar may result in secondary pores that can
also be filled with carbonate cement as shown by Umar et al. (2011). The latter
may have prevented the development of porosity in the Illawarra Coal
Measures. Another reason is that mechanical compaction also effectively
causes the collapse of the secondary porosity after dissolution (Wilkinson et al.,
2001; Dutton and Loucks, 2009). The abundance of mechanical compaction
may indicate that secondary porosity was more common than primary porosity.
However, if only the inside of the feldspar grain is subjected to solution porosity
(cf. Zhang et al., 2007) this may indicate why feldspar dissolution is not effective
in the development of porosity in some cases. In some samples, secondary
pores that result from dissolution of polycrystalline quartz are filled by carbonate
300

cement. Thus, they are also not effective in the development of porosity
(Appendix 6-49b).
Organic acids, CO2 and dehydration of clay minerals also contribute to the
dissolution of grains and cements and the generation of secondary porosity
(Xiaomin et al., 2004; Zhang et al., 2007). The maturation of organic matter is
the source of organic acids and CO2 (Chi et al., 2003; Zhang et al., 2007). Also,
Chi et al. (2003) showed that reactions between inorganic minerals and
meteoric water contribute to the dissolution of grains and cements and
generation of secondary porosity. Unstable grains are dissolved by percolation
of gravity-driven acidic meteoric water and groundwater, forming secondary
porosity (Wolela and Gierlowski-Kordesch, 2007). Also, meteoric water has a
main role in the dissolution of carbonate cement to generate secondary porosity
(Wang, 1992; Chi et al., 2003). Meteoric groundwater supports the presence of
dissolution at shallow depths (Bjorlykke, 1984; Mathisen, 1984; Umar et al.,
2011) whereas basin fluids cause dissolution at greater depths (Seibert et al.,
1984; Surdam et al., 1984; Umar et al., 2011).
During early diagenesis, detrital grains and cement may be dissolved by
meteoric waters (Morad et al., 2000; Ketzer et al., 2002; Mansurbeg et al.,
2008). During mid diagenesis, such grains and cements are dissolved due to
organic acids in the basin waters (Surdam et al., 1984; Wilkinson and
Haszeldine, 1996; Mansurbeg et al., 2008).
8-4-1-5 Quartz overgrowths
In the Illawarra Coal Measures, quartz overgrowths only have a minor influence
on primary porosity because of the abundance of ductile lithic grains and
carbonate cement that control porosity. Also, quartz grains are not abundant in
the Illawarra Coal Measures, thus there is little opportunity for quartz
overgrowth development. This suggests that the influence of quartz overgrowth
on porosity is low. Minor quartz cementation occurs in some pores and is not a
significant mechanism for the porosity loss. In the Kembla Sandstone,
Lawrence Sandstone and Loddon Sandstone, scanning electron microscope
studies showed that grain-coatings prevented quartz overgrowth development
301

(Appendix 7-1a, 7-2a e, 7-3a, e and 7-4c). The inhibition of quartz overgrowths
means that they have almost no influence on the preservation of primary
porosity (cf. Salem et al., 2005; Zhang et al., 2007).

8-4-2 Narrabeen Group


8-4-2-1 Compaction
In the Narrabeen Group, compaction has the greatest influence on thin section
porosity. Thus, porosity loss in the Narrabeen Group is high as result of both
mechanical (e.g. Higgs et al., 2007; Dutton and Loucks, 2009) and chemical
compaction (cf. Higgs et al., 2007). The poor sorting and common detrital clays
contribute to the porosity loss through mechanical compaction according Ramm
(1992).
In sandstone rich in detrital lithic fragments, chemical and mechanical
compaction is observed and reduces porosity (cf. Tobin et al., 2010). In these
samples, deformation of ductile grains and mica between quartz grains
illustrates porosity reduction by mechanical compaction (Appendix 6-30c).
Deformation of lithic fragments is common in the Narrabeen Group, resulting
from mechanical compaction, and plays a major role in porosity loss (cf.
Buyukutku and Sahinturk, 2005; Appendix 6-49c). Ductile grains are common in
all units and they account for the low porosity in most samples, as shown by
Gier et al. (2008). Alteration, extension of detrital mica and intergranular
pressure dissolution of quartz grains causes additional porosity loss in the
Narrabeen Group (cf. Hlal, 2008; Appendix 6-30c and 6-46a). The effect of
mechanical compaction on porosity occurs during early burial and is more
effective in the absence or rarity of carbonate cement in some samples of the
Coalcliff Sandstone and Scarborough Sandstone as shown by Umar et al.
(2011). In some units of the Narrabeen Group, clays predominate over quartz
grains and porosity is almost entirely lost as result of mechanical compaction
(cf. Ehrenberg et al., 2008). These factors indicate a strong relationship
between sandstone composition and porosity. Chemical compaction is also
significant to reduce porosity through the interpenetration of grains in the lithic302

rich sandstone (Appendix 6-50a). The preservation of porosity is reduced by


long, concavo-convex and sutured contacts that represent chemical compaction
(Appendix 6-31a). In samples with little carbonate cement, chemical compaction
continues to obliterate porosity and is more effective (cf. Kim and Lee, 2004). In
the Narrabeen Group, secondary porosity was developed in some sandstone
samples before compaction ended. Any secondary pores that developed were
compacted, thus secondary porosity was absent in thin section because of
compaction (Kim and Lee, 2004). The amount of porosity reduced by
mechanical and chemical compaction in the Narrabeen Group is similar to that
in the Illawarra Coal Measures and these losses are not unrecoverable (cf.
Wolela, 2009). Pseudomatrix occurs as result of mechanical compaction and
fills both small and large pores, thus further reducing porosity as shown by ElGhali et al. (2009a; Appendix 6-46b). Pseudomatrix is often more important for
porosity reduction where lithic grains are absent in the Narrabeen Group (cf. AlRamadan et al., 2004). Grain rearrangement is also an important porosity
reduction factor in the absence of lithic grains, particularly in quartzose samples
(Worden et al., 2000). Porosity in fine-grained deposits is mainly reduced by
mechanical compaction (e.g. Mansurbeg et al., 2008).
In the sandstone rich in detrital quartz, pore spaces are preserved because
mechanical compaction is prevented by the detrital grain framework. This
suggests that the influence of mechanical compaction on porosity was less in
this type of sample (Abouessa and Morad, 2009). Thus, in the Bulgo
Sandstone, the effect of compaction on primary porosity is less due to the
quartz-rich nature of parts of the unit, as shown by Smosna and Sager (2008).
This unit is characterised by a low abundance of ductile grains, thus compaction
is inhibited due to rigid grains which support the stable framework grains,
preserving primary porosity (cf. Sager, 2007). Thus, about 12% of primary
porosity occurs in this unit.
In the Narrabeen Group, thin section porosity is slightly less affected by
mechanical compaction than in the Illawarra Coal Measures. This is due to the
greater abundance of ductile grains and pseudomatrix in the Illawarra Coal
Measures compared with the Narrabeen Group.
303

8-4-2-2 Carbonate cementation


The influence of carbonate cement on thin section porosity is recorded as the
second main reason after the influence of compaction in the Narrabeen Group.
There is a very low negative correlation between carbonate cement and
porosity in the Narrabeen Group (Fig. 7-7a). Thus, porosity is high with low
carbonate cement whereas it is low with high carbonate cement. The
relationship between porosity and carbonate cement in the samples shows that
when the porosity is less than 2% carbonate cement is more than 20%,
whereas when the porosity is more than 15%, carbonate cement does not
exceed 5% (Fig. 7-7a). This suggests that carbonate cement significantly
affects both sandstone rich in detrital lithic grains and sandstone rich in detrital
quartz (e.g. Wolela and Gierlowski-Kordesch, 2007; Luo et al., 2009; Hammer
et al., 2010).
In sandstone rich in detrital lithic fragments and sandstone rich in detrital quartz,
early carbonate cement fills both large and small pores and reduces the primary
porosity, as shown by Umar et al. (2011). The widespread early carbonate
cement may completely fill the pores, but even minor early carbonate cement
can support the framework grains against compaction. All types of carbonate
cement, including ankerite, siderite, dolomite and calcite, are observed by
microscope as pore-filling cement, reducing primary porosity in the Narrabeen
Group (Appendix 6-22b-c, 6-23c, 6-25a, c, 6-26c and 6-29b). Ankerite cement
has the most influence on thin section porosity, followed by siderite cement
(Appendix 6-22b-c, 6-23c, 6-25a and c). In the Scarborough Sandstone,
carbonate cement shows a moderate negative correlation with thin section
porosity whereas there is only a low negative correlation between carbonate
cement and porosity in the Bulgo Sandstone (Figs 7-8b and 7-9). These
relationships indicate that porosity decreases with increasing carbonate cement.
In a similar study, El-Ghali et al. (2009a) showed that carbonate cement,
represented by calcite and dolomite, destroy porosity in the Early Triassic
Petrohan Terrigenous Group in northwest Bulgaria. In the present study,
porosity in sandstone rich in detrital lithic grains is influenced more by carbonate
cement than sandstone rich in detrital quartz.
304

In the Narrabeen Group, porosity loss due to carbonate cement may be


recoverable in both lithic-rich and quartzose sandstone. Early carbonate cement
contributes to the preservation of porosity by: a) preventing the development of
quartz cement and early mechanical compaction. In this case the framework is
supported against mechanical compaction by early carbonate cement that also
prevents porosity loss due to quartz cement. b) Carbonate cement may be
dissolved during later diagenesis, forming secondary porosity (e.g. Chi et al.,
2003; Zhang et al., 2007; Appendix 6-32c, 6-33b-c, 6-42c and 6-43a-b).
However, where early cementation was abundant and almost completely
occluded primary porosity, later dissolution was compensated by reprecipitation of calcite and no new porosity was created. New porosity was only
generated if the dissolution of this early calcite cement was not compensated by
new cement formation.
Also, incomplete pore-filling by carbonate cement supports the preservation of
porosity in some samples (Appendix 6-50b). Later dissolution of carbonate
cement can be observed and result in carbonate cement partly filling pores
(Appendix 6-32c, 6-33b-c, 6-42c and 6-43a-b). This process was often seen in
sandstone rich in detrital quartz. In other cases, carbonate grains can be
effective in the preservation or destruction of porosity they may dissolve and
result in secondary porosity (Appendix 6-50c) or they are dissolved and then reprecipitation as carbonate cement (cf. Hlal, 2008).
Carbonate cement reduces more porosity in the Illawarra Coal Measures than
in the Narrabeen Group because it is more common in the former. On the other
hand, more carbonate cement has been dissolved in the Narrabeen Group than
in the Illawarra Coal Measures, thus it provides more secondary porosity in the
Narrabeen Group than in the Illawarra Coal Measures.
8-4-2-3 Authigenic clay minerals
In the Narrabeen Group, porosity is commonly reduced by authigenic clay
minerals (cf. Zhang et al., 2008) and such porosity is unrecoverable in this
succession. Although, the influence of authigenic clays on porosity is effective, it
had less influence than compaction and carbonate cement. Scanning electron
305

microscope studies showed the types of pore-filling authigenic clay minerals


which reduce primary porosity in the Narrabeen Group (cf. Wolela and
Gierlowski-Kordesch, 2007). Vermicular booklets of kaolinite occur as porefillings and cause significant porosity loss (Appendix 7-5a-b and 7-6e).
Deformed kaolinite occurs as pseudomatrix that effectively decreases porosity
in the Narrabeen Group. Chlorite has been observed by scanning electron
microscope in both primary and secondary pores and it also causes loss of
porosity in the Narrabeen Group (Appendix 7-7b). Also, porosity is decreased
by pore-filling illite and mixed-layer illite-smectite in sandstones units (cf. AlRamadan et al., 2004; Appendix 7-6f and 7-6d). Pore-filling illite and mixedlayer illite/smectite have a greater influence on porosity loss in lithic-rich
sandstone whereas pore-filling kaolinite and chlorite cause the greatest porosity
loss in quartz-rich sandstone. This comparison depends on the results of
scanning electron microscope and X-ray diffraction analyses.
Grain-coating clays, represented by chlorite, illite and mixed layer illite/smectite,
help to preserve porosity in the Narrabeen Group in both lithic-rich and quartzrich sandstone. In similar studies, Salem et al. (2005) showed that porosity is
preserved in the Abu Madi Gas reservoirs in the Nile Delta Basin in Egypt
because of the role of grain-coating clays that contribute to the inhibition of the
growth of quartz overgrowths.
In the present study, grain-coating clays effectively inhibit the precipitation of
quartz overgrowth cement, thus low porosity may be preserved in parts of the
Narrabeen Group. Mixed layer illite-smectite is observed as a grain-coating on
quartz overgrowths in many samples (Appendix 7-5f and 7-6c). This indicates
that low porosity is still present in the Narrabeen Group. Also, small quantities of
grain-coating chlorite inhibits quartz overgrowths and increases the percentage
of preserved porosity in the Narrabeen Group (e.g. Baker et al., 2000; Salem et
al., 2005; Umar et al., 2011; Appendix 7-5e). Grain-coating illite clays are
present in the Narrabeen Group around quartz overgrowths (Appendix 7-7a).
Thus, they contribute to the preservation of porosity in these deposits (cf.
Storvoll et al., 2002; Mansurbeg et al., 2008; Abouessa and Morad, 2009). This
supports the abundance of quartz overgrowths in sandstone that has low grain306

coating clays. Grain-coating mixed layer illite-smectite plays the greatest role in
the preservation of porosity because it is more common than grain-coatings illite
or chlorite. Also, the illite-smectite coatings are thicker and more continuous
than other grain-coating clays. They have the ability to prevent the growth of
quartz cement during burial and, thus, retain porosity (e.g. Bloch et al., 2002;
Islam, 2009; Luo et al., 2009). No grain-coating kaolinite was observed around
quartz overgrowths in this succession. In sandstones, the opposite influence
occurs between clay coatings and porosity when associated with an absence of
compaction. In some units, particularly quartz-rich sandstone, grain-coating
clays that occur as thin or discontinuous layers are not significant in the
prevention of development of quartz overgrowths. Thus, porosity was not
observed in these units. The development of quartz cement is less affected by
detrital clay rims than by authigenic clay coatings which are more continuous in
the Narrabeen Group (Taylor et al., 2010). Detrital clay rims are generated by
mechanical infiltration (Taylor et al., 2010).
Scanning electron microscope studies showed that the influence of pore-filling
clay minerals in reducing porosity in the Narrabeen Group is similar to those in
the Illawarra Coal Measures. More grain-coating clays are seen in the
Narrabeen Group than in the Illawarra Coal Measures, thus they contribute
more to the preservation of porosity in the Narrabeen Group than in the
Illawarra Coal Measures.
8-4-2-4 Dissolution
In the Narrabeen Group, feldspar, rock fragments and carbonate cement are
dissolved in some samples during mid to late diagenesis as a mechanism
leading to the enhancement of porosity (cf. Chi et al., 2003). Secondary porosity
occurs in all units of the Narrabeen Group as a diagenetic feature resulting from
dissolved grains. Dissolution of feldspar is supported by the presence of
secondary porosity in the Narrabeen Group, particularly in the Coalcliff
Sandstone (e.g. Baker et al., 2000; Salem et al., 2000; Lima and De Ros, 2002;
El-Ghali et al., 2006a; Zhang et al., 2007; Zhang et al., 2008; Abouessa and
Morad, 2009; Hammer et al., 2010; Appendix 6-33a, 6-44b-c and 6-45a). Also,
Luo et al. (2009) showed a similar influence of dissolution of plagioclase in the
307

enhancement of secondary porosity in the Ordos Basin in China. In the present


study, plagioclase grains were more likely to be dissolved than K-feldspar
grains, thus the former is more effective in forming secondary porosity in the
Narrabeen Group. Lithic-rich sandstone, represented by the Coalcliff Sandstone
and Scarborough Sandstone, shows that lithic grain dissolution results in
secondary porosity (cf. Ketzer et al., 2003a; Islam, 2009; Appendix 6-32b, 643c, 6-44a and c). In the Bulgo Sandstone, most of the lithic grains are
dissolved, thus it contains more secondary porosity (Appendix 6-51a). Volcanic
rock fragments are more abundant and less stable than sedimentary rock
fragments, thus they are more importance in the formation of secondary
porosity. Dissolution of unstable grains is volumetrically subordinate in the
Narrabeen Group, similar to the Illawarra Coal Measures, indicating the role of
composition to support porosity by dissolution of feldspar and lithic detrital
grains. Carbonate cement is also dissolved in the Narrabeen Group forming
secondary porosity, particularly in the Coalcliff Sandstone and Wombarra
Claystone (cf. Al-Ramadan et al., 2004; Zhang et al., 2010; Appendix 6-32c, 633b-c, 6-42c and 6-43a-b). However, dissolution of unstable grains is more
common and importance than dissolution of carbonate cement in the Narrabeen
Group. The development of secondary porosity post-dates quartz overgrowths
and carbonate cement in the Narrabeen Group and is indicated by the presence
of oversized pores and corroded grain boundaries, including corrosion of quartz
overgrowths and carbonate cement. Determining the amount of secondary
porosity that formed by dissolution of carbonate cement is difficult in some
samples. On the contrary, it is easy to identify secondary porosity formed by
dissolved unstable grains.
In some cases, the dissolution of grains such as feldspar may not contribute to
the development of much porosity in the Narrabeen Group for two reasons.
Firstly, the solution porosity may occur inside the feldspar grain forming minor
intragranular porosity (cf. Zhang et al., 2007; Appendix 6-33a). Secondly,
secondary pores generated by dissolution of feldspars may be filled by
carbonate cement (cf. Umar et al., 2011; Appendix 6-51b) and/or authigenic
kaolinite (cf. Bertier et al., 2008; Appendix 6-47b) in some samples. These
reasons are enough to prevent the development of much secondary porosity
308

that is formed by dissolution of feldspar grains. Also, mechanical compaction


may cause the collapse of the secondary porosity after dissolution in the
Narrabeen Group, similar to the Illawarra Coal Measures (Wilkinson et al.,
2001; Dutton and Loucks, 2009).
Several factors play a role in the dissolution of grains and cements that form the
secondary porosity. These are organic acids, CO2 and dehydration of clay
minerals (Zhang et al., 2007) in addition to inorganic mineral and meteoric water
reactions (Chi et al., 2003). Dissolution may occur as a result of the original
pore fluids but is more commonly associated with acidic ground water invasion
(e.g. Al-Hajri et al., 1999; Al-Ramadan et al., 2004). Pore fluid is also important
in the dissolution of carbonate cement, such as calcite cement (McBride et al.,
1996).
The importance of dissolution of grains and cements in the production of
secondary porosity in the Narrabeen Group is greater than in the Illawarra Coal
Measures, even though lithic and feldspars grains are more common in the
Illawarra Coal Measures. This may be due to more abundant circulation of
meteoric water in the Narrabeen Group (e.g. Mansurbeg et al., 2008).
8-4-2-5 Quartz overgrowths
In this unit, thin section porosity is reduced by quartz overgrowths but they are
rare. In fact, in the Narrabeen Group a very low positive correlation is recorded
between quartz overgrowths and thin section porosity (Fig. 7-6b). In lithic-rich
sandstone, the abundance of carbonate cement and ductile grains reduces the
influence of quartz overgrowths on porosity. Quartz overgrowths mainly occur in
cleaner sandstone and hence they have a moderate positive correlation with
thin section porosity in the Scarborough Sandstone (Fig. 7-7b). This means that
when quartz overgrowths are present, they may reduce further physical and
chemical compaction and preserve more primary porosity in this unit. Thus, the
influence of compaction on porosity may be reduced by minor quartz
overgrowths, preserving porosity in some samples (cf. Abouessa and Morad,
2009). In the Scarborough Sandstone, the results showed that when the
porosity of a sandstone sample is 19.5% quartz overgrowths only form 3.7% of
309

the rock, whereas when the porosity is zero quartz overgrowths form less than
1.7% of the sample (Fig. 7-7b). Also, in lithic-rich sandstone, quartz
overgrowths have a low negative correlation with porosity. Here minor pore
spaces are filled with quartz overgrowths further reducing porosity (cf. Molenaar
et al., 2007). Grain-coating clays around detrital quartz grains are common in
the lithic-rich sandstone, as indicated by scanning electron microscope studies
(Appendix 7-6c). This process prevents the development of quartz overgrowths,
thus some primary porosity is preserved (e.g. Islam, 2009; Salem et al., 2005).
In quartz-rich sandstone, quartz overgrowths contribute to the process of
destroying thin section porosity (cf. Islam, 2009; Appendix 6-51c). They are
present as a pore-filling cement that reduces primary porosity. Also, early
carbonate cements are uncommon in these samples; consequently quartz
overgrowths effectively decrease primary porosity. Also, quartz cement inhibits
further compaction and fills pore space between grains, thus porosity is reduced
in quartz-rich sandstone (cf. Zhang et al., 2010). A negative correlation is
recorded between quartz cement and porosity in the Bulgo Sandstone (r2 = 0.2;
Fig. 7-9). This relationship shows that porosity decreases with the presence of
more quartz cement. Primary porosity is more influenced by quartz overgrowths
due to the lack of ductile grains and carbonate cement in quartz-rich sandstone.
In these samples, grain coating clays are also present and can control the
evolution of quartz overgrowths, preserving porosity particularly in the Bulgo
Sandstone (cf. Islam, 2009; Salem et al., 2005; Appendix 7-7a).

8-4-3 Hawkesbury Sandstone


8-4-3-1 Compaction
Lundegard (1992) showed that compaction is more important than cementation
in the loss of primary porosity in the average sandstone. Previous studies
showed that porosity decreased from 40% to 34% because of the
rearrangement of grains in well sorted quartz-rich sandstone (Lundegard, 1992;
Wilson and Stanton, 1994) or to 26% (Paxton et al., 2002; Molenaar et al.,
2007).
310

In the Hawkesbury Sandstone, compaction dominates the porosity reduction


during the early diagenetic phase (Frohlich et al., 2010). This porosity is
unrecoverable (cf. Wolela, 2009). Also, long, concavo-convex and sutured
contacts produced by pressure solution and chemical compaction are observed
commonly in the Hawkesbury Sandstone and contribute to primary porosity loss
(Appendix 6-37b). The role of chemical compaction in primary porosity
reduction occurs during mid to late diagenesis. Carbonate cement is rare to
absent in most of the Hawkesbury Sandstone, thus the role of chemical
compaction to reduce primary porosity is significant, as shown by Kim and Lee
(2004).
Worden et al (2000) showed that grain rearrangement and dissolution support
the impact of compaction on porosity in quartzose sandstones. In the present
study, the Hawkesbury Sandstone is a quartz-rich arenite and, thus grain
rearrangement is an important factor in reducing porosity.
Mechanical compaction occurs in the Hawkesbury Sandstone but had less
influence on porosity than chemical compaction. This is because of the lack of
ductile grains and grain deformation in the Hawkesbury Sandstone. Thus, the
presence of rigid grains plays a role in obstructing the importance of mechanical
compaction since rigid grains increase the stability of the framework and thus
contributes to the preservation of primary porosity as shown by Sager (2007).
Thus some primary porosity was preserved in the Hawkesbury Sandstone.
Also, the moderately open framework packing is evidence that primary porosity
was not greatly affected by mechanical compaction in this unit (cf. Tamar-Agha,
2009). These factors indicate that the remaining primary porosity was preserved
after mechanical compaction and it decreased with the occurrence of
cementation.
The influence of compaction on reservoir quality occurs more in siltstone than in
sandstone in the Hawkesbury Sandstone and is stronger in well sorted beds. It
is present at all depths.
In general, porosity was reduced more by mechanical compaction in the
Illawarra Coal Measures and Narrabeen Group than in the Hawkesbury
311

Sandstone as result of the abundance of ductile grains in the former units.


Chemical compaction is more prevalent in the Hawkesbury Sandstone
compared with the Narrabeen Group and Illawarra Coal Measures. This
indicates that chemical compaction had a greater influence on porosity in the
Hawkesbury Sandstone than in the Narrabeen Group and Illawarra Coal
Measures.
8-4-3-2 Quartz overgrowths
Deposition of quartz overgrowths is recognised in many studies as a
mechanism for porosity loss in sandstones (Worden et al., 2000; Baker et al.,
2000; Al-Ramadan et al., 2004; Xiaomin et al., 2004; Molenaar et al., 2007;
Zhang et al., 2007; Mansurbeg et al., 2008; Luo et al., 2009; Islam, 2009;
Armitage et al., 2010). In quartzose sandstone, porosity is strongly influenced
by the amount of quartz dissolution and cementation (Appendix 6-34a). This
influence occurs during mid diagenesis at temperatures of 80C or more
(Ehrenberg, 1990; Bjorkum and Nadeau, 1998; Bjorkum et al., 1998;
Ehrenberg, 2008). The porosity reduction resulting from precipitation of
authigenic quartz is associated with increased temperature in several studies
(e.g. Giles et al., 2000; Walderhaug, 2000; Taylor et al., 2010). According to
Walderhaug (2000), the role of temperature and the presence of clean quartz
grain surfaces as sites for quartz overgrowth development are two important
factors that can determine the amount of porosity loss under the influence of
quartz cementation.
In the Hawkesbury Sandstone, there is a very low positive correlation between
quartz overgrowths and porosity (Fig. 7-12a). This relationship indicates that
porosity is present when the pores are clean enough for quartz overgrowths.
Porosity can be observed at the termination points of quartz overgrowth
cements in this unit (Reed et al., 2005; Appendix 6-34a). Primary porosity is
preserved in the Hawkesbury Sandstone by the mechanism of quartz
cementation, which inhibits collapse of the grain framework (Wolela, 2009).
Quartz overgrowths in some samples of Hawkesbury Sandstone help cement
the framework and obliterate chemical compaction. In these cases, primary
porosity was preserved in these samples, supported by quartz overgrowths (cf.
312

El-Ghali et al., 2006b; Appendix 6-34a). However, the relationship between


quartz overgrowths and secondary porosity also shows a very low positive
value in the Hawkesbury Sandstone, indicating that secondary porosity can
increase with increasing quartz overgrowths in this porous sandstone (Fig. 713). This could suggest that quartz overgrowths post-date the dissolution of
carbonate cement (e.g. Forgotson et al., 2000). Porosity gain from feldspar
dissolution could also add to the low positive correlation between quartz cement
and secondary porosity in the Hawkesbury Sandstone (Bertier et al., 2008). In
contrast, there is a very low negative correlation between quartz cement and
porosity in the Illawarra Coal Measures (Fig. 7-2b). This difference in correlation
between quartz cement and porosity in the Hawkesbury Sandstone and the
Illawarra Coal Measures could suggest that the source of quartz cement for
these units is different (cf. Bertier et al., 2008).
Quartz cement is a mid to late stage diagenetic mineral that was derived from
pressure solution leading to the reduction of porosity. This indicates continuous
compaction in the sediment at least until the time of quartz cementation. In
general, quartz overgrowths are observed as pore-filling cements that destroy
porosity in the Hawkesbury Sandstone (cf. Kilda and Friis, 2002; Appendix 634a). This unit is poor in ductile grains and carbonate cement. Thus, quartz
cement is the main reason for porosity loss in the Hawkesbury Sandstone
(Worden et al., 2000). Quartz grains are very abundant in the Hawkesbury
Sandstone, leading to common quartz cement that reduces primary porosity.
This relationship indicates that composition has a significant impact on porosity.
Also, the cleanest sandstones are observed in the Hawkesbury Sandstone, thus
quartz cementation leads to more loss of porosity in these samples (Worden et
al., 2000). The rarity of grain-coating clay in the Hawkesbury Sandstone
indicates clearly that the development of quartz overgrowths continued after the
clay precipitation and caused the main reduction of porosity in this unit (cf. AlRamadan et al., 2004; Zhang et al., 2007). In a few samples, clay minerals such
as chlorite exist as grain-coatings around quartz grains (Appendix 7-11a). In
these cases, the clays effectively prevent the precipitation of quartz overgrowths
and support the preservation of porosity (cf. Worden and Morad, 2003; AlRamadan et al., 2004; Zhang et al., 2007; Bloch et al., 2002; Luo et al., 2009;
313

Islam, 2009; Salem et al., 2005). In a different study, Cocker et al. (2003)
showed that porosity was preserved in the Jauf reservoir in the Hawiyah Field
as a result of the inhibition of quartz overgrowth precipitation by thick illite grain
coatings. In the Hawkesbury Sandstone, the development sites for quartz
overgrowths are subject to obliteration by grain-coating illite, thus this process
can support the preservation of porosity in this unit (cf. Bloch et al., 2002;
Abouessa and Morad, 2009; Appendix 7-10e). Several studies have shown that
quartz overgrowths are prevented by the presence of microcrystalline quartz
rims on detrital quartz grains (Aase et al., 1996; Bloch et al., 2002; Ramm et al.,
1997; Jahren and Ramm, 2000; Aase and Walderhaug, 2005; Lima and De
Ros, 2002). In a few samples of Hawkesbury Sandstone, microcrystalline quartz
rims are important and contribute to the development of reservoir quality where
they prevent the growth of quartz overgrowths and pressure dissolution
(Appendix 7-12a-c). The distinction of microcrystalline quartz coatings in thin
section is not easy due to the very small crystal size whereas their distinction by
scanning electron microscope is easy. Low abundance of quartz overgrowths
and high porosity occur clearly in sandstone with abundant microcrystalline
quartz coatings.
Quartz overgrowths are a more effective cause of porosity loss in the
Hawkesbury Sandstone than in the Narrabeen Group and Illawarra Coal
Measures.
8-4-3-3 Authigenic clay minerals
Authigenic clay minerals occur as pore-filling cement and are one of the main
reasons for porosity reduction in the Hawkesbury Sandstone. Grain-coating
clays also support the occurrence of primary porosity in the Hawkesbury
Sandstone. Pore-filling clays and grain-coating clays are shown by scanning
electron microscope.
Vermicular booklets of kaolinite crystals are the most common authigenic clay
minerals and fill both primary and secondary pores, reducing total porosity in
the Hawkesbury Sandstone (Appendix 7-8b-d and 7-9a-b). In some samples,
kaolinite is deformed and generates pseudomatrix that effectively decreases
314

porosity. Well crystalline kaolinite is usually associated with common


microporosity between the kaolinite plates (Appendix 7-8a). Pore-filling illite is
present and prevents the developed porosity in the Hawkesbury Sandstone
(Appendix 7-9c). Pore-filling chlorite is rare but also reduces porosity in the
Hawkesbury Sandstone (cf. Wolela and Gierlowski-Kordesch, 2007; Appendix
7-10f). These porosity losses are unrecoverable in the Hawkesbury Sandstone.
A comparison between the influences of pore-filling kaolinite, pore-filling illite
and pore-filling chlorite on porosity indicates that pore-filling kaolinite has the
greatest influence, followed by pore-filling illite and then pore-filling chlorite,
which has the least influence. Many studies have shown the influence of porefilling clays on porosity. For example, Al-Ramadan et al. (2004) showed that
porosity in the Devonian Jauf Sandstone at Ghawar Field in eastern Saudi
Arabia was reduced as a result of illite and chlorite which fill pores.
The growth of grain-coating clays contributes to the prevention of quartz
overgrowth development, thus retaining primary porosity in the Hawkesbury
Sandstone (Zhang et al., 2008; Taylor et al., 2010; Morad et al., 2010). Quartz
cementation is affected by grain-coating kaolinite, grain-coating illite and graincoating chlorite which preserve primary porosity in the Hawkesbury Sandstone
(Appendix 7-8c-e, 7-10a-b, d-e and 7-11a). Grain-coating kaolinite is more
common, thus its influence is stronger on quartz overgrowths (Appendix 7-8c, e
and 7-10a-b). Thus grain-coating kaolinite contributes significantly to the
preservation of primary porosity in the Hawkesbury Sandstone. Also, graincoating illite around quartz overgrowths is observed, thus it retains primary
porosity in the Hawkesbury Sandstone (cf. Abouessa and Morad, 2009;
Appendix 7-8d and 7-10d-e). The influence of grain-coating kaolinite and illite
on quartz overgrowths is stronger than grain-coating chlorite because they
occur as thick and continuous layers between detrital grains and intergranular
pore space. These features support the importance of grain-coating kaolinite
and illite to inhibit quartz overgrowth and preserve primary porosity in the
Hawkesbury Sandstone. Clay coatings affect porosity and are associated with
rare compaction in some samples. The influence of patchy detrital clay rims on
quartz overgrowth precipitation is much less than the influence of authigenic
clays coatings which are continuous (Taylor et al., 2010). Thin or discontinuous
315

grain-coatings, which do not affect quartz overgrowths and do not support


porosity preservation in the Hawkesbury Sandstone, is similar to that in the
Illawarra Coal Measures and Narrabeen Group.
8-4-3-4 Carbonate cementation
In the Hawkesbury Sandstone, carbonate cement is a rare component in some
samples, thus its influence on reservoir quality is low or absent. In this unit,
porosity has a negative correlation with carbonate cements (r2 = 0.2; Fig. 712b). Pore-filling carbonate cement, represented by siderite and ankerite
cement, occurs as an early diagenetic mineral and contributes to the porosity
reduction in the Hawkesbury Sandstone (Appendix 6-35a-c and 6-36a-b).
8-4-3-5 Dissolution
In the Hawkesbury Sandstone, dissolution of the rare unstable grains, such as
feldspars and lithic grains, is not significant whereas dissolution of carbonate
cement is absent. This indicates that secondary porosity is uncommon in the
Hawkesbury Sandstone, which is characterised by primary porosity. In the
Hawkesbury Sandstone, compaction is the main mechanism in the reduction of
secondary porosity (Wilkinson et al., 2001; Dutton and Loucks, 2009).
Carbonate cement is rare in the Hawkesbury Sandstone, thus compaction has
more influence on secondary porosity. The presence of carbonate cement
prevents compaction. Dissolution of feldspar is low but when it is present, it
does not enhance porosity in the Hawkesbury Sandstone because it is
incomplete (Appendix 6-5c).

8-5 The influence of compaction and cementation on


porosity and permeability
Many studies have used the Lundegard (1992) method to assess the porosity
loss by compaction and cementation, e.g. Lima and De Ros (2002), Ahmad and
Bhat (2006) and Zhang et al. (2008).

316

Lundegard (1992) method suggests that:


COPL = compaction porosity loss.
= Pi (((100 Pi) Pmc)/ (100 Pmc))).
CEPL = cementation porosity loss.
= (Pi COPL) (C/Pmc).
Where
Pi = initial porosity (45%).
Pmc = minus cement porosity.
C = volume percent of pore-filling cement.

8-5-1 Illawarra Coal Measures


Compaction and cementation commonly control porosity preservation in the
Illawarra Coal Measures. A plot of compaction porosity loss versus cementation
porosity loss showed that compaction had a greater influence on porosity than
cementation. Thus, porosity is reduced by compaction more than by
cementation in most samples in this sequence.
According to the methodology of Lundegard (1992), compaction porosity loss
ranges between 0% and 44.2% (Appendix 5-29; Fig. 8-1) at average of 32.6%
(Table 8-1), whereas cementation porosity loss ranges from 0.8% to 45%
(Appendix 5-29; Fig. 8-1) at average of 11.5% (Table 8-1) in the Illawarra Coal
Measures. This indicates that an average of 32.6% of the porosity was reduced
by compaction whereas an average of 11.5% of the porosity was reduced by
cementation.
y=0.989x+44.018
R=0.9758

COPL

50.0
45.0
40.0
35.0
30.0
25.0
20.0
15.0
10.0
5.0
0.0
0.0

10.0

20.0 CEPL 30.0

40.0

50.0

Figure 8-1: Plot shows compaction porosity loss (COPL) versus


cementation porosity loss (CEPL) in the Illawarra Coal Measures.
317

8-5-2 Narrabeen Group


Compaction and cementation have contributed to the reduction of porosity in
the Narrabeen Group. The relationship between compaction porosity loss and
cementation porosity loss indicated that the role of compaction in the reduction
of porosity is the greatest in most samples.
In the Narrabeen Group, compaction porosity loss ranges between 0% and
41.8% (Appendix 5-30; Fig. 8-2) at average of 32.2% (Table 8-1), whereas
cementation porosity loss ranges from 1.5% to 45% (Appendix 5-30; Fig. 8-2) at
average of 9.1% (Table 8-1). Thus compaction has destroyed 32.2% of the
porosity whereas cementation destroyed 9.1% of the porosity.

y=0.9182x+40.552
R=0.7394

45.0
40.0
35.0

COPL

30.0
25.0
20.0
15.0
10.0
5.0
0.0
0.0

10.0

20.0

CEPL

30.0

40.0

50.0

Figure 8-2: Plot shows compaction porosity loss (COPL) versus


cementation porosity loss (CEPL) in the Narrabeen Group.

8-5-3 Hawkesbury Sandstone


The importance of compaction and cementation in the influence on porosity is
observed in the Hawkesbury Sandstone. The technique of Lundegard (1992)
showed that the highest percentage of porosity was destroyed by compaction in
this unit in all samples.
318

In the Hawkesbury Sandstone, compaction induced porosity loss is recorded


between 23.1% and 40.9% (Appendix 5-31; Fig. 8-3) at average of 31.7%
(Table 8-1), whereas cementation porosity loss ranges from 2.2% to 18.6%
(Appendix 5-31; Fig. 8-3) at average of 7.4% (Table 8-1). This is evidence that
31.7% of the porosity was lost due to compaction whereas 7.4% of the porosity
was lost due to cementation.
In general, compaction and cementation play a main role in the reduction of
porosity in the Illawarra Coal Measures, Narrabeen Group and Hawkesbury
Sandstone. Compaction had the most influence on porosity, according to the
technique of Lundegard (1992), in all three units.

45.0

y=0.5792x+35.98
R=0.1456

40.0
35.0

COPL

30.0
25.0
20.0
15.0
10.0
5.0
0.0
0.0

5.0

10.0
CEPL

15.0

20.0

Figure 8-3: Plot shows compaction porosity loss (COPL) versus


cementation porosity loss (CEPL) in the Hawkesbury Sandstone.

COPL
CEPL

ICM
32.6
11.5

NBG
32.1
9.1

HBSS
31.7
7.4

Table 8-1: Averages of compaction porosity loss (COPL) and


cementation porosity loss (CEPL) in the Illawarra Coal Measures,
Narrabeen Group and Hawkesbury Sandstone.

319

320

Chapter Nine
9-1 Reservoir Potential of the Southern Sydney Basin
Porosity and permeability were measured for the Illawarra Coal Measures,
Narrabeen Group and Hawkesbury Sandstone. Porosity and permeability are
important to determine groundwater and petroleum reservoirs. Porosity and
permeability, together with structure, control the potential for gas in these units.
Extractable groundwater and gas may be present in the units which contain
high porosity and permeability whereas it would be unavailable in units which
contain low porosity and permeability. Also, potential oil and gas sources will be
shown in this part.

9-1-1 Illawarra Coal Measures


9-1-1-1 Wilton Formation
Primary thin section porosity is absent in the Wilton Formation as result of
compaction and cementation. Secondary porosity has not developed as result
of the lack of dissolution. Carbonate cement is present infilling pores and also
authigenic clays such as kaolinite reduce the porosity and permeability. Most
samples in the Wilton Formation are fine-grained deposits formed in a floodplain
environment. Fine-grained deposits have low porosity and permeability, thus
they are unlikely to contain extractable water or gas. This supports the role of
facies and environment to control the possible occurrence of gas in the
formation. The results show that the Wilton Formation is a poor reservoir but
provides a good seal potential. Also, well logs showed that claystone forms
good seals in the Wilton Formation. Carbonaceous rocks are recorded in the
Wilton Formation, thus it is a potential source for petroleum.
9-1-1-2 Bargo Claystone
Shale is widespread in the Bargo Claystone. One sample of siltstone was
described in the Bargo Claystone. Primary thin section porosity is absent in this
sample due to the influence of compaction and cementation. Ankerite cement,
siderite cement, mixed-layer illite-smectite and kaolinite are effective in the
321

prevention of the development of secondary porosity. Dissolution is mainly


absent in this formation, thus secondary porosity is absent. These results show
that the Bargo Claystone, like the Wilton Formation, represents a poor
groundwater or gas reservoir and it also contains claystone seals. Common
shale and siltstone are recorded in the Bargo Claystone, thus porosity is low.
This indicates that gas is not available in the Bargo Claystone. These results
indicate the importance of diagenesis and facies in the inhibition of gas. These
factors reduce porosity, thus they hinder the presence of gas. Also, well logs did
not indicate the presence of porosity in the Bargo Claystone. This formation
may contain hydrocarbons which are sourced from carbonaceous rocks.
9-1-1-3 Darkes Forest Sandstone
Most samples are fine-grained deposits such as fine-grained sandstone,
siltstone and shale. Thus, thin section porosity is not present in this unit and
permeability is only 0.01 md. Also, compaction, carbonate cement and
authigenic clays have affected the thin section porosity. Ankerite, calcite and
siderite cements and mixed layer illite-smectite have destroyed thin section
porosity in the Darkes Forest Sandstone. These factors all reduce porosity and
permeability in the Darkes Forest Sandstone. Thus, this formation is similar to
the Bargo Claystone and Wilton Formation which are poor reservoirs and good
seals. Also, the well logs showed that the Darkes Forest Sandstone is a good
seal. The Darkes Forest Sandstone is not a source for hydrocarbons since it
does not include carbonaceous shale.
9-1-1-4 Allans Creek Formation
No thin section porosity has been determined in the Allans Creek Formation
whether primary or secondary porosity. Also, the presence of compaction and
cementation has destroyed primary porosity whereas the absence of dissolution
prevents the development of secondary porosity. This indicates that
groundwater and gas would not be available in the Allans Creek Formation.
Fine-grained deposits are common in this formation and they are also moderate
sorted. The Allans Creek Formation represents a poor reservoir but would be a
good seal due to its low porosity. Well logs showed that fine-grained deposits
322

are good seals in the Allans Creek Formation. The Allans Creek Formation
could contain oil or gas because of the presence of carbonaceous rocks which
are a potential source for hydrocarbons.

9-1-1-5 Kembla Sandstone


In the Kembla Sandstone, thin section porosity ranges from 0 to 4.7% at
average of 1.1% whereas permeability does not exceed 0.01 md. Compaction
and cementation have had a major influence on reservoir quality in the Kembla
Sandstone. Large and small pores are filled by carbonate cement. Also, mixed
layer illite-smectite, chlorite and kaolinite are observed by SEM as pore-filling
cements in the Kembla Sandstone. These factors have a large affect on
porosity and permeability. Thus, the Kembla Sandstone is a poor reservoir but
may contain gas in some areas.

a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

b)

Fig. 9-1: a) and b) Porosity in the Kembla Sandstone

The lower Kembla Sandstone consists of channel deposits which are porous,
thus gas or water can found in this formation. The environment of deposition in
the Kembla Sandstone is not different from the environments of many other
typical fluvial groundwater and hydrocarbon reservoirs. In the Kembla
Sandstone, medium- to coarse-grained sandstone is porous and permeable,
thus it contains groundwater and possibly gas. This is shown in well logs
between 420.9 m and 422.8 m (Well 126, Fig. 9-1a). Also, well logs showed the
323

presence of porosity in conglomerate samples between 390.5 m and 391.3 m


(Well 125, Fig. 9-1b). In Well 127, porosity is observed between 371.8 m and
379.5 m according well logs (Fig. 9-2). Also, the upper Kembla Sandstone
contains siltstone and shale which would act as seals according thin section
and core logging. In general, the Kembla Sandstone is recorded as a potential
groundwater and gas reservoir.
Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 9-2: Porosity in the Kembla Sandstone


For example, sample (25) in well 124 (394.08-394.11 m, Fig. 9-3) is a coarsegrained sandstone and includes 4.7% thin section porosity. This sample is also
porous according to the well logs. Sample (50) in well 30 (878.63 878.66 m) is
siltstone and did not contain porosity according to well logs and thin sections.
Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 9-3: Porosity in the Kembla Sandstone

Carbonaceous shale is absent from the lower Kembla Sandstone which is


characterized by medium- and coarse-grained sandstone. This indicates that
the lower Kembla Sandstone is not a potential source for hydrocarbons.
The Kembla Sandstone is underlain by the Allans Creek Coal Member. This
coal seam, like the other coals in the southern Sydney Basin, contains a large
quantity of dissolved gas which is why gas drainage is needed before the coal
324

seams are mined. This coal has a well developed cleat that would allow
migration of both water and gas to the Kembla Sandstone.
9-1-1-6 Eckersley Formation
9-1-1-6-1 Unnamed Member Three
In the Unnamed Member Three, thin sections indicated that porosity is absent
because of compaction and cementation. Pore-filling carbonate cement is
present and prevents the preservation of primary porosity. Authigenic clays
such as mixed layer illite-smectite and illite are also present as pore-fillings and
grain-coatings, thus they also reduce porosity. This indicates that carbonate
cement and authigenic clays reduce the potential for groundwater or gas in the
Unnamed Member Three.
Fine-grained deposits are common in the Unnamed Member Three which is a
poor reservoir but would act as good seal. Also, well logs showed that these
fine-grained deposits are good seals. Also, gas may be found in the Unnamed
Member Three because of the occurrence of carbonaceous shale in this
formation.
9-1-1-6-2 Unnamed Member Two
Compaction and cementation also play a role in the destruction of thin section
porosity in the Unnamed Member Two. Thus, they contribute to a low reservoir
potential in the Unnamed Member Two. Carbonate cement, represented by
ankerite, siderite and dolomite, fills pores between detrital grains, thus it
reduces porosity. Mixed layer illite-smectite also fills pores and coats grains,
preventing the growth of secondary porosity. These factors reduce porosity,
thus groundwater is not available in the Unnamed Member Two which, like
Unnamed Member Three, is characterized by fine-grained deposits. Thus the
Unnamed Member Two is also a poor reservoir and a good seal. Also, porosity
is not available in the fine-grained deposits according to core logging.
The Unnamed Member Two is a potential source for hydrocarbons as indicated
by carbonaceous shale which occurs in this formation.
325

9-1-1-6-3 Lawrence Sandstone Member


In the Lawrence Sandstone Member, thin section porosity occurs between 0
and 15% at average of 1.4% whereas permeability is only 0.01 md.
Compaction, carbonate cement, quartz overgrowths and authigenic clays are
significant factors in the destruction of the primary porosity in the Lawrence
Sandstone. Carbonate cement is observed as a pore-filling cement in thin
section. Also, authigenic clays, particularly mixed layer illite-smectite, fill pores
in the Lawrence Sandstone. Thus, porosity and permeability are destroyed by
compaction and cementation in the Lawrence Sandstone. This indicates that
the Lawrence Sandstone has a poor reservoir potential but may contain traces
of groundwater and gas. Porosity and permeability are higher in the Lawrence
Sandstone than in the Kembla Sandstone, thus the porosity volume is greater in
the Lawrence Sandstone than in the Kembla Sandstone.
Sandstone is common throughout the Lawrence Sandstone which consists of
channel deposits. The environment of deposition in the Lawrence Sandstone is
similar to the environments of typical fluvial groundwater and hydrocarbon
reservoirs. Medium- to- coarse-grained sandstone is common in the Lawrence
Sandstone along with the presence of minor siltstone and shale. Medium- tocoarse-grained sandstone is more porous and permeable than in the siltstone
and shale. This indicates that potential volumes of groundwater and gas are
greater in the medium- to- coarse-grained sandstone than in the siltstone and
shale. In well logs, porosity was determined in coarse-grained sandstone
between 370.2 m and 372.2 m (Well 125, Fig. 9-4a, Porosity 1). Also, porosity
was determined in medium-grained sandstone between 639 m and 643.1 m
(Well 42, Fig. 9-4b) according to well logs. In general, the Laurence Sandstone
is recorded as having reservoir potential.
For example, thin sections showed that a sample of medium-grained sandstone
(37) in well 125 (367.15-367.18 m, Fig. 9-4a, Porosity 2) contains porosity of
15%, thus it probably contains groundwater or gas. Well logs confirmed the
presence of porosity in this sample. Sample (19) in well 127 (354.86-354.89 m)
is siltstone and did not contain porosity according to well logs and thin section.
326

a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

b)

Fig. 9-4: a) and b) Porosity in the Lawrence Sandstone Member

Carbonaceous shale is not apparent in the Laurence Sandstone Member, thus,


hydrocarbons would not be generated in this unit.
The Cape Horn Coal Member underlies the Lawrence Sandstone in the
Illawarra Coal Measures. This coal also has a well developed cleat that would
allow migration of both water and gas to the Lawrence Sandstone Member.
9-1-1-7 Loddon Sandstone
The Loddon Sandstone has thin section porosity ranging from 0 to 9.5% at
average of 2.6% with permeability only reaching 0.01 md. Reservoir quality is
affected by compaction and cementation in the Loddon Sandstone. Carbonate
cement is present, reducing porosity and permeability. SEM showed mixed
layer illite-smectite, chlorite and illite as pore-filling cements that reduce porosity
and permeability. In the Loddon Sandstone, the occurrence of porosity and
permeability indicates that groundwater is present. This indicates that the
Loddon Sandstone is a moderate groundwater and gas reservoir. Thin section
porosity is higher in the Loddon Sandstone than in the Lawrence Sandstone
and Kembla Sandstone, thus the potential volume of groundwater and gas is
greater in the Loddon Sandstone than in the Lawrence Sandstone and Kembla
Sandstone.

327

The Loddon Sandstone was deposited as the channel deposits. It contains


common medium- to coarse-grained sandstone. Also, siltstone and shale are
present in the Loddon Sandstone. The environment of deposition in the Loddon
Sandstone is not different from the environments for other typical fluvial
groundwater and hydrocarbon reservoirs. Medium- to coarse-grained sandstone
contains the porosity in the Loddon Sandstone whereas siltstone and shale are
good seals in the Loddon Sandstone. For example and according to well logs,
porosity is present in medium-grained sandstone between 630.4 m and 631.7 m
(Well 42, Fig. 9-5a) and also between 829.2 m and 830.8 m (Well 30, Fig. 9-5b).
Coarse-grained sandstone contains the porosity at depths between 347.5 m
and 349.2 m (Well 127, Fig. 9-5c, porosity 1). In Fig. 9-5d, well logs showed the
occurrence of porosity in whole the Loddon Sandstone in well 156. In general,
the Loddon Sandstone is recorded as a groundwater and gas reservoir.

a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

b)

c)

d)

Fig. 9-5: a), b), c) and d) Porosity in the Loddon Sandstone


328

This indicates that well logs confirmed the presence of porosity in samples
which contain thin section porosity. For example, sample (18) in well 127
(350.25-350.27 m, Fig. 9-6c, porosity 2) is coarse sandstone and includes 9.5%
of thin section porosity.
Also, carbonaceous shale was not recorded in the Loddon Sandstone which is
rich in sandstone. This means that the Loddon Sandstone is not a potential
source for hydrocarbons.
The Loddon Sandstone is underlain by the Balgownie Coal which includes a
large quantity of dissolved gas. The both water and gas may migrate from the
Balgownie Coal to the Loddon Sandstone.
9-1-1-8 Bulli Coal
The Bulli Coal seam contains a large quantity of dissolved gas which is why gas
drainage is needed before the coal seam is mined. The Bulli Coal is the main
source of gas in the Narrabeen Group. Gas probably migrated from the Bulli
Coal into the Coalcliff Sandstone and some of the overlying units.
9-1-1-9 Summary
In summary, thin section porosity and permeability showed that the Kembla
Sandstone, Lawrence Sandstone Member and Loddon Sandstone are suitable
hosts for groundwater and gas. The Loddon Sandstone contains a larger
volume of porosity than the Kembla Sandstone and Lawrence Sandstone
because thin section porosity is higher in the Loddon Sandstone. Also, the
Wilton Formation, Bargo Claystone, Darkes Forest Sandstone, Allans Creek
Formation, Unnamed Member Three and Unnamed Member Two would provide
good hydrocarbon and water seals within the coal measure succession.

9-1-1-10 Hydrocarbons in the Illawarra Coal Measures


Thermal history models were used by Faiz (1993) to study gas generation in the
Illawarra Coal Measures. Kinetic parameters were used to study hydrocarbon
generation from coal in the Illawarra Coal Measures (Faiz 1993). Espitalie et al.
329

(1988) had previously used kinetic parameters in a study of the Brent Coal in
the Northern Hemisphere. Oil, condensate and gas are recorded as the main
types of hydrocarbons that result from maturation of coal. Some authors such
as Espitalie et al (1988) and Burnham and Braun (1990) showed that long chain
hydrocarbons (oil and condensate) and gases were generated from coal as the
earliest hydrocarbons. According to Faiz (1993), in the Illawarra Coal Measures,
coal seams had macro-pores that allowed the migration of oil and condensate
during the early stages of coalification. Thus coal produced early biogenic gas
and subsequently-formed CO2. The presence of common macro-pores
contributed to the expulsion of gases. Gases are of thermogenic origin in the
Illawarra Coal Measures.

According to Faiz (1993), biochemical coalification was common in the Sydney


Subgroup when temperatures were less than 50C during the Late Permian and
Early Triassic and the coal was affected by bacterial activity. This period
includes production of biogenic gas (CH4 and CO2). H2O and CO2 were
produced during the very early stages of biochemical coalification whereas CH4
was produced slightly later during the early stages of biochemical coalification.
H2O and CO2 are produced by the action of aerobic bacteria whereas CH4 is
produced by the action of anaerobic bacteria. Coals in the Sydney Subgroup
produced biogenic gas during the Late Permian and Early Triassic. Several
factors suggest the expulsion of much of the gas from the system, including the
shallow burial depths, high permeabilities and the occurrence of connate water
in large amounts.
According to Faiz (1993), the main phase of thermal maturation was recorded
when temperatures exceeded 50C during subsequent burial, and this
maturation included physico-chemical variations. Also, H2O, CO2 and CH4 are
produced by dehydration and decarboxylation reactions which were present
during coalification during Early Triassic to Middle Jurassic between 50C and
100C. H2O and CO2 are common whereas CH4 is subordinate. Also, N2 and
H2S were produced during coal maturation but are uncommon. CH4 was also
produced in the Middle Jurassic to the Late Cretaceous which is known as the
main stage of hydrocarbon generation.
330

According to Faiz (1993), coal seams produced mainly CO2 during the Early
Triassic to Middle Jurassic with plentiful water. Also, factors that contributed to
the exclusion of CO2 included continuous burial and compaction of the
succession. Bray and Foster (1980) indicated the production of CO2 is mainly
associated with diagenesis of clay and feldspar minerals. The Middle Jurassic
was characterized by thermogenic hydrocarbon generation when the coal
measures were at depths exceeding 2000 m with temperatures in excess of
100C (Faiz, 1993). Also, the coal rank tended to increase and hydrocarbon
generation took place at the highest rate associated with high temperatures
during the Early Cretaceous and Late Cretaceous (Faiz, 1993).
According to Eadington et al. (1991), who studied the fluid history and clay
diagenesis in the Narrabeen Group, the Jurassic and Late Cretaceous periods
were characterised by active migration of hydrocarbons. Faiz (1993) showed
that the Illawarra Coal Measures generated important hydrocarbons in the
Middle Jurassic.
Also, Faiz (1993) studied hydrocarbon generation from coals in the Sydney
Subgroup in the northern, eastern and western parts of the basin in four drill
holes. He indicated that the coals generated hydrocarbons at rates of about 40
mg/g of coal in the northern area, 60 to 52 mg/g of coal in the eastern areas and
15 mg/g of coal in the western part of the basin.

9-1-2 Narrabeen Group


9-1-2-1 Coalcliff Sandstone
In the Coalcliff Sandstone, thin section porosity varies between 0 and 16% with
average of 4.5% whereas permeability is limited at 0.01 md. In the Coalcliff
Sandstone, low reservoir quality occurs as result of compaction and
cementation. Carbonate cement and authigenic clays effectively block many
pores. Carbonate cement fills large and small pores. Also, mixed layer illitesmectite and chlorite are pore-filling components in the Coalcliff Sandstone.

331

They reduce porosity and permeability, thus they reduce reservoir potential in
the Coalcliff Sandstone.
These results indicate that the Coalcliff Sandstone is a poor reservoir but it has
some porosity. The presence of porosity and permeability indicates that the
Coalcliff Sandstone is suitable for gas accumulations. Mullard (1995) showed
that hydrocarbons were generated from the Coalcliff Sandstone, thus gas was
generated in the Narrabeen Group. Hamilton and Galloway (1989) also
indicated that the Coalcliff Sandstone is poor reservoir.
Environment of deposition in the Coalcliff Sandstone is not very different from
the environments of other typical fluvial groundwater and hydrocarbon
reservoirs. The Coalcliff Sandstone is represented by medium- and coarsegrained sandstone which is moderately to well sorted. These units show low to
moderate porosity and permeability. Siltstone and shale are also present in the
Coalcliff Sandstone. They do not include porosity and permeability and are
lithological seals. Well logs showed the presence of porosity in conglomerate
between 340.1 m and 342 m (Well 125, Fig. 9-6a) and also in medium to
coarse-grained sandstone between 330.9 m and 337.6 m (Well 127, Fig. 9-6b).
Medium-grained sandstone also contains a porous zone between 597 m and
598.6 m (Well 42, Fig. 9-6c) according well logs. In general, medium- and
coarse-grained sandstone are more common than in the Illawarra Coal
Measures, thus the Coalcliff Sandstone has a better reservoir potential for
groundwater and gas.
For example, thin section showed that the sample of medium-grained
sandstone (25) in well 30 (813.32-813.35 m, Fig. 9-6d) contains a porosity of
10.5%. Well logs confirmed the presence of porosity in this sample. Sample
(10) in well 124 (346.16-346.19 m) is fine-grained sandstone and did not
contain much porosity according to well logs and thin section.
The abundance of sandstone, combined with the absence of carbonaceous
shale, shows that the Coalcliff Sandstone is not a potential source for
hydrocarbons.
332

a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

b)

c)

d)

Fig. 9-6: a), b), c) and d) Porosity in the Coalcliff Sandstone

9-1-2-2 Wombarra Claystone


In the Wombarra Claystone, thin section porosity reaches up to 13.5% with
average of 3.5%, whereas permeability varies between 0.07 and 0.1 md with
average of 0.1 md. Thin section porosity was recorded in sandstone beds in the
Wombarra Claystone. Carbonate cement and authigenic clays prevent the
development of much porosity in the Wombarra Claystone. Carbonate cement
is observed as pore-fillings in thin section whereas authigenic clays fill pore
spaces. These factors cause the loss of porosity and permeability in sandstone
beds, thus available spaces for porosity are rare in the sandstone beds. These
333

results indicate that sandstone beds in the Wombarra Claystone would act as
poor reservoirs but may contain groundwater and gas.
In the Wombarra Claystone, most samples consist of shale but a few sandstone
beds are characterized by porosity. Shale beds are less porous and permeable,
thus they are good seals. Well logs demonstrated that sandstone beds contain
porosity. For example, coarse-grained sandstone has porosity between 336.9 m
and 337.9 m (Well 125, Fig. 9-7a). Also, medium-grained sandstone contains
porosity between 327.9 m and 331.6 m (Well 124, Fig. 9-7b) according well
logs. In well 156, well logs showed that gas occurs between 417.2 m and 421.3
m in sandstone beds in the Wombarra Claystone (Fig. 9-7c).

a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

b)

c)

Fig. 9-7: a), b) and c) Porosity in the Wombarra Claystone


For example, sample sandstone (4) in well 126 (354.64-354.67 m, Fig. 9-8)
contains porosity of 13.5%, thus it probably contains porosity according to both
334

thin section and well logs. In general, the Wombarra Claystone forms a good
seal due to the widespread occurrence of shale beds.

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 9-8: Porosity in the Wombarra Claystone


The Wombarra Claystone is not a good potential source for hydrocarbons
because the shale is not carbonaceous.

9-1-2-3 Scarborough Sandstone


In the Scarborough Sandstone, thin section porosity is higher than in the
underlying units, ranging from 0 to 19.5% with average 6.5%. Permeability
varies between 0.04 and 0.05 md with average 0.05 md. Reservoir quality is
reduced by compaction, siderite cement, quartz overgrowths and authigenic
clays. Siderite cement fills large and small pores, thus it has reduced porosity.
Mixed layer illite-smectite and illite are seen as pore-filling components by
scanning electron microscope. Diagenetic alterations have reduced porosity
and permeability, thus they are effectively trapping the groundwater and gas.
These results indicate that the Scarborough Sandstone is a poor to moderate
reservoir and probably contains groundwater and gas. Thin section porosity is
higher in the Scarborough Sandstone than in the Coalcliff Sandstone. Thus, the
volume of porosity in the Scarborough Sandstone is greater than in the Coalcliff
Sandstone.
Medium- and coarse-grained sandstone is the dominant components in the
Scarborough Sandstone and it is more porous and permeable than the siltstone
and shale interbeds which are also observed. The volume of porosity is also
greater in medium- and coarse-grained sandstone than in siltstone and shale
335

according to well logs. Between 279 m and 280.8 m (Well 127, Fig. 9-9a),
porosity is determined in coarse-grained sandstone. Also, porosity occurs in
medium-grained sandstone between 296.6 m and 298 m (Well 124, Fig. 9-9b)
and between 327.3 m and 333.1 m (Well 126, Fig. 9-9c). The environment of
deposition for the Scarborough Sandstone is not very different from the
environments of typical fluvial groundwater and hydrocarbon reservoirs.

a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

b)

c)

Fig. 9-9: a), b) and c) Porosity in the Scarborough Sandstone

In some samples, well logs and thin section porosity confirmed that porosity is
found. For example, sample (38) in well 42 (518.33-518.36 m, Fig. 9-10) is a
coarse-grained sandstone and includes 19.5% thin section porosity. This
sample contains porosity according well logs and thin section.

336

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

Fig. 9-10: Porosity in the Scarborough Sandstone


The Scarborough Sandstone would not act as a potential source for
hydrocarbons because of the abundance of sandstone and the absence of
carbonaceous shale.

9-1-2-4 Stanwell Park Claystone


Most samples consist of shale in the Stanwell Park Claystone. Two samples of
sandstone were described in thin section. These two samples showed that
porosity ranges from 3-10.5% at average of 6.8% whereas permeability is 0.03
md. Sandstone beds in the Stanwell Park Claystone include carbonate cement
and authigenic clays which reduce porosity and permeability. Sandstone beds
in the Stanwell Park Claystone may contain groundwater or gas indicated by the
presence of porosity and permeability. Well logs confirmed that two samples
sandstone, described by thin section contain porosity.
In general, the Stanwell Park Claystone includes common shale beds which are
a sealing lithology. Thus, the Stanwell Park Claystone is a poor reservoir but is
a good seal. In this study, the Stanwell Park Claystone is not a potential source
for hydrocarbons because the absence of carbonaceous shale.
9-1-2-5 Bulgo Sandstone
Thin section porosity is the highest in the Bulgo Sandstone, varying between
4.2% and 17.3% with average of 12%. Permeability reaches to 79.4 md with
average of 32.5 md. Siderite cement, quartz overgrowth and authigenic clays
reduce reservoir quality in the Bulgo Sandstone. They fill pore spaces in most
337

samples to reduce porosity. The Bulgo Sandstone contains higher thin section
porosity and permeability than the Scarborough Sandstone and Coalcliff
Sandstone. This indicates that the porosity volume for groundwater or gas in the
Bulgo Sandstone is greater compared with the Scarborough Sandstone and
Coalcliff Sandstone. Thin section porosity was confirmed by well logs which
showed the presence of porosity in the Bulgo Sandstone. For example, sample

a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

b)

c)

d)

Fig. 9-11: a), b), c) and d) Porosity in the Bulgo Sandstone


(25) in well 42 (378.17-378.20 m, Fig. 9-11a) is coarse-grained sandstone and
includes 16.5% of thin section porosity. Well logs confirmed the presence of
porosity in this sample.
338

Environment of deposition in the Bulgo Sandstone is not very different from the
environments of other typical fluvial groundwater and hydrocarbon reservoirs.
The Bulgo Sandstone is characterized by medium- and coarse-grained
sandstone with the presence of small amounts of siltstone and shale. Mediumand coarse-grained sandstone contain more porosity and permeability than
siltstone and shale that represent sealing lithologies. Between 339.7 and 346 m
(Well 42, Fig. 9-11b) and between 395.3 and 397 m (Well 42, Fig. 9-11c), well
logs showed the presence of porosity in medium-grained sandstone and
coarse-grained sandstone, respectively. In well 156, porosity occurs in coarsegrained sandstone between 332.7 and 334.7 m (Fig. 9-11d).
In general, the Bulgo Sandstone is a moderate to good reservoir and contains
groundwater and possibly gas. The abundance of sandstone with the lack of
carbonaceous shale means that the Bulgo Sandstone is not a potential source
for hydrocarbons.
9-1-2-6 Bald Hill Claystone
Shale beds are widespread in the Bald Hill Claystone. One sandstone sample
was analysed in thin section. It showed that thin section porosity does not
exceed 1.5%. Permeability is 0.03 md in the Bald Hill Claystone. Siderite
cement played a role in the destruction of thin section porosity in one sandstone
sample in the Bald Hill Claystone. Also, authigenic clays reduced porosity in the
same sample.
In general, the Bald Hill Claystone is poor a reservoir but is a good seal
because the widespread of shale beds. Carbonaceous shale occurs in the Bald
Hill Claystone which is the potential source for hydrocarbons.
9-1-2-7 Newport Formation
Fine-grained deposits such fine-grained sandstone, siltstone and shale occurs
commonly in the Newport Formation which contains thin section porosity at
average of 0.8% with permeability at average of 0.2 md. Siderite cement and
authigenic kaolinite destroy the porosity and permeability in the Newport
339

Formation. Also, the thin section results and well logs demonstrate that the
Newport Formation is similar to the Bald Hill Claystone. Both represent poor
reservoirs but act as good seals and are potential sources for hydrocarbons.
In summary, the Bulgo Sandstone, Scarborough Sandstone and Coalcliff
Sandstone are potential groundwater and gas reservoirs. The Newport
Formation, Bald Hill Claystone, Stanwell Park Claystone and Wombarra
Claystone are good seals in the Narrabeen Group.

9-1-3 Hawkesbury Sandstone


The Hawkesbury Sandstone is characterized by medium- to coarse-grained
sandstone with a lesser presence of fine sandstone, siltstone and shale. This
unit has thin section porosity ranging from 0 to 19.3% with average of 8.2%.
Also, permeability is recorded between 0.2 and 22.3 md with average 7.9 md.
Reservoir quality is lost by quartz overgrowths which are present as pore-filling
cement. This indicates that quartz overgrowths reduce porosity and
permeability, thus they reduce the amount of available groundwater in the
Hawkesbury Sandstone.
The environment of deposition in the Hawkesbury Sandstone mainly is not
different from the environments of typical fluvial aquifer and hydrocarbon traps.
Medium- and coarse-grained sandstone are common in the Hawkesbury
Sandstone whereas siltstone and shale are rare. Well logs showed that
medium- and coarse-grained sandstone contains porosity whereas siltstone and
shale are poor. This indicates the importance of facies and environment in the
determination of available space for groundwater. In well logs, porosity is
present in medium-grained sandstone between 53 and 57.1 m (Well 115, Fig. 912a). Also, it occurs in coarse-grained sandstone between 64 and 67 m (Well
156, Fig. 9-12b) and between 227.4 and 239.2 m (Well 42, Fig. 9-12c)
according the well logs.
In described samples, thin sections showed the presence of porosity in
sandstone beds. Well logs indicated that these samples contain porosity. For
example, sample (4) in well 18a (243.19-243.22 m, Fig. 9-12d) is coarse340

grained sandstone and includes 17% thin section porosity. This sample
contains porosity according to well logs and is rich in porosity according thin
section.
a)

Gamma
GAM

Density-Caliper
DENP DENL CADE

Porosity
DEPO RPOR

Sonic
VL2F VL6F

Resistivity
FE1 FE2

b)

c)

d)

Fig. 9-12: a), b), c) and d) Porosity in the Hawkesbury Sandstone


The results indicated that the Hawkesbury Sandstone is moderate to poor
reservoir and may contain groundwater. It is unlikely that the Hawkesbury
Sandstone is the potential source for oil or gas since the included shales would
only provide local seal potential.

341

In summary, the Hawkesbury Sandstone is higher in thin section porosity than


the Narrabeen Group and Illawarra Coal Measures. Thus, the Hawkesbury
Sandstone contains more available space for groundwater than the Narrabeen
Group and Illawarra Coal Measures.

9-2 Discussion
Several factors can control hydrocarbon reservoirs in the Illawarra Coal
Measures, Narrabeen Group and Hawkesbury Sandstone. These factors are
lithology, diagenetic alteration, depositional environment and structures. These
factors control porosity and permeability, thus they can control hydrocarbon
reservoirs.
In the Illawarra Coal Measures, thin section porosity is completely absent in the
Wilton Formation, Bargo Claystone, Darkes Forest Sandstone, Allans Creek
Formation, Unnamed Member Three and Unnamed Member Two which are
characterized by shale, siltstone and fine-grained sandstone. Thus, these
formations are less porous and permeable, they are poorly sorted and finegrained. Carbonate cement and authigenic clays are present in the shale,
siltstone and fine-grained sandstone as potential barriers to prevent the
preservation of porosity and permeability. Ankerite cement and mixed layer
illite/smectite have considerable influence on reservoir quality, thus pore spaces
available for hydrocarbons occurrences are absent. These factors represented
by lithology, diagenetic alteration, sorting and grain size indicate that these
formations are lithological seals and confining layers in the Illawarra Coal
Measures.
Also, changes in facies and grain size play a role in influencing gas migration
and groundwater movement in the Illawarra Coal Measures. Freeze and Cherry
(1979) showed that the hydraulic gradient is changed in aquifers as result of
variations in facies. The Kembla Sandstone, Lawrence Sandstone Member and
Loddon Sandstone were deposited in fluvial environments that were not
dissimilar to the environments of many other typical fluvial groundwater and
hydrocarbon reservoirs. Medium- to coarse-grained sandstone, siltstone and
342

shale units are present in the Kembla Sandstone, Lawrence Sandstone and
Loddon Sandstone. Medium- to coarse-grained sandstone units are more
common in the channel facies than siltstone or shale. They are moderately to
well sorted, with rounded to sub-rounded grains. These features can preserve
moderate porosity in medium- to coarse-grained sandstone and indicate that
these units probably contain gas or water (Wolela, 2009). Well logs confirmed
the presence of porosity in the medium- to coarse-grained sandstone.
Diagenetic alteration has influenced the reservoir quality, and hence has
affected the possibility of hydrocarbon occurrences. Carbonate cement is also
commonly present in the medium- to coarse-grained sandstone where it fills
pore spaces, reducing thin section porosity. Clay minerals are present in two
forms as detrital matrix and authigenic clays. The influence of the latter on
porosity is stronger than the former (Sams and Martijn, 2001). Mechanical
compaction is also common in the medium- to coarse-grained sandstone and
further reduces thin section porosity. All these factors affect reservoir quality,
and thus the pore space available for gas or water is probably low in the
medium- and coarse-grained sandstone. Dickey (1981) showed that the
migration of hydrocarbons is reduced by the presence of clay minerals. This
indicates that mechanical compaction, carbonate cement and authigenic clays
all lead to the production of effective seals. Dissolution of grains results some
secondary porosity in the medium- and coarse-grained sandstone, thus
increasing the pore spaces available for gas or water. According to Ryan
(2005), an abundance of volcanic rock fragments in a sandstone commonly
leads to low porosity, destruction of hydrocarbon reservoirs and the production
of permeability traps. In this study, volcanic rock fragments are common in
sandstone beds in the Kembla Sandstone, Lawrence Sandstone and Loddon
Sandstone, thus they are contain hydrocarbon traps adjacent to low porosity
zones.
Thin section porosity is absent in the floodplain shale and siltstone units that
form lithological seals and confining layers in the Kembla Sandstone, Lawrence
Sandstone and Loddon Sandstone. These shale and siltstone units are poorly
sorted and fine-grained. Schoen (1996) showed that the preservation of primary
porosity can be affected by grain size and sorting. This indicates that poor
343

sorting and fine grains prevent the preservation of visible porosity in shale and
siltstone units. Also, secondary mixed layer illite/smectite, kaolinite and illite are
distributed through the shale and siltstone units and further reduce the porosity
and permeability. Thus, pore spaces available for gas or water probably are
absent in these units. This indicates that shale and siltstone beds form local
vertical permeability seals in and between the Kembla Sandstone, Lawrence
Sandstone and Loddon Sandstone.
The Tongarra Coal, Allans Creek Coal Member, Wongawilli Coal, Hargrave
Coal Member, Cape Horn Coal Member, Balgownie Coal Member and Bulli
Coal all contain coal seam gases and are recognized as the main source of gas
in the Illawarra Coal Measures. They result gas to the Kembla Sandstone,
Lawrence Sandstone and Loddon Sandstone which are underlain by Allans
Creek Coal Member, Cape Horn Coal Member and Balgownie Coal Member,
respectively. Thus gas may migrate from the Allans Creek Coal Member to the
overlying Kembla Sandstone and from the Cape Horn Coal Member to the
overlying Lawrence Sandstone. Also, gas may migrate from the Balgownie Coal
Member to the overlying Loddon Sandstone.
Gentle synclines and faults are present in the Illawarra Coal Measures and
have influenced the migration of gas or movement of water. Synclines may be
recorded as the potential barriers to gas flow in the Illawarra Coal Measures.
Also, faults are important for draining or preserving hydrocarbons in a reservoir.
In some individual units, reservoirs may be restricted by faults, which are
present as seals and break anticlinal reservoirs according to Dickey (1981).
Also, faults have probably influenced groundwater movement in the Illawarra
Coal Measures. Faults can prevent free groundwater movement and form semiisolated groundwater potential within the Illawarra Coal Measures. This
indicates that the path of the groundwater movement can be changed as result
of the occurrence of faults.
Mullard (1995) showed that the Stanwell Park Claystone, Bald Hill Claystone
and Wombarra Claystone would be good seals in the Narrabeen Group. In this
study, reservoir quality such as porosity and permeability disappear in the
344

Stanwell Park Claystone, Bald Hill Claystone and Newport Formation. Shale
and siltstone beds are dominant in these formations and they are poorly sorted
and fine-grained. Thus, they have little porosity and permeability. Siderite,
ankerite, kaolinite and mixed layer illite/smectite are present and reduce
porosity and permeability in these formations. The factors including lithology,
sorting, grain size and diagenetic alteration reduce porosity and permeability,
and thus the pore space available for hydrocarbon occurrences is low. This
indicates that the Stanwell Park Claystone, Bald Hill Claystone and Newport
Formation are very poor reservoirs but are good seals in the Narrabeen Group.
The environment of deposition in the Coalcliff Sandstone, Scarborough
Sandstone and Bulgo Sandstone was fluvial and is not very different from the
environments of many other typical fluvial groundwater and hydrocarbon
reservoirs. The Coalcliff Sandstone, Scarborough Sandstone and Bulgo
Sandstone contain medium- and coarse-grained sandstone, shale and siltstone.
Medium- and coarse-grained sandstone are more common than shale and
siltstone in these formations and are moderately to well sorted with rounded to
sub-rounded grains. They have low to moderate thin section porosity which is
higher in the Bulgo Sandstone than in the Coalcliff Sandstone and Scarborough
Sandstone. Thus, these formations are poor to moderate reservoirs and
probably contain gas or water (Wolela, 2009). Also, Mullard (1995) in his study,
indicated that these formations could be hydrocarbon reservoirs in the
Narrabeen Group. There are potential barriers to reduce the occurrence of gas
or water in the medium- and coarse-grained sandstone. These are represented
by mechanical compaction, carbonate cement, authigenic clays and quartz
overgrowths. Mechanical compaction occurs commonly in these medium- to
coarse-grained sandstones and reduces porosity and permeability. Carbonate
cement, authigenic clays and quartz overgrowths also fill pore spaces and
reduce porosity and permeability in the medium- and coarse-grained sandstone.
Thus the pore spaces available for gas or water are probably low in these units
as result of diagenetic alteration. Minor secondary porosity occurs as a result of
dissolution of grains and may provide some pore spaces for gas or water in the
medium- and coarse-grained sandstone. In Adigrat Sandstone, Wolela (2009)
showed that medium- to coarse-grained sandstone can be porous and
345

permeable and may be characterized by the presence of oil and gas. Well logs
showed the presence of porosity in the medium- to coarse-grained sandstone in
this study. In this study, the Coalcliff Sandstone, Scarborough Sandstone and
Bulgo Sandstone are rich in volcanic rock fragments which may form
hydrocarbon traps according to Ryan (2005).
The associated shale and siltstone units have very low thin section porosity and
consist of very fine poorly sorted grains in the Coalcliff Sandstone, Scarborough
Sandstone and Bulgo Sandstone. Schoen (1996) showed that the preservation
of primary porosity can be affected by grain size and sorting. Thus, these fine
units represent poor reservoirs and are recognized as confining layers in these
formations. Secondary clay minerals, particularly mixed layer illite/smectite, are
common in the shale and siltstone units and prevent the preservation of
porosity, thus pore spaces are unavailable for gas or water in these units. This
indicates that the shale and siltstone beds form local vertical permeability seals
in the Coalcliff Sandstone, Scarborough Sandstone and Bulgo Sandstone.
Shale, siltstone and sandstone are also observed in the Wombarra Claystone
which is characterized by a dominance of shale units. Thin section porosity in
the shale and siltstone units is further reduced by carbonate cement and clays,
thus visible porosity is zero. The shale and siltstone units are also poorly sorted
and fine-grained, thus porosity is very low in these units. This indicates that the
shale and siltstone units form lithological seals and confining layers in the
Wombarra Claystone. However, the few included sandstone beds have low thin
section porosity and are more porous and permeable than the shale or siltstone
units. The sandstone beds are moderately to well sorted with rounded to subrounded grains. This indicates that the sandstone beds could support gas or
water resources in the Narrabeen Group (Wolela, 2009). The influence of
diagenetic alteration on reservoir quality can affect hydrocarbon occurrences in
sandstone beds. Pore-filling carbonate cement, particularly siderite and
ankerite, is present in the sandstone beds and causes additional porosity loss.
Authigenic clays are also present as pore-filling cements and they also cause a
loss of porosity. These factors reduce the pore spaces available for gas or
water in sandstone beds.
346

The Bulli Coal underlies the Coalcliff Sandstone and is a source of gas in the
Narrabeen Group. The gas migrates from the Bulli Coal into the Coalcliff
Sandstone. The latter has more gas than in the Scarborough Sandstone or
Bulgo Sandstone. Ryan (2005) confirmed that the Illawarra Coal Measures may
be the source for gas in the Narrabeen Group. He showed evidence that gas in
the Coalcliff Sandstone and in the coal measures has a similar composition. In
this study, the base of the Coalcliff Sandstone contains porous and permeable
beds. Thus, gas probably migrated through these beds to the upper parts of the
Coalcliff Sandstone.
The migration of gas or movement of water may be affected by folds and faults
in the Narrabeen Group. Synclines have been recorded as potential barriers to
gas flows in the Narrabeen Group. Faults probably may be present as seals and
break anticlinal reservoirs, thus they reduce the reservoirs in Narrabeen Group
(Dickey, 1981). This may suggests that faults are important in the destruction of
hydrocarbon reservoirs. In the Narrabeen Group groundwater movement is
probably controlled by faults which can prevent free groundwater movement
and form semi-isolated groundwater reservoirs. This indicates that the presence
of faults probably changed the path of groundwater migration in the Narrabeen
Group.
Thin section porosity is high in the Hawkesbury Sandstone which is near the
land surface and is recognized as having good groundwater potential. In this
study, the Hawkesbury Sandstone is not a homogeneous aquifer as result of
local variations in the porosity percentages (Grey and Ross, 2003; Freed, 2005;
Gentz, 2006; Chambers, 2006). Rainfall is the main source of groundwater in
the Hawkesbury Sandstone, which is recharged by direct infiltration. This shows
the presence of a positive relationship between porosity and aquifer potential. It
also indicates that primary and secondary porosity allow groundwater flow
(Pritchard et al., 2004) and forms the paths for water movement in the
Hawkesbury Sandstone. Freed (2005) showed that the Hawkesbury Sandstone
contains primary and secondary porosity and is a heterogeneous aquifer. In this
study, groundwater flow is affected by primary porosity more than by secondary
porosity. Also, changes in facies and grain size play a role in influencing
347

groundwater movement. Freeze and Cherry (1979) showed that the hydraulic
gradient is changed in aquifers as a result of the variations in facies. In this
study, two types of units are determined based on reservoir quality in the
Hawkesbury Sandstone. They are coarse-grained deposits and fine-grained
deposits, with the former being more common than the later. Coarse-grained
deposits are represented by medium- and coarse-grained sandstone whereas
fine-grained deposits are represented by shale and siltstone units. The mediumand coarse-grained sandstone units are thick, massive and trough crossbedded. These units are mainly composed of quartz grains and are represented
by coarsening upwards and fining upwards trends. The sandstone mainly
consists of well sorted, rounded to sub-rounded grains. Thus, medium- and
coarse-grained sandstone units are characterized by containing visible thin
section porosity. In general, medium- and coarse-grained sandstone units are
probably characterized by moderate to good groundwater storage and flow
potential. Good aquifer can be indicated by well sorted and coarse grains which
preserve a greater quantity of primary porosity (Bowman, 1971; Schoen, 1996).
In the Hawkesbury Sandstone, medium- to coarse-grained sandstone is
characterized by these features. Also, water flow rates are probably increased
by the abundance of primary porosity in these well sorted and coarse-grained
sandstones according to the interpretation of Freed (2005). Diagenetic
alteration such as chemical compaction, quartz overgrowths, authigenic clays
and carbonate cement are recorded as potential barriers to flow and may
reduce water quality in the medium- to coarse-grained sandstone. Cement fills
many space pores and is present as a barrier to the development of porosity
and permeability. Also, packing and compaction reduce porosity and
permeability. This indicates that these factors, particularly pore-filling cement
and the tightly packed nature of the grains, affect the movement and storage of
water in medium- and coarse-grained sandstone (Branagan, 2000). Water
quality is affected by clays which may include dispersed clay and structural clay.
Dispersed clays fill some pore spaces and have a strong influence on
sandstone permeability. Structural clays exist as clay grains and are present
between grain contacts. The influence of clay grains on water quality is very low

348

(Sams and Martijn, 2001). Also, iron oxide can affect porosity and permeability
in the Hawkesbury Sandstone and is present as potential barriers to flow.
Thin section porosity is completely absent in shale and siltstone units in the
Hawkesbury Sandstone which are poorly sorted and fine-grained. Clays
represented by kaolinite, mixed layer illite/smectite and illite are observed in
most shale and siltstone beds according X-ray diffraction and scanning electron
microscope. Thus, shale and siltstone beds are less porous and permeable,
and they form groundwater-poor confining layers in the Hawkesbury Sandstone.
Grain size, mineralogy and well logs contribute to the determination of
impermeable shale beds in the Hawkesbury Sandstone.
The groundwater movement may be affected by synclines and faults in the
Hawkesbury Sandstone. Faults can prevent free groundwater movement and
can form semi-isolated groundwater potential. This indicates that the presence
of faults probably changes the path of groundwater movement in the
Hawkesbury Sandstone. Also, faults are probably present as potential barriers
in the Hawkesbury Sandstone. Synclines can be recognized as places for water
accumulation in the Hawkesbury Sandstone.

349

Table 9-1: Summary information from each individual unit.

350

Table 9-1 (cont.): Summary information from each individual unit

351

Table 9-1 (cont.): Summary information from each individual unit

352

Chapter Ten
Conclusions
The Illawarra Coal Measures is mostly litharenite with rare sublitharenite,
whereas the Narrabeen Group is litharenite and sublitharenite to rare
quartzarenite. The Hawkesbury Sandstone is richer in quartz than the
underlying Narrabeen Group and Illawarra Coal Measures and falls in
quartzarenite to sublitharenite fields. These variations indicate that the Illawarra
Coal Measures have a similar provenance to the Narrabeen Group and it is
different from the Hawkesbury Sandstone. The Illawarra Coal Measures and
Narrabeen Group plot in the lithic recycled to transitional recycled and
quartzose recycled provenance fields whereas the Hawkesbury Sandstone
shows a craton interior to quartzose recycled provenance. A few samples from
the Narrabeen Group plot in the craton interior field. In the Illawarra Coal
Measures and Narrabeen Group, quartz and lithic grains are derived from the
New England Fold Belt and an eastern volcanic arc with minor contributions
from the Lachlan Fold Belt. In the Hawkesbury Sandstone, the cratonic Lachlan
Orogen is the main source of the quartz grains.
Petrographic data demonstrated that lithic grains are more common in the
Illawarra Coal Measures and Narrabeen Group than in the Hawkesbury
Sandstone, which is characterized by quartz grains. Feldspar grains are sparse
in all three units but are least common in the Hawkesbury Sandstone.
Quartz includes monocrystalline and polycrystalline quartz grains whereas
feldspar grains consist of K-feldspar and plagioclase. Rock fragments are
volcanic or sedimentary, with an absence of volcanic rock fragments in the
Hawkesbury Sandstone. Mica includes more common muscovite than biotite.
The heavy minerals comprise hematite, hornblende, rutile, zircon and
tourmaline in trace percentages.
Thin section and scanning electron microscope analyses were used to describe
the alteration and diagenesis in the southern Sydney Basin. Results showed
353

that carbonate cement is most common in the Illawarra Coal Measures and the
Narrabeen Group whereas quartz overgrowths are the dominant cement in the
Hawkesbury Sandstone.
In the Illawarra Coal Measures, authigenic minerals included quartz
overgrowths, authigenic clay minerals, carbonate cement, authigenic feldspar,
authigenic pyrite and iron oxide cement. Quartz overgrowths are uncommon
whereas authigenic clay minerals are characterized by kaolinite, illite, mixedlayer illite-smectite and chlorite. Siderite is the most common carbonate cement
in the Illawarra Coal Measures.
In the Narrabeen Group, authigenic minerals included quartz overgrowths,
authigenic clay minerals, carbonate cement, authigenic feldspar, authigenic
pyrite and iron oxide cement. Authigenic clay minerals are common, and are
represented by kaolinite, illite, mixed-layer illite-smectite and chlorite. Quartz
overgrowths are rare. Siderite, ankerite, calcite and dolomite are present as
carbonate cements but siderite is dominant.
Thin section porosity is much higher in the Hawkesbury Sandstone than in the
Illawarra Coal Measures and Narrabeen Group.
In the Illawarra Coal Measures, thin section porosity is observed in some
samples whereas density porosity is represented in most samples. In the
Narrabeen Group and Hawkesbury Sandstone, thin section porosity and density
porosity are present in most samples.
Helium porosity was measured in twenty one samples from the Illawarra Coal
Measures, Narrabeen Group and Hawkesbury Sandstone. In comparison with
thin section porosity the helium porosity is shows greater values.
The study noted the presence of two types of porosity in the Illawarra Coal
Measures, Narrabeen Group and Hawkesbury Sandstone. They are primary
and secondary porosity. The latter occurs at higher percentages than former in
the Illawarra Coal Measures and Narrabeen Group whereas primary porosity is
more common than secondary porosity in the Hawkesbury Sandstone.

354

The influence of mechanical compaction on thin section porosity is stronger in


the Narrabeen Group and Illawarra Coal Measures than in the Hawkesbury
Sandstone. Chemical compaction has a greater affect on thin section porosity in
the Hawkesbury Sandstone than in the Illawarra Coal Measures and Narrabeen
Group.
The Hawkesbury Sandstone is richer in the quartz grains than the Illawarra Coal
Measures and Narrabeen Group, and hence quartz overgrowths are more
common in the Hawkesbury Sandstone. Thus, thin section porosity is more
affected by quartz overgrowths in the Hawkesbury Sandstone than the Illawarra
Coal Measures and Narrabeen Group.
Carbonate cement is prevalent in the Illawarra Coal Measures and Narrabeen
Group, thus it has an important influence on thin section porosity. In contrast,
carbonate cement is rare in the Hawkesbury Sandstone, thus its influence on
thin section porosity in this formation is low.
In the Illawarra Coal Measures and Narrabeen Group, and to a lesser extent in
the Hawkesbury Sandstone, pore-filling clays are dominant and reduce thin
section porosity in all these units. Also, grain-coating clays preserve thin section
porosity in these units but their importance is greatest in the Narrabeen Group.
In the Illawarra Coal Measures and Narrabeen Group, lithic and feldspar grains
are more common than in the Hawkesbury Sandstone, thus secondary porosity
caused by unstable grain dissolution is more abundant in the Illawarra Coal
Measures and Narrabeen Group than in the Hawkesbury Sandstone.
In the Illawarra Coal Measures, individual units represented by the Wilton
Formation, Bargo Claystone, Darkes Forest Sandstone, Allans Creek
Formation, Unnamed Member Three and Unnamed Member Two have very low
thin section porosity values. These formations are characterized by poorly
sorted shale, siltstone and fine-grained sandstone which have low visible
porosity and permeability. Thus, these formations are recorded as effective
water and hydrocarbon seals and confining layers in the Illawarra Coal
Measures. On the other hand, the Kembla Sandstone, Lawrence Sandstone
355

and Loddon Sandstone are represented by medium- and coarse-grained


sandstone units which have low thin section porosity and are moderately to well
sorted. Thus, they would contain gas or water in the subsurface. In the Illawarra
Coal Measures, coal seam gas is commonly presented in the Tongarra Coal,
Allans Creek Coal Member, Wongawilli Coal, Hargrave Coal Member, Cape
Horn Coal Member, Balgownie Coal Member and Bulli Coal. These formations
probably are the source of gas in the Kembla Sandstone, Lawrence Sandstone
and Loddon Sandstone which are more porous and permeable. For example,
gas may migrate from the Allans Creek Coal Member to the overlying Kembla
Sandstone and from the Cape Horn Coal Member to the overlying Lawrence
Sandstone. Also, gas may migrate from the Balgownie Coal Member to the
overlying Loddon Sandstone. Porosity is higher in the Loddon Sandstone than
in the Kembla Sandstone and Lawrence Sandstone. Thus, the volume available
for gas or water in the Loddon Sandstone is probably greater than in the
Kembla Sandstone and Lawrence Sandstone.
In the Narrabeen Group, the Wombarra Claystone contains common shale and
siltstone and minor thin sandstone beds. Thin section porosity is absent in shale
and siltstone units which form seals and confining layers. Thin section porosity
occurs in the more porous and permeable sandstone beds in the Wombarra
Claystone. These beds probably represent minor gas or water resources in the
Narrabeen Group. Visible porosity and permeability are completely absent in
the Stanwell Park Claystone, Bald Hill Claystone and Newport Formation.
These formations include mainly shale and siltstone units which are poorly
sorted and are characterized by abundant clays. Visible porosity and
permeability are absent in shale and siltstone units since all intergranular
spaces are filled with clay. Thus, the Stanwell Park Claystone, Bald Hill
Claystone and Newport Formation are very poor reservoirs but represent good
seals in the Narrabeen Group. On the other hand, the Coalcliff Sandstone,
Scarborough Sandstone and Bulgo Sandstone are characterized by mediumand coarse-grained sandstone with only minor amounts of shale and siltstone.
In these formations, the medium- and coarse-grained sandstone is moderately
to well sorted and shows low to moderate thin section porosity. Thus, they
represent poor to moderate reservoirs and probably contain gas or water. The
356

Bulgo Sandstone contains the highest percentage of thin section porosity in the
Narrabeen Group. Thus, the volume available for gas or water probably is the
greatest in the Bulgo Sandstone. The Coalcliff Sandstone directly overlies the
Bulli Coal, thus it may contains larger amounts of gas than in the Scarborough
Sandstone and Bulgo Sandstone. Thus, coal seam gas in the Illawarra Coal
Measures may be the source of gas in the Narrabeen Group.
Good porosity aquifers are recorded in the Hawkesbury Sandstone as result of
the presence of both primary and secondary porosity in these beds. The
Hawkesbury Sandstone has high thin section porosity in most units, thus it has
a good groundwater potential. The common thin section porosity occurs in well
sorted medium- and coarse-grained sandstone units which are rich in quartz.
These sandstone units may coarsen upwards or fine upwards and thus show
vertical variations in porosity. These features indicate that medium- and coarsegrained sandstone units are probably characterized by moderate to good
groundwater storage and flow potential. Fine-grained deposits represented by
impermeable shale and siltstone units are poorly sorted and are very poor in
thin section porosity. Thus, they are recognized as confining layers and provide
very poor groundwater storage in the Hawkesbury Sandstone. These shale and
siltstone beds form local vertical permeability seals in the Hawkesbury
Sandstone.

357

358

References
Aagaard, P., Jahren, J. S., Harstad, A. O., Nilsen, O., and Ramm, M., 2000.
Formation of grain-coating chlorite in sandstones. Laboratory synthesized
vs. natural occurrences. Clay Minerals, 35, 261-269.
Aase, N. E., Bjorkum, P. A., and Nadeau, P. H., 1996. The effect of grain-coating
microquartz on preservation of reservoir porosity. AAPG Bulletin, 80, 16541673.
Aase, N. E., and Walderhaug, O., 2005. The effect of hydrocarbons on quartz
cementation; diagenesis in the Upper Jurassic sandstones of the Miller
Field, North Sea, revisited. Petroleum Geoscience, 11, 215-223, doi:
10.1144/1354-079304-648.
Abouessa, A. and Morad, S., 2009. An integrated study of diagenesis and
depositional facies in tidal sandstones: Hawaz Formation (Middle
Ordovician), Murzuq Basin, Libya. Journal of Petroleum Geology, 32, 39-66.
Ahmad, A. H. M., and Bhat, G. M., 2006. Petrofacies, provenance and diagenesis
of the dhosa sandstone member (Chari Formation) at Ler, Kachchh subbasin, Western India. Journal of Asian Earth Sciences, 27, 857-872.
Ajdukiewicz, J. M., Nicholson, P. H., and Esch, W. L., 2010. Prediction of deep
reservoir quality using early diagenetic process models in the Jurassic
Norphlet Formation, Gulf of Mexico. AAPG Bulletin, 94, 1189-1227.
Akaegbobi, I. M., and Adeleye, M. A., 2001. Diagenesis and its implications on the
reservoir quality of parts of Ajalli Sandstone in the Cretaceous Anambra
Basin, southeastern Nigeria. Annual Meeting Expanded Abstracts American
Association of Petroleum Geologists. 2001, 3.
Akarish, A. I. M. and El-Gohary, A. M., 2008. Petrography and geochemistry of
lower Paleozoic sandstones, East Sinai, Egypt: implications for provenance
and tectonic setting. Journal of African Earth Sciences, 52, 43-54.
Al-Gailani, M. B., 1981. Authigenic mineralizations at unconformities: implication for
reservoir characteristics. Sedimentary Geology, 29, 89-115.
Al-Hajri, S. A., Filatoff, J., Wender, L. E., and Norton, A. K., 1999. Stratigraphy and
operational palynology of the Devonian System in Saudi Arabia. GeoArabia
Manama, 4, 53-68.
Al-Harbi, O. A. and Khan, M. M., 2008. Provenance, diagenesis, tectonic setting
and geochemistry of Tawil Sandstone (Lower Devonian) in central Saudi
Arabia. Journal of Asian Earth Sciences, 33, 278-287.
Ali, A. D., and Tuner, P., 1982. Authigenic K-feldspar in the Bromsgrove Sandstone
Formation (Triassic) of central England. Journal of Sedimentary Petrology,
52, 187-198.
Al-Juboury, A. I., McCann, T. and Ghazal, M. M., 2009. Provenance of Miocene
sandstones in northern Iraq: constraints from framework petrography, bulkrock geochemistry and mineral chemistry. Russian Geology and
Geophysics, 50, 517-534.
Allen, D. E., 2008. Origin and distribution of porosity in the Bloyd Sandstone,
southern Boston Mountains, Arkansas. Abstracts with Programs Geological
Society of America, 40, 6.
359

Allen, J. L., and Johnson, C. L., 2010. Facies control on sandstone composition
(and influence of statistical methods on interpretations) in the John Henry
Member, Straight Cliffs Formation, southern Utah, USA. Sedimentary
Geology, 230, 60-76.
Al-Ramadan, K., Morad, S., Proust, J. N. and Al-Aasm, I., 2005. Distribution of
diagenetic alterations in siliciclastic shoreface deposits within a sequence
stratigraphic framework: evidence from the Upper Jurassic, Boulonnais, NW
France. Journal of Sedimentary Research, 75, 943-959.
Al-Ramadan, K. A., Hussain, M., Imam, B. and Saner, S., 2004. Lithologic
characteristics and diagenesis of the Devonian Jauf Sandstone at Ghawar
Field, eastern Saudi Arabia. Marine and Petroleum Geology, 21, 1221-1234.
Amireh, B. S., 1991. Mineral composition of the Cambrian-Cretaceous Nubian
Series of Jordan; provenance, tectonic setting and climatological
implications. Sedimentary Geology, 71, 99-119.
Anani, C., 1999. Sandstone petrology and provenance of the Neoproterozoic
Voltaian Group in the southeastern Voltaian Basin, Ghana. Sedimentary
Geology, 128, 83-98.
Anjos, S. M. C., De Ros, L. F. and Silva, C. M. A., 2003. Chlorite authigenesis and
porosity preservation in the Upper Cretaceous marine sandstones of the
Santos Basin, offshore eastern Brazil. In: Worden, R. H., and Morad, S.
(Eds), Clay Mineral Cements in Sandstones. International Association of
Sedimentologists Special Publication, 34, 291-316.
Anon, 1997. 3-D reservoir characterization of the House Creek Oil Field, Powder
River Basin, Wyoming, U.S. Geological Survey Digital Data Series DDS-33.
http://pubs.usgs.gov/dds/dds-033/USGS_3D/ssx_txt/heterog.htm.
Arditto, P. A., 1987a. Eustasy, sequence stratigraphic analysis and peat formation:
A model for widespread Late Permian coal deposition in the Sydney Basin,
NSW. Proceedings of the twenty first Newcastle Symposium on Advances in
the Study of the Sydney Basin, University of Newcastle, Newcastle,
Australia, 11-17.
Arditto, P. A., 1987b. Potential hydrocarbon plays in the Late Permian succession
of the southern Sydney Basin based on well log and seismic sequence
analysis. Exploration Geophysics 18, 355-366.
Arditto, P. A., 1991. A sequence stratigraphic analysis of the Late Permian
succession in the Southern Coalfield, Sydney Basin, New South Wales.
Australian Journal of Earth Sciences, 38, 125-137.
Arditto, P. A., 2003. High-resolution sequence stratigraphy of the Sydney
Subgroup, southern Sydney Basin; preliminary findings. In: Hutton, A. C.,
Jones, B. G., Carr, P. F., Ackerman, B and Switzer, A. C. (Eds),
Proceedings of the Thirty fifth symposium on Advances in the study of the
Sydney Basin. University of Wollongong, Wollongong, N.S.W., Australia,
333-339.
Armitage, P. J., Worden, R. H., Faulkner, D. R., Aplin, A. C., Butcher, A. R., and
Iliffe, J., 2010. Diagenetic and sedimentary controls on porosity in Lower
Carboniferous fine-grained lithologies, Krechba Field, Algeria; a petrological
360

study of a cap rock to a carbon capture site. Marine and Petroleum Geology,
27, 1395-1410.
Ashley, G. M. and Duncan, I. J., 1977. The Hawkesbury Sandstone; a critical
review of proposed environmental models. Geological Society of Australia,
Journal 24, 117-119.
Asiedu, D. K., Suzuki, S. and Shibata, T., 2000. Provenance of sandstones from
the Lower Cretaceous Sasayama Group, Inner Zone of southwest Japan.
Sedimentary Geology, 131, 9-24.
Bai, G. P., 1988. Diagenesis of the sandstones in the Narrabeen Group. Advances
in the study of the Sydney Basin, Proceedings of the twenty second
sumposium, Department of Geology, the University of Newcastle. pp. 47-53.
Bai, G. P., 1991. Petrology, diagenesis and reservoir potential of Narrabeen Group
sandstones, Sydney Basin, NSW. PhD thesis, University of Sydney,
Sydney.
Bai, G. P., and Keene, J. B., 1996. Petrology and diagenesis of Narrabeen Group
sandstones, Sydney Basin, New South Wales. Australian Journal of Earth
Sciences, 43, 525-538.
Baker, G., 1956. Pelletal claystone from the Southern Coalfield, New South Wales,
Australia. Journal of Science, 18, 126-127.
Baker, J. C., Havord, P. J., Martin, K. R., and Ghori, K. A. R. , 2000. Diagenesis
and petrophysics of the Early Permian Moogooloo Sandstone, southern
Carnarvon Basin, Western Australia. AAPG Bulletin, 84, 250-265.
Bakken, I., Midtbo, R. E. A., and Rykkje, J. M., 2004. The effect of porosity
preserving clay coating on permeability; examples from deeply buried
sandstone reservoirs offshore Norway. Annual Meeting Expanded Abstracts
American Association of Petroleum Geologists, 13, 9.
Balfe, P. E., Draper, J. J., Scott, S. G., and Belcher, R. L., 1988. Bowen Basin solid
geology 1:500 000 map. Geological Survey of Queensland, Brisbane.
Bamberry, W., Hutton, A., and Jones, B., 1989. Depositional systems of the
Illawarra Coal Measures, southern Sydney Basin. Advances in the study of
the Sydney Basin; Proceedings of the Twenty-third Symposium, Department
of Geology, University of Newcastle, 61-68.
Bamberry, W. J., 1992. Stratigraphy and sedimentology of the Late Permian
Illawarra Coal Measures, Southern Sydney Basin, New South Wales. PhD.
thesis, University of Wollongong (unpublished), Wollongong.
Bamberry, W. J., Hutton, A. C., and Jones, B. G., 1995. The Permian Illawarra
Coal Measures, southern Sydney Basin, Australia; a case study of deltaic
sedimentation. In: Oti, J. S. (Eds), Geology of deltas. Balkema Utigevers,
Rotterdam, 153-166.
Bamberry, W. J. and Doyle, R. P., 1987. A tuffaceous origin for the Burragorang
Claystone Member and its lateral equivalents. Advances in the study of the
Sydney Basin, twenty first Symposium proceedings, The University of
Newcastle, New South Wales, Newcastle, 123-127.
Barton, M. D., 1994. Outcrop characterization of architecture and permeability
structure in fluvial deltaic sandstones, Cretaceous Ferron Sandstone, Utah:
Ph.D. thesis, University of Texas at Austin, 259 p.
361

Basu, A., 1976. Petrology of Holocene fluvial sand derived from plutonic source
rocks: implications to paleoclimatic interpretation Journal of Sedimentary
Petrology, 46, 694-709.
Basu, A., 1985. Influence of climate and relief on compositions of sands released
at source areas. In: Zuffa G. G. (Ed). Provenance of Arenites NATO ASI
Series, reidel Publication Company, 1-18.
Basu, A., Young, S. W., Lee, J., Sutter, W., Calvin, J., and Mack, G. H., 1975. Reevaluation of the use of undulatory extinction and polycrystallinity in detrital
quartz for provenance interpretation. Journal of Sedimentary Research, 45,
873-882.
Bayliss, P., Loughnan, F. C., and Standard, J. C., 1965. Dickite in the Hawkesbury
Sandstone of the Sydney Basin, Australia. American Mineralogist, 50, 418426.
Beckett, J., Hamilton, D. S., and Weber, C. R., 1983. Permian and Triassic
stratigraphy and sedimentation in the GunnedahNarrabriCoonabarabran
region. Geological Survey of New South Wales Quarterly Notes, 51, 1-16.
Behrenbruch, P., and Biniwale, S., 2005. Characterisation of clastic depositional
environments and rock pore structures using the Carman-Kozeny equation:
Australian sedimentary basins. Journal of Petroleum Science and
Engineering, 47, 175-196.
Berger, A., Gier, S., and Krois, P., 2009. Porosity-preserving chlorite cements in
shallow-marine volcaniclastic sandstones: evidence from Cretaceous
sandstones of the Sawan gas field, Pakistan. AAPG Bulletin, 93, 595-615.
Berner, R. A., 1981. A review geochemical classification of sedimentary
environments. Journal of Sedimentary Petrology, 51, 359-365.
Bernet, M. and Gaupp, R., 2005. Diagenetic history of Triassic sandstone from the
Beacon Supergroup in central Victoria Land, Antarctica. New Zealand
Journal of Geology and Geophysics, 48, 447-458.
Bertier, P., Swennen, R., Lagrou, D., Laenen, B. and Kemps, R., 2008. Palaeoclimate controlled diagenesis of the Westphalian C & D fluvial sandstones in
the Campine Basin (north-east Belgium). Sedimentology, 55, 1375-1417.
Billault, V., Beaufort, D., Baronnet, A., and Lacharpagne, J. C., 2003. A
nanopetrographic and textural study of grain-coating chlorites in sandstone
reservoirs. Clay Minerals, 38, 315-328.
Bjorkum, P. A., Oelkers, E. H., Nadeau, P. H., Walderhaug, O., and Murphy, W.
M., 1998. Porosity prediction in quartzose sandstones as a function of time,
temperature, depth, stylolite frequency, and hydrocarbon saturation. AAPG
Bulletin, 82, 637-648.
Bjorkum, P. A., and Nadeau, P. H., 1998. Temperature controlled
porosity/permeability reduction, fluid migration, and petroleum exploration in
sedimentary basins. Australian Petroleum Production and Exploration
Association Journal, 38, 453-464.
Bjorlykke, K., 1984. Formation of secondary porosity; how important is it? In:
Surdam, R., Mcdonald, D. A. (Eds), Clastic diagenesis. Association of
Petroleum Geologists Memoir, vol. 37. American Association of Petroleum
Geologists,Tulsa, Oklahoma, 77-286.
362

Bjorlykke, K., and Egeberg, P. K., 1993. Quartz cementation in sedimentary


basins. AAPG Bulletin, 77, 1538-1548.
Bjorlykke, K., Aagaard, P., Dypvik, H., Hastings, D., and Harper, A., 1986.
Diagenesis and reservoir properties of Jurassic sandstones from the
Haltenbanken area, offshore mid-Norway, Norwegian Petroleum Society. In:
Spencer, A. (Eds), Habitat of hydrocarbons on the Norwegian continental
shelf, Norwegian Petroleum Society. Graham and Trotman, pp. 275286.
Bjorlykke, K., Ramm, M., and Saigal, G., 1989. Sandstone diagenesis and porosity
modification during basin evolution. International Journal of Earth Sciences,
68, 1151-1171.
Blatt, H., 1967. Original characteristics of clastic quartz grains. Journal of
Sedimentary Petrology, 37, 401-424.
Blatt, H., 1979. Diagenetic process in sandstones. In: Scholle, P. A., and SellWood, B. W. (Eds), Aspects of Diagenesis. SEPM Special Publication, 26,
141-157.
Blatt, H., Middleton, G., and Murray, R., 1980. Origin of Sedimentary Rocks,
second ed. Prentice Hall Inc., New York
Blevin, J., Hall, L., Chapman, J., and Pryer, L., 2007. Sydney Basin Reservoir
Prediction Study, F/T Project Code: MR705. 1-102.
Bloch, S., Lander, R. H., and Bonnell, L., 2002. Anomalously high porosity and
permeability in deeply buried sandstone reservoirs: origin and predictability.
AAPG Bulletin, 86, 301-328.
Bloomfield, J. P., Moreau, M. F., and Newell, A. J., 2006. Characterization of
permeability distributions in six lithofacies from the Helsby and Wilmslow
sandstone formations of the Cheshire Basin, UK. Geological Society Special
Publications, 263, 83-101.
Bo, T., and Lu, Z. Y., 2000. The characteristics, reservoir quality and types of
lacustrine braided deltas. Sedimentary Geology and Tethyan Geology, 20,
78-84 (in Chinese with English abstract).
Boggs, S., 2001. Principles of Sedimentology and Stratigraphy. Prentice Hall,
London. 726 pp.
Boggs, S., 2009. Petrology of Sedimentary Rocks. Cambridge University Press,
New York. 600 pp.
Boles, J. R., and Franks, S. G., 1979. Clay diagenesis in Wilcox sandstones of
Southwest Texas; implications of smectite diagenesis on sandstone
cementation. Journal of Sedimentary Petrology, 49, 55-70.
Borgohain, P., Borah, C., and Gilfellon, G. B., 2010. Sandstone diagenesis and its
impact on reservoir quality of the Arenaceous Unit of Barail Group of an
oilfield of Upper Assam Shelf, India. CURRENT SCIENCE, 98, 82-88.
Bowman, H. N., 1970. Palaeoenvironment and revised nomenclature of the upper
Shoalhaven Group and Illawarra Coal Measures in the Wollongong-Kiama
area, New South Wales. Records of the Geological Survey of New South
Wales, 12, 163-182.
Bowman, H. N., 1972. Natural slope stability in the City of Greater Wollongong.
Geological Survey of New South Wales Records, 14, 159-222.
363

Bowman, H. N., 1974. Geology of the Wollongong, Kiama and Robertson 1: 50,000
sheets. Geological Survey New South Wales, Sydney, Sydney. pp.179.
Bowman, H. N., 1980. Southern Coalfield, upper Shoalhaven Group and Illawarra
Coal Measures. In: Herbert, C., and Helby, R. (Eds), A Guide to the Sydney
Basin. Geological Survey of New South Wales, Bulletin 26, 117-132.
Brakel, A. T., 1989. Sydney Basin: Permian coal measures-stratigraphy and
sedimentation. In: Harrington H. J., Brakel A. T., Hunt J. W., Wells A. T.,
Middleton M. F., OBrien P. E., Hamilton D. S., Beckett J., Weber C. R.,
Radke S., Totterdell J. M., Swaine D. J. and Schmidt P. W. (Eds), Permian
Coals of Eastern Australia. Bureau of Mineral Resources Bulletin, 231, 9-41.
Branagan, D. F. (1969) The Gosford Formation - Palm Beach to Long Reef.
Abstracts for the Third Symposium on Advances in the study of the Sydney
basin. University of Newcastle.
Branagan, D. F., 2000. The Hawkesbury Sandstone; its origins and later life. In:
McNally, G. H., and Franklin, B. J. (Eds), Sandstone city; Sydney's
dimension stone and other sandstone geomaterials; proceedings of a
symposium held during the 15th Australian geological convention.
Geological Society of Australia Sydney, N.S.W., Australia, 23-38.
Branagan, D. F., Packham, G. H. and Webby, B. D., 1966. Notes on the Narrabeen
Group (Triassic) coastal section north of Long Reef, Sydney Basin.
Australian Journal of Science, 29, 117-118.
Bray, E. E., and Foster, W. R., 1980. A process for primary migration of petroleum.
American Association of Petroleum Geologists Bulletin, 64, 107-111.
Briggs, D. J. C., 1993. Permian depositional sequences of the Sydney Bowen
Basin. Proc. 27th Newcastle Symposium on Advances in the Study of the
Sydney Basin, University of Newcastle, Newcastle, Australia, 247-254.
Bruce, A. G., and Wong, P. M., 2002. Permeability prediction from well logs using
an evolutionary neural network. Petroleum Science and Technology, 20,
317-331.
Bunny, M. R., 1966. Proposed sub-division of the Sydney Sub-group, Southern
Coalfield. Rep. geol. Surv. N.S.W. , GS 1966/005 (unpubl.).
Bunny, M. R., 1972. Geology and coal resources of the Southern catchment coal
reserve, southern Sydney Basin, New South Wales. Geological Survey of
New South Wales, Bulletin 22, 146 pp.
Bunny, M. R., and Herbert, C., 1971. The lower Triassic Newport Formation,
Narrabeen Group, southern Sydney Basin. Records of the Geological
Survey of New South Wales, 13, 61-81.
Burley, S., Kantorowicz JD, Waugh B., 1985. Clastic diagenesis. In: Brenchley PJ,
Williams. BPJ. (Eds), Sedimentology: recent developments and applied
aspects. Geological Society Special Publication, 18, 189-226.
Burley, S. D., and Kantorowicz, J. D., 1986. Thin section and SEM textural criteria
for the recognition of cement-dissolution porosity in sandstones.
Sedimentology, 33, 587-604.
Burnham, A. K., and Braun, R. L., 1990. Development of a detailed model of
petroleum formation, destruction and expulsion from lacustrine and marine
source rocks. Organic Geochemistry, 16, 27-39.
364

Buyukutku, A. G., and Bagci, A. S., 2005. Clay controls on reservoir properties in
sandstone of Kuzgun formation and its relevance to hydrocarbon
exploration, Adana basin, southern Turkey. Journal of Petroleum Science
and Engineering, 47, 123-135.
Buyukutku, A. G., and Sahinturk, O., 2005. The diagenesis of the upper Oligocene
and Mio-Pliocene sandstones and its relevance to hydrocarbon exploration
in the East Anatolia Basin, Turkey. Journal Geological Society of India, 65,
468-478.
Buyukutku, A. G., and Sahinturk, O., 2006. The effect of ductile-lithic sand grains,
overpressure and secondary dissolution on porosity and permeability and
their relevance to hydrocarbon exploration in Askale Sub-Basin, east
Anatolia, Turkey. Energy Sources, Part A, 28, 1027-1038.
Caetano-Chang, M. R., and Wu, F. T., 2003. Diagenese de arenitos da formacao
Piramboia no centro-leste paulista. Translated Title: Sandstone diagenesis
in the Piramboia Formation, east-central Sao Paulo State. Geociencias Sao
Paulo, 33-39.
Cagatay, M. N., Saner, S., Al-Saiyed, I., and Carrigan, W. J., 1996. Diagenesis of
the Safaniya Sandstone Member (mid-Cretaceous) in Saudi Arabia.
Sedimentary Geology, 105, 221-239.
Cameron, R. G., Meakin, N. S., Pogson, D. J., Colquhoun, G. P., Yoo, E. K.,
Barron, L. M., Henderson, G. A. M., Warren, A. Y. E., and Morgan, E. J.,
1999.Sydney-Gunnedah Basin. In: Meakin, N. S., Morgan, E. J., Barron, L.
M., Cameron, R. G., Colquhoun, G. P., Downes, P. M., Glen, R. A.,
Henderson, G. A. M., Jagodzinski, E., Krynen, J. P., Leys, M. T. C.,
Percival, I. G., Pogson, D. J., Raymond, O. L., Robson, D. F., Scott, M. M.,
Watkins, J. J., Warren, A. Y. E., Wyborn, D., Yoo, E. K., Facer, R. A. and
Stewart, J. R. (Eds), Dubbo 1:250,000 geological sheet. Geological Survey
of New South Wales, Dept. of Mineral Resources Sydney, Australia, 281312.
Caracciolo, L., Le Pera, E., Muto, F., and Perri, F., 2011. Sandstone petrology and
mudstone geochemistry of the Peruc-Korycany Formation (Bohemian
Cretaceous Basin, Czech Republic). International Geology Review, 53,
1003-1031.
Carr, P. F., 1983. A reappraisal of the stratigraphy of the upper Shoalhaven Group
and lower Illawarra Coal Measures, southern Sydney Basin, New South
Wales. Proceedings of the Linnean Society of New South Wales, 106, 287297.
Cavazza, W., and Ingersoll, R.V., 2005. Detrital Modes of the Ionian Forearc Basin
fill (Oligocene-Quaternary) reflect the tectonic evolution of the CalabriaPeloritani Terrane (Southern Italy). Journal of Sedimentary Research, 75,
268-279.
Chambers, T., 2006. Stage 2 Drilling and Pilot Testing Program, Upper NepeanMonitoring Well Installation Construction and Piezometer Completion.
Sydney Catchment Authority Report Number: 5065 RP04.

365

Chi, G., Giles, P. S., Williamson, M. A., Lavoie, D., and Bertrand, R., 2003.
Diagenetic history and porosity evolution of Upper Carboniferous
sandstones from the Spring Valley #1 well, Maritimes Basin, Canada
implications for reservoir development. Journal of Geochemical Exploration,
80, 171-191.
Choquette, P. W., and Pray, L., 1970. Geologic nomenclature and classification of
porosity in sedimentary carbonates. American Association Petroleum
Geologists Bulletin, 54, 207-250.
Cocker, J. D., Knox, W. O. B., Lott, G. K., and Milodowski, A. E., 2003. Petrologic
controls on reservoir quality in the Devonian Jauf Formation sandstones of
Saudi Arabia. GeoFrontier, 1, 6-11.
Conaghan, P. J., 1980. The Hawkesbury Sandstone: gross characteristics and
depositional environment. In: Herbert, C., and Helby, R. (Eds), A Guide to
the Sydney Basin. Bulletin Geological Survey of New Sooth Wales, 26, 188253.
Conaghan, P. J., Jones, J. G., Mcdonnell, K. L., and Royce, K., 1982. A dynamic
fluvial model for the Sydney Basin. Journal of the Geological Society of
Australia, 29, 55-70.
Conaghan, P. J., and Jones, J. G., 1975. The Hawkesbury Sandstone and the
Brahmaputra; a depositional model for continental sheet sandstones.
Geological Society of Australia Journal, 22, 275-283.
Conolly, J. R., 1964. Trough cross-stratification in the Hawkesbury Sandstone.
Australian Journal of Science, 27, 113-115.
Conolly, J. R., 1965. Large scale cross-stratification in Triassic sandstone, Sydney,
Australia. Journal of Sedimentary Petrology, 35, 765-768.
Conolly, J. R., 1969. Models for Triassic deposition in the Sydney Basin. Special
Publication Geological Society of Australia, 2, 209-223.
Conolly, J. R., and Ferm, J.C., 1971. Permo-Triassic sedimentation patterns,
Sydney Basin, Australia. American Association of Petroleum Geologists,
Bulletin, 55, 2018-2032.
Cooper, M. R., Evans, J., Flint, S. S., Hogg, A. J. C., and Hunter, R. H. , 2000.
Quantification of detrital, authigenic and porosity components of the
Fontainebleau Sandstone: a comparison of conventional optical and
combined scanning electron microscope-based methods of modal analyses.
Special Publication of the International Association of Sedimentologists, 29,
89-101.
Cowan, D. S., 1974. Deformation and metamorphism of the Franciscan subduction
zone complex northwest of Pacheco Pass, California. Geological Society of
America Buletinl, 85, 1623-1634.
Cowan, E. J., 1993. Longitudinal fluvial drainage patterns within a foreland basinfill; Permo-Triassic Sydney Basin, Australia. Sedimentary Geology, 85, 557577.
Crook, K. A. W., 1957. Cross-stratification and other sedimentary features of the
Narrabeen group. Proceedings of the Linnean Society of New South Wales,
82, 157-166.
366

Curtis, C. D., and Coleman, M. L., 1986. Controls on the precipitation of early
diagenetic calcite, dolomite and siderite concretions in complex depositional
sequences. In: Gautier, D. L. (Eds), Roles of Organic Matter in Sediment
Diagenesis. SEPM, Special Publication 38, 23-33.
Cusack, A., 1991. The Eckersley Formation deposition and diagenesis. B. Sc
(Hons), University of Wollongong, Wollongong.
Dabbagh, M. E., and Rogers, J. J., 1983. Depositional environments and tectonic
significance of the Wajid Sandstone of southern Saudi Arabia. Journal of
African Earth Sciences 1, 47-57.
Das, B. K., and Haake, B, 2003. Geochemistry of Rewalsar Lake sediments,
Lesser Himalaya, India; implications for source-area weathering,
provenance and tectonic setting. Geosciences Journal, 7, 299-312.
Datta, B., 2005. Provenance, tectonics and palaeoclimate of Proterozoic
Chandarpur sandstones, Chattisgarh Basin: a petrographic view. Journal of
Earth System Science, 114, 227-245.
de Ros, L. F., Sgarbi, G. N. C., and Morad, S., 1994. Multiple authigenesis of Kfeldspar in sandstones; evidence from the Cretaceous Areado Formation,
Sao Francisco Basin, central Brazil. Journal of Sedimentary Research, 64,
778-787.
De Ros, L. F., Morad, S., and Paim, P. S. G., 1994. The role of detrital composition
and climate on the diagenetic evolution of continental molasses: evidence
from the Cambro--Ordovician Guaritas sequence, southern Brazil.
Sedimentary Geology, 92, 197-228.
De Waal, J. A., Bil, K. J., Kantorowicz, J. D., and Dicker, A. I. M., 1988.
Petrophysical core analysis of sandstones containing delicate illite. Log
Analyst, 317-331.
Deen, D. J., 1999. Facies analysis and distribution of the Narrabeen Group,
southern Sydney Basin, New South Wales. B. Sc (Hons), University of
Wollongong, Wollongong.
Dehghani, M. H., 1994. Sedimentology, genetic stratigraphy and depositional
environment of the Permo-Triassic succession in the southern Sydney
Basin, Australia. PhD. thesis, University of Wollongong, Wollongong.
Dehghani, M. H., and Jones, B. G., 1994a. Evolution of the Narrabeen Group
succession in the southern Sydney Basin. Geological Society of Australia,
Abstract No. 37, "12th" Australian Geological Convention, September 1994,
Perth, 81.
Dehghani, M. H., and Jones, B. G., 1994b. Sedimentology of Coalcliff Sandstone,
southeastern Sydney Basin: fluvial interpretation based on bounding
surfaces and architectural elements. Proceedings of the Twenty-eighth
Symposium on Advances in the Study of the Sydney Basin, pp. 236-243.
University of Newcastle, Newcastle.
Dehghani, M. H., and Jones, B. G., 1994c. Scarborough Sandstone in the
southeastern Sydney Basin; an Early Triassic sandy to gravelly bed-load
fluvial deposit. Proceedings of the Twenty-eighth Symposium on Advances
in the Study of the Sydney Basin, 244-251. University of Newcastle,
Newcastle.
367

Dey, S., Rai, A. K., and Chaki, A., 2009. Palaeoweathering, composition and
tectonics of provenance of the Proterozoic intracratonic Kaladgi-Badami
Basin, Karnataka, southern India: evidence from sandstone petrography and
geochemistry. Journal of Asian Earth Sciences, 34, 703-715.
Dickey, P. A., 1981. Petroleum Development Geology. Penn Well Publishing
Company, Tulsa, Oklahoma, USA. 2nd ed.
Dickinson, W. R., 1970. Interpreting detrital modes of greywacke and arkose.
Journal of Sedimentary Petrology, 40, 695-707.
Dickinson, W. R., Beard, L. S., Brakenridge, G. R., Erjavec, J. L., Ferguson, R. C.,
Inman, K. F., Knepp, R. A., Lindberg, F. A., and Ryberg, P. T., 1983.
Provenance of North American Phanerozoic sandstones in relation to
tectonic setting. Geological Society of America Bulletin, 94, 222-235.
Dickinson, W. R., 1985. Interpreting provenance relations from detrital modes of
sandstones In: Zuffa, G. G. (Eds), Provenance of Arenites. Nato ASI Series,
Reidel, Dor-drecht (book), 148, 333-361.
Dickson, T. W., 1967. Stratigraphy of the Narrabeen Group in the southern
Coalfield. Geological Survey of N.S.W. Report, GS 1967/355, 8 pp.
(unpubl.).
Dickson, T. W., 1969. Stratigraphy of the Narrabeen Group in the Southern
Coalfields, N.S.W. In Abstracts for the Second Symposium on Advances in
the Study of the Sydney Basin, 1967, pp. 23-24. Department of Geology,
University of Newcastle, Newcastle.
Diessel, C. F. K., 1966. Micro-cross-bedding in symmetrical ripple marks.
Australian Journal of Science, 28, 438.
dos Anjos, S. M. C., De Ros, L. F., de Souza, R. S., de Assis Silva, C. M., and
Sombra, C. L., 2000. Depositional and diagenetic controls on the reservoir
quality of Lower Cretaceous Pendencia sandstones, Potiguar Rift Basin,
Brazil. AAPG Bulletin, 84, 1719-1742.
Doyle, J. D., and Sweet, M.L., 1995. Three-dimensional distribution of lithofacies,
bounding surfaces, porosity, and permeability in a fluvial sandstone - Gypsy
Sandstone of northern Oklahoma. AAPG Bulletin, 79, 70-96.
Doyle, L. J., Carder, K. L., and Stewart, R. G., 1983. The hydraulic equivalence of
micas. Journal of Sedimentary Petrology, 53, 643-648.
Dragovich, D., 2000. Weathering mechanisms and rates of decay of Sydney
dimension sandstone. In: Mcnally, G. H., and Franklin, B. J. (Eds),
Sandstone city; Sydney's dimension stone and other sandstone
geomaterials; Environmental, Engineering and hydrogeology group (EEHG)
Geological Society of Australia. 5, 74-83.
Dutta, P. K., 2007. First-cycle sandstone composition and colour of associated
fine-grained rocks as an aid to resolve Gondwana stratigraphy in peninsular,
India. In: Arribas, J., Critelli, S., Johnsson, M.J. (Eds), Sedimentary
provenance and petrogenesis: perspectives from petrography and
geochemistry. Geological Society of America Special Paper, 420, 241-252.
Dutton, S. P., and Willis, B.J., 1998. Comparison of outcrop and subsurface
sandstone permeability distribution, Lower Cretaceous Fall River Formation,
368

South Dakota and Wyoming. Journal of Sedimentary Research, 68, 890900.


Dutton, S. P., 2008. Calcite cement in Permian deep-water sandstones, Delaware
Basin, west Texas: origin, distribution, and effect on reservoir properties.
AAPG Bulletin, 92, 765-787.
Dutton, S. P., and Diggs, T. N., 1990. History of quartz cementation in the Lower
Cretaceous Travis Peak Formation, East Texas. Journal of Sedimentary
Research, 60, 191-202.
Dutton, S. P., and Loucks, R. G., 2009. Diagenetic controls on evolution of porosity
and permeability in lower Tertiary Wilcox Sandstone from shallow to
ultradeep (200-6700 m) burial, Gulf of Mexico Basin, U.S.A. Marine and
Petroleum Geology, 27, 69-81.
Eadington, P. J., Hamilton, P. J., and Bai, G. P., 1991. Fluid history analysis - a
new concept for prospect evaluation Australian Petroleum Exploration
Association, Journal, 31, 282-294.
Ehrenberg, S. N., 1990. Relationship between diagenesis and reservoir quality in
sandstones of the Garn Formation, Haltenbanken, mid-Norwegian
continental shelf. AAPG Bulletin, 74, 1538-1558.
Ehrenberg, S. N., Aagaard, P., Wilson, M. J., Fraser, A. R., and Duthie, D. M. L.,
1993. Depth-dependant transformation of kaolinite to dickite in sandstone of
the Norwegian continental shelf. Clay minerals, 28, 325-352.
Ehrenberg, S. N., 1993. Preservation of anomalously high porosity in deeply buried
sandstones by grain-coating chlorite: examples from the Norwegian
continental shelf. AAPG Bulletin, 77, 1260-1286.
Ehrenberg, S. N., and Jakobsen, K. G., 2001. Plagioclase dissolution related to
biodegradation of oil in Brent Group sandstones (Middle Jurassic) of
Gullfaks Field, northern North Sea. Sedimentology, 48, 703-721.
Ehrenberg, S. N., Nadeau, P. H., and Steen, O., 2008. A megascale view of
reservoir quality in producing sandstones from the offshore Gulf of Mexico.
AAPG Bulletin, 92, 145-164.
El-Ghali, M. A. K., Mansurbeg, H., Morad, S., Al-Aasm, I., and Ajdanlisky, G.,
2006a. Distribution of diagenetic alterations in fluvial and paralic deposits
within sequence stratigraphic framework: Evidence from the Petrohan
Terrigenous Group and the Svidol Formation, Lower Triassic, NW Bulgaria.
Sedimentary Geology, 190, 299-321.
El-Ghali, M. A. K., Mansurbeg, H., Morad, S., Al-Aasm, I., and Ramseyer, K.,
2006b. Distribution of diagenetic alterations in glaciogenic sandstones within
a depositional facies and sequence stratigraphic framework: evidence from
the Upper Ordovician of the Murzuq Basin, SW Libya. Sedimentary
Geology, 190, 323-351.
El-Ghali, M. A. K., Morad, S., Mansurbeg, H., Caja, M. A., Ajdanlijsky, G., Ogle, N.,
Al-Aasm, I., and Sirat, M., 2009a. Distribution of diagenetic alterations within
depositional facies and sequence stratigraphic framework of fluvial
sandstones: Evidence from the Petrohan Terrigenous Group, Lower
Triassic, NW Bulgaria. Marine and Petroleum Geology, 26, 1212-1227.
369

El-Ghali, M. A. K., Morad, S., Mansurbeg, H., Caja, M. A., Sirat, M., and Ogle, N.,
2009b. Diagenetic alterations related to marine transgression and
regression in fluvial and shallow marine sandstones of the Triassic
Buntsandstein and Keuper sequence, the Paris Basin, France. Marine and
Petroleum Geology, 26, 289-309.
El-Ghali, M. A. K., Tajori, K. G., Mansurbeg, H., Ogle, N., and Kalin, R. M., 2006c.
Origin and timing of siderite cementation in Upper Ordovician glaciogenic
sandstones from the Murzuq Basin, SW Libya. Marine and Petroleum
Geology, 23, 459-471.
Elias, A. R. D., De Ros, L. F., Mizusaki, A. M. P., and Anjos, S. M. C., 2004.
Diagenetic patterns in eolian/coastal sabkha reservoirs of the Solimoes
Basin, northern Brazil. Sedimentary Geology, 169, 191-217.
Emerson, D. W., 2000. The petrophysics of the Hawkesbury Sandstone; a
preliminary study. In: Mcnally, G. H., and Franklin, B. J. (Eds), Sandstone
city; Sydney's dimension stone and other sandstone geomaterials;
proceedings of a symposium held during the 15th Australian geological
convention. Geological Society of Australia Sydney, N.S.W., Australia, 5,
197-213.
Espejo, I. S., and Lopez-Gamundi, O. R., 1994. Source versus depositional
controls on sandstone composition in a foreland basin; the El Imperial
Formation (Mid Carboniferous-Lower Permian), San Rafael Basin, western
Argentina. Journal of Sedimentary Research, 64, 8-16.
Espitalie, J., Ungerer, P., Marquis, F., and Irwin, H., 1988. Primary cracking of
kerogens: experimental and kinetic modelling. In: Matavelli, L., and Novelli,
L. (Eds), Advances in Organic Geochemistry 1987. 13, 893-899.
Estupinan, J., Marfil, R., Delgado, A., and Permanyer, A., 2007. The Impact of
carbonate cements on the reservoir quality in the Napo Formation
sandstones (Cretaceous Oriente Basin, Ecuador). Geologica Acta, 5, 89107.
Evans, R., Mory, A. J., and Tait, A. M., 2007. An outcrop gamma ray study of the
Tumblagooda Sandstone, Western Australia. Journal of Petroleum Science
and Engineering, 57, 37-59.
Exon, N. F., 1976. Geology of the Surat Basin in Queensland. Bureau of Mineral
Resources Bulletin, 166.
Faiz, M. M., 1993. Thermal history and geological controls on the distribution of
coal seam gases in the Southern Sydney Basin, Australia. PhD thesis,
University of Wollongong, Wollongong.
Folk, R. L., 1968. Petrology of Sedimentary Rocks. University of Texas, Austin,
U.S.A. 170 pp.
Fontanelli, P. D. R., De Ros, L. F., and Remus, M. V. D., 2009. Provenance of
deep-water reservoir sandstones from the Jubarte oil field, Campos Basin,
Eastern Brazilian Margin. Marine and Petroleum Geology, 26, 1274-1298.
Forgotson, J. M., Liu, H., Blatt, H., Suneson, N. H., and Schreiner, W. P., 2000.
Influence of initial calcite grains and diagenesis on porosity development in
Spiro Sandstone of South Haleyville, South Hartshorne, and South Panola
370

fields, Pittsburg and Latimer counties, Oklahoma. Oklahoma Geological


Survey Circular 103, 141-155.
Forkner, R. M., Pate, K. A., Higginbotham, L., Himstedt, J., and Dworkin, S. I.,
2000. Provenance and diagenesis of the Wolf City Sandstone, central
Texas, U.S.A. Abstracts with Programs Geological Society of America, 32,
305.
Franklin, B. J., 2000. Sydney dimension sandstone: The value of petrography in
Stone selection and assessing durability. In: Mcnally, G. H., and Franklin,
B.J. (Eds), Sandstone City Sydney, dimension stone and other sandstone
geomaterials. Environmental engineering and hydrogeology group (EEHG)
Geological society of Australia, 98-117.
Franklin, S. P., and Tieh, T. J., 1989. Petrography, diagenesis, and reservoir
properties of the Dakota Sandstone of West Lindrith Field, Rio Srriba
County, New Mexico. In: Colson, E. B. (Eds), Petrogenesis and
petrophysics of selected sandstone reservoirs in the Rocky Mountain region.
Rocky Mountain Association of Geologists, Denver, 117-133.
Freed, S. J., 2005. The reservoir characteristics of the Hawkesbury Sandstone in
the Southern Highlands in relation to Sydney Water Shortages. BEnvSc
Honours Thesis, University of Wollongong, Wollongong.
Freeze, R. A., and Cherry, J. A., 1979. Groundwater. Prentice Hall, Englewood
Cliffs, New Jersey.
Frohlich, S., Redfern, J., Petitpierre, L., Marshall, J. D., Power, M., and Grech, P.
V., 2010. Diagenetic evolution of incised channel sandstones; implications
for reservoir characterisation of the Lower Carboniferous Marar Formation,
Ghadames Basin, western Libya. Journal of Petroleum Geology, 33, 3-18.
Galloway, M. C., 1972. Statistical analyses of regional heavy mineral variation,
Hawkesbury Sandstone and Narrabeen Group (Triassic), Sydney Basin.
Journal of the Geological Society of Australia, 19, 65-76.
Gentz, M. L., 2006. A pre-mining study of the Hawkesbury Sandstone and aquifer
characeristics of potential longwall mining area, Appin area 3. BEnvSc
Honours thesis, University of Wollongong, Wollongong.
Getaneh, W., 2002. Geochemistry provenance and depositional tectonic setting of
the Adigrat Sandstone, northern Ethiopia. Journal of African Earth Sciences,
35, 185-198.
Ghazi, S., and Mountney, N. P., 2011. Petrography and provenance of the Early
Permian fluvial Warchha Sandstone, Salt Range, Pakistan. Sedimentary
Geology, 233, 88-110.
Gier, S., Worden, R. H., Johns, W. D., and Kurzweil, H., 2008. Diagenesis and
reservoir quality of Miocene sandstones in the Vienna Basin, Austria. Marine
and Petroleum Geology, 25, 681-695.
Giles, M. R., 1987. Mass transfer and problems of secondary porosity creation in
deeply buried hydrocarbon reservoirs. Marine and Petroleum Geology, 4,
188-201.
Giles, M. R., Indrelid, S. L., Beynon, G. V., and Amthor, J., 2000. The origin of
large-scale quartz cementation: evidence from large data sets and coupled
371

heat-fluid mass transport modelling. Special Publication of the International


Association of Sedimentologists, 29, 21-38.
Girard, J.-P., Munz, I. A., Johansen, H., Hill, S., and Canham, A., 2001. Conditions
and timing of quartz cementation in Brent reservoirs, Hild Field, North Sea:
constraints from fluid inclusions and SIMS oxygen isotope microanalysis.
Chemical Geology, 176, 73-92.
Goldbery, R., and Holland, W. N., 1973. Stratigraphy and sedimentation of redbed
facies in Narrabeen Group of Sydney Basin, Australia. The American
Association of Petroleum Geologists Bulletin, 57, 1314-1334.
Gonzalez-Acebron, L., Arribas, J., and Mas, R., 2007. Provenance of fluvial
sandstones at the start of Late Jurassic-Early Cretaceous rifting in the
Cameros Basin (N. Spain). Sedimentary Geology, 202, 138-157.
Gonzalez-Acebron, L., Arribas, J., and Mas, R., 2010. Role of sandstone
provenance in the diagenetic albitization of feldspars: a case study of the
Jurassic Tera Group sandstones (Cameros Basin, NE Spain). Sedimentary
Geology, 229, 53-63.
Gorniak, K., 1997. The role of diagenesis in the formation of kaolinite raw materials
in the Santonian sediments of the North-Sudetic Trough (Lower Silesia,
Poland). Applied Clay Science, 12, 313-328.
Gotze, J., and Blankenburg, H.-J., 1994. The genesis of the Hohenbocka quartz
sand (Eastern Germany) - new results of mineralogical and geochemical.
Zentralbl. Geol. Palaontol. Teil 1 1992, 1217-1231.
Gould, K., Pe-Piper, G., and Piper, D. J. W., 2010. Relationship of diagenetic
chlorite rims to depositional facies in Lower Cretaceous reservoir
sandstones of the Scotian Basin. Sedimentology, 57, 587-610.
Grades, Z. E. D. A., 2003. Sedimentology and reservoir geology of the MiddleUpper Cretaceous strata in Unity and Heglig Fields in SE Muglad Rift Basin,
Sudan. Doctor rerum naturalium, der Technischen Universitt
Bergakademie Freiberg eingereichte.
Grain, E. R., 2000. Myth-interpretation. Canadian Well Logging Society,
ross@spec2000.net 403-845-2527.
Greene, T. J., Carroll, A. R., Wartes, M., Graham, S. A., and Wooden, J. L., 2005.
Integrated provenance analysis of a complex orogenic terrane: Mesozoic
uplift of the Bogda Shan and inception of the Turpan-Hami Basin, NW
China. Journal of Sedimentary Research, 75, 251-267.
Grevenitz, P., Carr, P., and Hutton, A., 2003. Origin, alteration and geochemical
correlation of Late Permian airfall tuffs in coal measures, Sydney Basin,
Australia. International Journal of Coal Geology, 55, 27-46.
Grey, I., and Ross, J., 2003. Groundwater Investigation for Contingency Drought
Relief in the Sydney Region, Sydney Water Corporation and Sydney
Catchment Authority, Parsons Brinckerhoff.
Griffith, A., 1986. Fluvial architecture, sedimentology, petrology and diagenesis of
the Hawkesbury Sandstone. B.Sc (Hons), University of Wollongong,
Wollongong.

372

Grigsby, J. D., 2001. Origin and growth mechanism of authigenic chlorite in


sandstones of the lower Vicksburg Formation, south Texas. Journal of
Sedimentary Research, 71, 27-36.
Hall, J. S., Mozley, P., Davis, J. M., and Roy, N. D., 2004. Environments of
formation and controls on spatial distribution of calcite cementation in PlioPleistocene fluvial deposits, New Mexico, USA. Journal of Sedimentary
Research, 74, 643-653.
Hallsworth, C. R., and Chisholm, J. I., 2008. Provenance of Late Carboniferous
sandstones in the Pennine Basin (UK) from combined heavy mineral, garnet
geochemistry and palaeocurrent studies. Sedimentary Geology, 203, 196212.
Hamilton, D. S., Galloway, W. E., and Reynolds, S. A., 1987. Depositional systems
of the Narrabbeen Group. Proceedings of the 21st Symposium on Advances
in the Study of the Sydney Basin, University of Newcastle, 115-122.
Hamilton, D. S., and Galloway, W. E., 1989. New exploration techniques in the
analysis of diagenetically complex reservoir sandstones, Sydney Basin,
NSW. The APEA Journal, 29, 235-257.
Hammer E., Mork, M. B. E., and Naess, A., 2010. Facies controls on the
distribution of diagenesis and compaction in fluvial-deltaic deposits. Marine
and Petroleum Geology, 27, 1737-1751.
Hanlon, F. N., Osborne, G.D., and Raggatt, H.G., 1953. The Narrabeen Group - its
subdivisions and correlation between the south coast and the Narrabeen Wyong district. Journal and Proceedings of the Royal Society of New South
Wales, 87, 106-120.
Hanlon, F. N., 1956. Geology of the southern Coalfield. Illawarra district. Annual
Report, Department of Mines N.S.W., for 1952, 95-102.
Harber, L. F., 1915. Geology and mineral resources of the Southern Coalfield.
Memoir of the Geological Survey of N.S.W., Geology, 7.
Harris, R. G. and Bustin, R. M., 2000. Diagenesis, reservoir quality, and production
trends of Doig Formation sand bodies in the Peace River area of western
Canada. Bulletin of Canadian Petroleum Geology, 48, 339-359.
Hartmann, B. H., Juhasz-Bodnar, K., Ramseyer, K., and Matter, A., 2000.
Polyphased quartz cementation and its sources: a case study from the
Upper Palaeozoic Haushi Group sandstones, Sultanate of Oman. Special
Publication of the International Association of Sedimentologists, 29, 253270.
Hartmann, L. A., Endo, I., Suita, M. T. F., Santos, J. O. S., Frantz, J. C., Carneiro,
M. A., Mcnaughton, N. J., and Barley, M. E., 2006. Provenance and age
delimitation of Quadrilatero Ferrifero sandstones based on zircon U-Pb
isotopes. Journal of South American Earth Sciences, 20, 273-285.
Haworth, R. J., 2003. The shaping of Sydney by its urban geology. Quaternary
International, 103, 41-55.
Heald, M. T., and Larese, R. E., 1974. Influence of coatings on quartz cementation.
Journal of Sedimentary Research, 44, 1269-1274.

373

Henry, D. J., and Dutrow, B. L., 1992. Tourmaline in a low grade clastic
metasedimentary rock: an example of the petrographic potential of
tourmaline. Contributions to Mineralogy and Petrology, 112, 203-218.
Herbert, C., and Uren, R., 1972. Duffys Forest shale deposit - proposed drilling
programme. Geological Survey of N.S.W., Report, GS1972/199.
Herbert, C., 1980a. Wianamatta Group and Mittagong Formation. In: Herbert, C.,
and Helby, R. (Eds), A Guide to the Sydney Basin. Bulletin Geological
Survey of New South Wales, 26, 254-272.
Herbert, C., 1980b. Narrabeen Group at North Head, Port Jackson. Quarterly
Notes Geological Survey of New South Wales, 38, 1-7.
Herbert, C., 1983. Geology of the Sydney 1:100,000 sheet 9130. Geological
Survey of New South Wales,
Herbert, C., 1987. Petroleum prospectivity in the Sydney Basin. Advances in the
Study of the Sydney Basin Proceedings of the Symposium, 21, 107-113.
Herbert, C., 1993a. Marine environments in the Early to Middle Triassic Narrabeen
Group. Proceeding of the 27th Symposium on Advances in the Study of the
Sydney Basin, 61-68. University of Newcastle, Newcastle.
Herbert, C., 1993b. Early to Middle Triassic fluvial to marine transition in the
Sydney basin. In: Fifth International Conference on Fluvial Sedimentology
Brisbane. 49.
Herbert, C., 1995. Sequence stratigraphy of the Late Permian coal measures in the
Sydney Basin. Australian Journal of Earth Sciences, 42, 391-405.
Herbert, C., 1996. Sedimentary environments in the Early to Middle Triassic
Narrabeen Group. In: Bamberry, W. J., and Herbert, C. (Eds), Fluvial
Successions of the Late Permian to Triassic rocks of the northern Sydney
Basin. 40-56. Australasian Sedimentologists Group Field Guide 9.
Geological Society of Australia, Sydney.
Herbert, C., 1997. Sequence stratigraphic analysis of Early and Middle Triassic
alluvial and estuarine facies in the Sydney Basin, Australia. Australian
Journal of Earth Sciences, 44, 125-143.
Heritage, Y., 2005. Oil in the Douglas Park Syncline, Southern Coalfield, NSW.
B.Sc. (Hons) thesis, Wollongong University, Wollongong.
Hiatt, E. E., Kyser, T. K., Fayek, M., Polito, P., Holk, G. J., and Riciputi, L. R., 2007.
Early quartz cements and evolution of paleohydraulic properties of basal
sandstones in three Paleoproterozoic continental basins: evidence from in
situ 18O analysis of quartz cements. Chemical Geology, 238, 19-37.
Hiatt, E. E., Kyser, K., and Dalrymple, R. W., 2003. Relationships among
sedimentology, stratigraphy, and diagenesis in the Proterozoic Thelon
Basin, Nunavut, Canada: implications for paleoaquifers and sedimentaryhosted mineral deposits. Journal of Geochemical Exploration, 80, 221-240.
Higgs, K. E., Arnot, M. J., Browne, G. H., and Kennedy, E. M., 2010. Reservoir
potential of Late Cretaceous terrestrial to shallow marine sandstones,
Taranaki Basin, New Zealand. Marine and Petroleum Geology, 27, 18491871.
Higgs, K. E., Zwingmann, H., Reyes, A. G., and Funnell, R. H., 2007. Diagenesis,
porosity evolution, and petroleum emplacement in tight gas reservoirs,
374

Taranaki Basin, New Zealand. Journal of Sedimentary Research, 77, 10031025.


Hill, M. B. L., Armstrong, M., Cozens, S. and Byrnes, J., 1994. Tracing the Bulli and
Balgownie seams across the Sydney Basin. Procedings of the 28th
Newcastle Symposium on Advances in the Study of the Sydney Basin,
University of Newcastle, Newcastle, Australia, 142-150.
Hillier, S., Fallick, A. E., and Matter, A., 1996. Origin of pore-lining chlorite in the
aeolian Rotliegend of northern Germany. Clay Minerals, 31, 153-171,
DOI:10.1180/CLAYMIN.1996.031.2.02.
Hindrix, M. S., 2000. Evolution of Mesozoic sandstone compositions, southern
Junggar, northern Tarim, and western Turpan Basins, northwest China; a
detrital record of the ancestral Tian Shan. Journal of Sedimentary Research,
70, 520-532.
Hlal, O. A., 2008. Diagenesis and reservoir-quality evolution of paralic, shallow
marine and fluvio-lacustrine deposits. links to depositional facies and
sequence
stratigraphy.
Acta
Universitaties
Upsaliensis.
Digital
comprehensive summaries of Uppsala dissertations from the Faculty of
Science and Technology 448, 65 pp. Uppsala. ISBN 978-91-554-7237-5.
Hossain, H. M. Z., Roser, B. P., and Kimura, J. I., 2010. Petrography and wholerock geochemistry of the Tertiary Sylhet succession, northeastern Bengal
Basin, Bangladesh: provenance and source area weathering. Sedimentary
Geology, 228, 171-183.
Hower, J. E., Eslinger, E., Hower, M. E., and Perry, E. A., 1976. Mechanism of
burial metamorphism of argillaceous sediment: 1. Mineralogical and
chemical evidence. Geological Society of America Bulletin, 87, 725-737.
Hubert, J. F., Reed, A. A., and Carey, P. J., 1976. Paleogeography of the East
Berlin Formation, Newark Group, Connecticut Valley. American Journal of
Science, 276, 1183-1207.
Hutton, A., Johnson, M., and Gentz, M., 2006. Petrography of Hawkesbury
Sandstone and water movement. Proceedings of the Thirty-sixth
Symposium on Advances in the Study of the Sydney Basin, 95-102.
Hutton, A. C., 1990. Geophysical logs and correlation in the southern Sydney
Basin. Proc. 24th Newcastle Symposium on Advances in the Study of the
Sydney Basin, University of Newcastle, Newcastle, Australia,17-24.
Hutton, A. C., and Bamberry, J., 1999. Stratigraphy and terminology for the
southern Coalfield. Coalfield Geology Council of New South Wales. Bull, 1,
120-152.
Hutton, A. C., Bamberry, W. J., and Jones, B. G., 1992. A new stratigraphy for the
Illawarra Coal Measures, southern Sydney Basin. In: Jessell, M. W. (Eds),
Earth sciences, computers and the environment. Abstracts Geological
Society of Australia, 32, 111-112.
Irwin, H., and Hurst, A. R., 1983. Application of geochemistry to sandstone
reservoir studies. In: Brooks, J. (Eds), Petroleum Geochemistry and
Exploration of Europe. 12. Geological society of Londdon, 127-146 (special
publication).
375

Islam, M. A., 2009. Diagenesis and reservoir quality of Bhuban sandstones


(Neogene), Titas Gas Field, Bengal Basin, Bangladesh. Journal of Asian
Earth Sciences, 35, 89-100.
Jahren, J., and Ramm, M., 2000. The porosity-preserving effects of
microcrystalline quartz coatings in arenitic sandstones; examples from the
Norwegian continental shelf. In: Worden, R. H., and Morad, S. (Eds), Quartz
Cementation in Sandstones. International Association of Sedimentologists
Special Publication, 29, 271-280.
Jaquet, J. B., Card, G.W., and Harper, L.F., 1905. The geology of the Kiama Jamberoo district. Geological Survey of N.S.W. Record, 8, 1-66.
Jargal, L., and Lee, Y. I., 2006. Detrital modes of the east Gobi Basin (Ondor-Bogd
area) sandstones (Late Jurassic-Early Cretaceous) in southeastern
Mongolia and their geological implications. Geosciences Journal, 10, 1-16.
Jesus, J., 1999. The Wongawilli Coal of the Appin region Sydney Basin, New
South Wales. B. Sc (Hons) thesis, University of Wollongong, Wollongong.
Jia, J., and Gu, J., 2002. Control factors and porosity evolution of high-quality
sandstone reservoirs of Kela-2 gas field in Kuqa Depression. Chinese
Science Bulletin, 47, 100-106.
Jin, Z., Li, F., Cao, J., Wang, S., and Yu, J., 2006. Geochemistry of Daihai Lake
sediments, Inner Mongolia, north China; implications for provenance,
sedimentary sorting, and catchment weathering. Geomorphology, 80, 147163.
Johnson, D. P., 1995. Palaeoclimate and depositional settings of Australian coal
measures. In: Ward, C. R., Harrington, H.J., Mallet, C.W., and Beeston,
J.W. (Eds), Geology of Australian Coal Basins. Geological Society of
Australia Coal Geology Group. Special Publication 1.
Johnson, M. D., 2006. Solutional weathering of the Hawkesbury Sandstone and
cliff instability. BEnvSc (Hons) Thesis, University of Wollongong,
Wollongong.
Jones, B. G., 1983. Illawarra Coal Measures, Sydney Basin, N.S.W. In: Williams,
B. P. J., and Moore, P. S. (Eds), Fluvial Sedimentology Workshop; papers.
Geological Society of Australia, Australasian Sedimentologists Specialist
Group Adelaide, South Aust., Australia, 366-375.
Jones, B. G., and Rust, B.R., 1983. Massive sandstone facies in the Hawkesbury
Sandstone, a Triassic fluvial deposit near Sydney, Australia. Journal of
Sedimentary Petrology, 53, 1249-1260.
Jones, B. G., 1986. Southern Sydney basin field excursion. Department of
Geology, University of Wollongong (unpublished), Wollongong.
Jones, B. G., 1990. The Shoalhaven Group: implication for subsequent coal
measures deposition in the southern Sydney Basin. In: Hutton a, C., and
Depers, A. M. (Eds), Proceeding of Workshop on Southern and Western
Coalfield of the Sydney Basin, University of Wollongong. 11-18.
Joplin, G. A., Hanlon, F.N., and Noakes, L. C., 1952. Wollongong - 4 miles
Geological series. Explanatory Notes, Bureau of Mineral Resources
Geology and Geophysics Australia.
376

Karim, A., Pe-Piper, G., and Piper, D. J. W., 2010. Controls on diagenesis of Lower
Cretaceous reservoir sandstones in the western Sable Subbasin, offshore
Nova Scotia. Sedimentary Geology, 224, 65-83.
Keller, W. D., 1970. Environmental aspects of clay minerals. Journal of
Sedimentary Petrology, 40, 788-813.
Kennedy, J. L., 1999. A detailed facies analysis of the Sydney Subgroup, Late
Permian Illawarra Coal Measures in the eastern Southern Coalfield,
Wollongong, NSW. B. Sc (Hons), University of Wollongong, Wollongong.
Ketzer, J. M., Holz, M., Morad, S., and Al-Aasm, I. S., 2003a. Sequence
stratigraphic distribution of diagenetic alterations in coal-bearing, paralic
sandstones: evidence from the Rio Bonito Formation (early Permian),
southern Brazil. Sedimentology, 50, 855-877.
Ketzer, J. M., and Morad, S., 2006. Predictive distribution of shallow marine, lowporosity (pseudomatrix-rich) sandstones in a sequence stratigraphic
framework - example from the Ferron Sandstone, Upper Cretaceous, USA.
Marine and Petroleum Geology, 23, 29-36.
Ketzer, J. M., Morad, S., and Amorosi, S., 2003b. Predictive diagenetic claymineral distribution in siliciclastic rocks within a sequence stratigraphic
framework. In: Worden, R. H., and Morad, S. (Eds), Clay Mineral Cements
in Sandstones. : International Association of Sedimentologists, Special
Publication, 34, Blackwell, Oxford, 42-59.
Ketzer, J. M., Morad, S., Evans, R., and Al-Aasm, I. S., 2002. Distribution of
diagenetic alterations in fluvial, deltaic, and shallow marine sandstones
within a sequence stratigraphic framework: evidence from the Mullaghmore
Formation (Carboniferous), NW Ireland. Journal of Sedimentary Research,
72, 760-774.
Kilda, L., and Friis, H., 2002. The key factors controlling reservoir quality of the
Middle Cambrian Deimena Group sandstone in West Lithuania. Bulletin of
the Geological Society of Denmark, 49, 25-39.
Kim, J. C., Lee, Y. I., and Hisada, K.-I., 2007. Depositional and compositional
controls on sandstone diagenesis, the Tetori Group (Middle Jurassic-Early
Cretaceous), central Japan. Sedimentary Geology, 195, 183-202.
Kim, Y., and Lee, Y. I., 2004. Diagenesis of shallow marine sandstones, the Lower
Ordovician Dongjeom Formation, Korea: response to relative sea-level
changes. Journal of Asian Earth Sciences, 23, 235-245.
Klett, M., Eichorst, F., and Schaefer, A., 2002. Facies interpretation from well logs
applied to the Tertiary Lower Rhine Basin fill. Netherlands Journal of
Geosciences/Geologie en Mijnbouw, 81, 167-176.
Kok, M. V., Gokcal, B., and Ersoy, G., 2005. Reservoir analysis by well log data.
Energy Sources, 27, 399-404.
Kordi, M., Turner, B., and Salem, A. M. K., 2011. Linking diagenesis to sequence
stratigraphy in fluvial and shallow marine sandstones: evidence from the
Cambrian-Ordovician lower sandstone unit in southwestern Sinai, Egypt.
Marine and Petroleum Geology, 28, 1554-1571.

377

Korsch, R. J., 1984. Sandstone compositions from the New England Orogen,
eastern Australia; implications for tectonic setting. Journal of Sedimentary
Petrology, 54, 192-211.
Krainer, K., and Spotl, C., 1989. Detrital and authigenic feldspar in Permian and
Early Triassic sandstones, eastern Alps (Australia). Sedimentary Geology,
62, 59-77.
Krassay, A. A., Korsch, R. J., and Drummond, B. J., 2009. Meandarra Gravity
Ridge: symmetry elements of the gravity anomaly and its relationship to the
Bowen - Gunnedah - Sydney Basin system. Australian Journal of Earth
Sciences, 56, 355-379.
Kunledare, M., 2007. The "buckyball effect"; framework grain stabilization and
secondary porosity preservation in the Cambrian Galesville Sandstone,
Illinois Basin. Abstracts Annual Meeting American Association of Petroleum
Geologists. 2007, 78.
Kwon, Y.-I., and Boggs, S., 2002. Provenance interpretation of Tertiary sandstones
from the Cheju Basin (NE East China Sea): a comparison of conventional
petrographic and scanning cathodoluminescence techniques. Sedimentary
Geology, 152, 29-43.
Land, L. S., and Dutton, S. P., 1978. Cementation of a Pennsylvanian deltaic
sandstone; isotopic data. Journal of Sedimentary Petrology, 48, 1167-1176.
Langford, R., and Depret, P.-A., 2001. Stratigraphic controls on diagenesis and
porosity in the Cedar Mesa Sandstone, southeastern Utah. Annual Meeting
Expanded Abstracts American Association of Petroleum Geologists. 2001,
112.
Lanson, B., Beaufort, D., Berger, G., Bauer, A., Cassagnabere, A., and Meunier,
A., 2002. Authigenic kaolin and illitic minerals during burial diagenesis of
sandstones: a review. Clay Minerals, 37, 1-22.
Lee, J., 2000. Hydrogeology of the Hawkesbury Sandstone in the southern
highlands of NSW in relation to Mesozoic horst-graben tectonics and
stratigraphy. In: Boyd, R., Diessel, C. F. K., and Francis, S. (Eds),
Proceedings of the Thirty fourth Newcastle Symposium on Advances in the
Study of the Sydney Basin. University of Newcastle, N.S.W., Department of
Geology Newcastle, N.S.W., Australia, 34, 131-144.
Lee, Y. I., and Lim, D. H., 2008. Sandstone diagenesis of the Lower Cretaceous
Sindong Group, Gyeongsang Basin, southeastern Korea: implications for
compositional and paleoenvironmental controls. Island Arc, 17, 152-171.
Lemon, N. M., and Cubitt, C. J., 2003. Illite fluorescence microscopy: a new
technique in the study of illite in the Merrimelia Formation, Cooper Basin,
Australia. Clay Mineral Cements in Sandstones, 34, 411-424.
Li, C., Xie, Y., Liu, S., Huang, H., and Li, M., 2002. Factors controlling the ultra low
porosity and permeability sandstone reservoir rocks of the Yanchang
Formation in Fuxian area, north Shaanxi, China. Chengdu Ligong Xueyuan
Xuebao = Journal of Chengdu Institute of Technology, 29, 285-289.
Li, R., Lib, S., Jin F., Wan, Y., and Zhang, S., 2004. Provenance of Carboniferous
sedimentary rocks in the northern margin of Dabie Mountains, central China
and the tectonic significance; constraints from trace elements, mineral
378

chemistry and SHRIMP dating of zircons. Sedimentary Geology, 166, 245264.


Lima, R. D., and De Ros, L. F., 2002. The role of depositional setting and
diagenesis on the reservoir quality of Devonian sandstones from the
Solimoes Basin, Brazilian Amazonia. Marine and Petroleum Geology, 19,
1047-1071.
Liu, K., Boult, P., Painter, S., and Paterson, L., 1996. Outcrop analog for sandy
braided stream reservoirs; permeability patterns in the Triassic Hawkesbury
Sandstone, Sydney Basin, Australia. AAPG Bulletin, 80, 1850-1866.
Loughnan, F. C., 1962. Some tonstein-like rocks from New South Wales, Australia.
Neues Jahrbuch fuer Mineralogie, Abh, 99, 29-44.
Loughnan, F. C., 1963. A petrological study of a vertical section of the Narrabeen
Group at Helensburgh, N.S.W. Journal of the Geological Society of
Australia, 10, 177-192.
Loughnan, F. C., 1969. Gosford Formation, b) red beds in the geology of New
South Wales. Journal of the Geological Society of Australia, 16, 403-404.
Loughnan, F. C., 1970. Flint clay in the coal-barren Triassic of the Sydney Basin,
Australia. Journal of Sedimentary Petrology, 40, 822-828.
Loughnan, F. C., 1971. Kaolinite claystones associated with the Wongawilli Seam
in the southern part of the Sydney Basin. Journal of the Geological Society
of Australia, 18, 293-302.
Loughnan, F. C., Goldbery, R., and Holland, W. N., 1974. Kaolinite clayrocks in the
Triassic Banks Wall Sandstone of the western Blue Mountain, New South
Wales. Journal of the Geological Society of Australia, 21, 393-402.
Loughnan, F. C., Ko Ko, M., and Bayliss, P., 1964. The red-beds of the Triassic
Narrabeen Group. Journal of the Geological Society of Australia, 11, 65-77.
Lundegard, P. D., 1992. Sandstone porosity loss - a "big picture" view of the
importance of compaction. Journal of Sedimentary Petrology, 62, 250-260.
Luo, J. L., Morad, S., Salem, A., Ketzer, J. M., Lei, X. L., Guo, D. Y., and Hlal, O.,
2009. Impact of diagenesis on reservoir-quality evolution in fluvial and
lacustrine-deltaic sandstones: evidence from Jurassic and Triassic
Sandstones from the Ordos Basin, China. Journal of Petroleum Geology,
32, 79-102.
Mack, G. H., 1978. The survivability of labile light-mineral grains in fluvial, aeolian
and littoral marine environments; the Permian Cutler and Cedar Mesa
Formations, Moab, Utah. Sedimentology, 25, 587-604.
Mackey, G. N., 2009. Provenance of the South Texas Paleocene-Eocene Wilcox
Group, western Gulf of Mexico basin: insights from sandstone modal
compositions and detrital zircon geochronology. M.Sc. thesis, University of
Texas at Austin,
Maher, H. D., Hays, T., Shuster, R., and Mutrux, J., 2004. Petrography of Lower
Cretaceous sandstones on Spitsbergen. Polar Research, 23, 147-165.
Majdi, A., Beiki, M., Pirayehgar, A., and Hosseinyar, G., 2010. Identification of well
logs with significant impact on prediction of oil and gas reservoirs
permeability using statistical analysis of RSE values. Journal of Petroleum
Science and Engineering, 75, 91-99.
379

Makowitz, A., and Milliken, K. L., 2003. Quantification of brittle deformation in burial
compaction, Frio and Mount Simon Formation sandstones. Journal of
Sedimentary Research, 73, 1007-1021.
Mansurbeg, H., Caja, M. A., Marfil, R., Morad, S., Remacha, E., Garcia, D., MartinCrespo, T., El-Ghali, M. A. K., and Nystuen, J. P., 2009. Diagenetic
evolution and porosity destruction of turbiditic hybrid arenites and siliciclastic
sandstones of foreland basins: evidence from the Eocene Hecho Group,
Pyrenees, Spain. Journal of Sedimentary Research, 79, 711-735.
Mansurbeg, H., Morad, S., Salem, A., Marfil, R., El-Ghali, M. A. K., Nystuen, J. P.,
Caja, M. A., Amorosi, A., Garcia, D., and La Iglesia, A., 2008. Diagenesis
and reservoir quality evolution of palaeocene deep-water, marine
sandstones, the Shetland-Faroes Basin, British continental shelf. Marine
and Petroleum Geology, 25, 514-543.
Maraschin, A. J., Mizusaki, A. M. P., and De Ros, L. F., 2004. Near-surface Kfeldspar precipitation in Cretaceous Sandstones from the Potiguar Basin,
northeastern Brazil. The Journal of Geology, 112, 317-334.
Marenssi, S. A., Net, L. I., and Santillana, S. N., 2002. Provenance, environmental
and paleogeographic controls on sandstone composition in an incised-valley
system; the Eocene La Meseta Formation, Seymour Island, Antarctica.
Sedimentary Geology, 150, 301-321.
Marfil, R., Rossi, C., Lozano, R. P., Permanyer, A., and Ramseyer, K., 2000.
Quartz cementation in Cretaceous and Jurassic reservoir sandstones from
the Salam oil field, Western Desert, Egypt: constraints on temperature and
timing of formation from fluid inclusions. Special Publication of the
International Association of Sedimentologists, 29, 163-182.
Marfil, R., Scherer, M., and Turrero, M. J., 1996. Diagenetic processes influencing
porosity in sandstones from the Triassic Buntsandstein of the Iberian
Range, Spain. Sedimentary Geology, 105, 203-219.
Mathisen, M. E., 1984. Diagenesis of Plio-Pleistocene nonmarine sandstones,
Cagayan Basin, Philippines; early development of secondary porosity in
volcanic sandstones. In: Surdam, R., Mcdonald, D. A. (Eds), Clastic
Diagenesis. Association of Petroleum Geologists Memoir, vol. 37. American
Association of Petroleum Geologists,Tulsa, Oklahoma, 177-193.
Matlack, K. S., Houseknecht, D. W., and Applin, K. R., 1989. Emplacement of clay
into sand by infiltration. Journal of Sedimentary Petrology, 59, 77-87.
Matsumoto, R., and Iijima, A., 1981. Origin and diagenetic evolution of Ca-Mg-Fe
carbonates in some coalfields of Japan. Sedimentology, 28, 239-259.
Mattson, A., and Chan, M. A., 2004. Facies and permeability relationships for
wave-modified and fluvial-dominated deposits of the Cretaceous Ferron
Sandstone, central Utah. AAPG Studies in Geology, 50, 251-275.
Mayne, S. L., Nicholas, E., Bigg-Wither, A. L., Rasidi, J. S., and Raine, M. J., 1974.
Geology of the Sydney Basin: a review. Bureau of Mineral Resources,
Geology and Geophysics, Australia Bulletin, 149.
McCaulay, G. E., Burley, S. D., Fallick, A. E., and Kusznir, N. J., 1994.
Palaehydrodynamic fluid flow regimes during diagenesis of the Brent Group
in the Hutton reservoir, constraints from oxygen isotope studies of
380

authigenic kaolin and reserves flexural modeling. Clay Minerals, 29, 609629.
McBride, E. F., Abdel-Wahab, A., and Salem, A. M. K., 1996. The influence of
diagenesis on the reservoir quality of Cambrian and Carboniferous
sandstones, Southwest Sinai, Egypt. Journal of African Earth Sciences, 22,
285-300.
McBride, E. F., Land, L. S., Diggs, T. N., and Mack, L. E., 1988. Petrography,
stable isotope geochemistry and diagenesis of Miocene sandstones,
Vermilion Block 31, offshore Louisiana. Gulf Coast Association of Geological
Society Transactions, 38, 513-523.
McDonnell, K. S., 1969. The Gosford Formation in the Terrigal - Bouddi district.
Abstract 4th Symposium on Advances in the Study of the Sydney Basin.
University of Newcastle.
McElroy, C. T., 1952. Evidence of the intrusive nature of the Berkeley latite,
Wollongong District, N.S.W. Australian Journal of Science, 15, 100.
McElroy, C. T., 1954. Petrology of sandstones of the southern Coalfield. M.Sc.
Thesis, University of Sydney (unpublished).
McElroy, C. T., 1957. Petrology of sandstone of the Southern Coalfield. Technical
Report Department of Mines. N.S.W, 2 (1954), 29-44.
McElroy, C. T., 1958. Occurrence of the Gosford Formation, Narrabeen Group in
the Western Coalfield. Technical Report Department of Mines. N.S.W, 3
(1955), 81.
McElroy, C. T., 1961. Notes on the field use of heavy mineral studies in the
Wollombi-Broke district. Technical Report Department of Mines. N.S.W, 6
(1958), 99-100.
McElroy, C. T., 1962. Explanatory notes, Sydney geological sheet. 2nd Ed. Bureau
Mineral Resources, Canberra.
McKibbin, D., and Smith, P. C., 2000. Sandstone hydrogeology of the Sydney
region. In: Mcnally, G. H., and Franklin, B. J. (Eds), Sandstone city;
Sydney's dimension stone and other sandstone geomaterials; proceedings
of a symposium held during the 15th Australian geological convention.
Geological Society of Australia Sydney, N.S.W., Australia, 5, 83-97.
McKinley, J. M., Atkinson, P. M., Lloyd, C. D., Ruffell, A. H., and Worden, R. H.,
2011. How porosity and permeability vary spatially with grain size, sorting,
cement volume, and mineral dissolution in fluvial Triassic sandstones: the
value of geostatistics and local regression. Journal of Sedimentary
Research, 81, 844-858.
McKinley, J. M., Worden, R. H., and Ruffell, A. H., 2001. Contact diagenesis: the
effect of an intrusion on reservoir quality in the Triassic Sherwood
Sandstone Group, Northern Ireland. Journal of Sedimentary Research, 71,
484-495.
Meisler, H., Leahy, P. P., and Knobel, L. L., 1984. Effect of eustatic sea-level
changes on saltwater-freshwater in the North Atlantic Coast Plain. United
States Geological Survey Water-Supply Paper 2255 27pp.
Miall, A. D., 1977. A review of the braided river depositional environments. Earth
Science Reviews, 13, 1-62.
381

Miall, A. D., 1978a. Lithofacies types and vertical profile models in braided river
deposits: a summary. Canadian Society of Petroleum Geologists Memoir, 5,
597-604.
Miall, A. D., (Ed), 1978. Fluvial Sedimentology. Canadian Society of Petroleum
Geologists Memoir, 5, 859 pp.
Miall, A. D., and Jones, B.G., 2003. Fluvial architecture of the Hawkesbury
Sandstone (Triassic), near Sydney, Australia. Journal of Sedimentary
Research, 73, 531-545.
Miao, J., Zhu, Z., Liu, W., and Lu, H., 2000. The characteristics of secondary
porosity in Eogene sandstone in the Jiyang Depression, Shandong, China,
and the factors controlling it. Xibei Daxue Xuebao = Journal of Northwest
University, 30, 154-158.
Michaelsen, P., and Henderson, R. A., 2000. Sandstone petrofacies expressions of
multiphase basinal tectonics and arc magmatism: Permian-Triassic north
Bowen Basin, Australia. Sedimentary Geology, 136, 113-136.
Midtbo, R. E. A., Rykkje, J. M., and Ramm, M., 2000. Deep burial diagenesis and
reservoir quality along the eastern flank of the Viking Graben. evidence for
illitization and quartz cementation after hydrocarbon emplacement. Clay
Minerals, 35, 227-237.
Milliken, K. L., 1998. Carbonate diagenesis in non-marine foreland sandstones at
the western edge of the Alleghanian overthrust belt, Southern Appalachians.
In: Morad, S. (Eds), Carbonate cementation in sandstones. International
Association of Sedimentologists Special Publication, 26, 87-105.
Milliken, K. L., 2001. Diagenetic heterogeneity in sandstone at the outcrop scale,
Breathitt Formation (Pennsylvanian), eastern Kentucky. AAPG Bulletin, 85,
795-815.
Milliken, K. L., 2003. Microscale distribution of kaolinite in Breathitt Formation
sandstones (Middle Pennsylvanian); implications for mass balance. Special
Publication of the International Association of Sedimentologists, 34, 343360.
Milliken, K. L., 2004. Late diagenesis and mass transfer in sandstone-shale
sequences. In: Mackenzie, F. T. (Eds), Sediments, Diagenesis, and
Sedimentary Socks. Elsevier Pergamon Oxford, United Kingdom, 7, 159190.
Mills, K. J., Moelle, K., and Branagan, D., 1989. Faulting near Mooney Mooney
Bridge, NSW. Abstracts of 23rd Symposium on Advances in the Study of the
Sydney Basin, University of Newcastle, Department of Geology, 217-224.
Moffitt, R. S., 2000. A compilation of the geology of the Southern Coalfield. Notes
to accompany the 1:100 000 Southern Coalfield Geology Map. Geological
Survey of New South Wales, Sydney, GS1998/277. 92pp.
Mohaghegh, S., Balan, B., and Ameri, S., 1997. Permeability determination from
well log data. SPE Formation Evaluation, September 1997, 170-174.
Molenaar, N., Cyziene, J., and Sliaupa, S., 2007. Quartz cementation mechanisms
and porosity variation in Baltic Cambrian sandstones. Sedimentary Geology,
195, 135-159.
382

Morad, S., 1986. Albitization of K-feldspar grains in Proterozoic arkoses and


greywackes from southern Sweden. Neues Jahrbuch fuer Mineralogie,
Monatshefte 1986, 145-186.
Morad, S., 1990. Mica alteration reactions in Jurassic reservoir sandstones from
the Haltenbanken area, offshore Norway. Clays and Clay Minerals, 38, 584590.
Morad, S., 1998. Carbonate cementation in sandstone: distribution patterns and
geochemical evolution. International Association Sedimentologist, Special
Publication, 26, 1-26.
Morad, S., Al-Ramadan, K., Ketzer, J. M., and De Ros, L. F., 2010. The impact of
diagenesis on the heterogeneity of sandstone reservoirs: A review of the
role of depositional facies and sequence stratigraphy. AAPG Bulletin, 94,
1267-1309.
Morad, S., Ben Ismail, H. N., De Ros, L. F., Al-Aasm, I. S. & Serrhini, N. E.,
1994.Diagenesis and formation water chemistry of Triassic reservoir
sandstones from southern Tunisia. Sedimentology, 41, 1253-1272.
Morad, S., Ketzer, J. M., and De Ros, L. F., 2000. Spatial and temporal distribution
of diagenetic alterations in siliciclastic rocks: implications for mass transfer
in sedimentary basins. Sedimentology, 47, 95-120.
Moraes, M. A. S., and De Ros, L. F., 1990. Infiltrated clays in fluvial Jurassic
sandstones of Reconcavo Basin, northeastern Brazil. Journal of
Sedimentary Petrology, 60, 809-819.
Moraes, M. A. S., and De Ros, L. F., 1992. Depositional, infiltrated and authigenic
clays in fluvial sandstones of the Jurassic Sergi Formation, Reconcavo
Basin, northeastern Brazil. Origin, Diagenesis, and Petrophysics of Clay
Minerals in Sandstones. In: Houseknecht David, W., Pittman Edward, D.
(Eds), Society for Sedimentary Geology. 47. SEPM, Tulsa, OK, United
States, 197-208.
Mork, M. B. E., and Moen, K., 2007. Compaction microstructures in quartz grains
and quartz cement in deeply buried reservoir sandstones using combined
petrography and EBSD analysis. Journal of Structural Geology, 29, 18431854.
Morton, A. C., 1991. Geochemical studies of detrital heavy minerals and their
application to provenance research. In: Morton, A. C., Todd, S. P., and
Haughton, P. D. W. (Eds), Developments in Sedimentary Provenance
Studies. Geological Society of London London, United Kingdom, 57, 31-45.
Morton, A. C., and Hallsworth, C., 1999. Processes controlling the composition of
heavy mneral assemblages in sandstones. Sedimentary Geology, 124, 330.
Morton, A. C., Whitham, A. G., and Fanning, C. M., 2005. Provenance of Late
Cretaceous to Paleocene submarine fan sandstones in the Norwegian Sea:
integration of heavy mineral, mineral chemical and zircon age data.
Sedimentary Geology, 182, 3-28.
Mozley, P. S., 1989. Relation between depositional environment and the elemental
composition of early diagenetic siderite. Geology 17, 704-706, doi:
10.1130/0091-7613(1989)017<0704: RBDEAT>2.3.CO; 2.
383

Mullard, B., 1995. New South Wales Petroleum Potential. Coal and Petroleum
Geology Branch. Geological Survey of New South Wales, Department of
Mineral Resources.
Nadeau, P. H., 1998. An experimental study of the effects of diagenetic clay
minerals on reservoir sands. Clays and Clay Minerals, 46, 18-26.
Nadeau, P. H., and Hurst, A., 1991. Application of back-scattered electron
microscopy to the quantification of clay mineral microporosity in sandstones.
Journal of Sedimentary Research, 61, 921-925.
Nascimento, M. D. S., Goes, A. M., Macambira, M. J. B., and Brod, J. A.,
2007.Provenance of Albian sandstones in the Sao Luis-Grajau Basin
(northern Brazil) from evidence of Pb-Pb zircon ages, mineral chemistry of
tourmaline and palaeocurrent data. Sedimentary Geology, 201, 21-42.
Nichols, G., 2009. Sedimentology and stratigraphy. Wiley-Blackwell, 419.
Noda, A., Takeuchi, M., and Adachi, M., 2004. Provenance of the Murihiku
Terrane, New Zealand: evidence from the Jurassic conglomerates and
sandstones in Southland. Sedimentary Geology, 164, 203-222.
Ochoa, M., Arribas, J., Mas, R., and Goldstein, R. H., 2007. Destruction of a fluvial
reservoir by hydrothermal activity (Cameros Basin, Spain). Sedimentary
Geology, 202, 158-173.
Odigi, M. I., and Amajor, L. C., 2010. Geochemistry of carbonate cements in
Cretaceous sandstones, southeast Benue Trough, Nigeria: implications for
geochemical evolution of formation waters. Journal of African Earth
Sciences, 57, 213-226.
Oelkers, E. H., Bjorkum, P. A., and Murphy, W. M., 1996. A petrographic and
computational investigation of quartz cementation and porosity reduction in
North Sea sandstones. American Journal of Science, 296, 420-452.
Ord, A., Vardoulakis, I., and Kajewski, R., 1991. Shear band formation in Gosford
Sandstone. International Journal of Rock Mechanics and Mining Sciences &
Geomechanics Abstracts, 28, 397-409.
Osae, S., Asiedu, D. K., Banoeng-Yakubo, B., Koeberl, C., and Dampare, S. B.,
2006. Provenance and tectonic setting of Late Proterozoic Buem
sandstones of southeastern Ghana: Evidence from geochemistry and
detrital modes. Journal of African Earth Sciences, 44, 85-96.
Osborne, G. D., 1948. A review of some aspects of the stratigraphy, structure and
physiography of the Sydney Basin. Proceedings Linnean Society of N.S.W.,
73, 4-37.
Packham, G. H., 1976. Narrabeen Palm Beach Triassic sedimentation.
International Geological Congress, 25th, Sydney - Excursion Guide 16B, 15
pp.
Parry, W. T., Chan, M. A., and Nash, B. P., 2009. Diagenetic characteristics of the
Jurassic Navajo Sandstone in the Covenant oil field, central Utah thrust belt.
AAPG Bulletin, 93, 1039-1061.
Paxton, S. T., Szabo, J. O., Ajdukiewicz, J. M., and Klimentidis, R. E., 2002.
Construction of an intergranular volume compaction curve for evaluating
and predicting compaction and porosity loss in rigid-grain sandstone
reservoirs. AAPG Bulletin, 86, 2047-2067.
384

Pay, M. D., Astin, T. R., and Parker, A., 2000. Clay mineral distribution in the
Devonian-Carboniferous sandstones of the Clair Field, west of Shetland,
and its significance for reservoir quality. Clay Minerals, 35, 151-162.
Pells, P. J. N., 2002. Development in the design of Tunnles and Caverns in the
Triassic Rocks of the Sydney Basin. International Journal of Rock
Mechanics and Mining Sciences, 39, 569-587.
Perri, F., Cirrincione, R., Critelli, S., Mazzoleni, P., and Pappalardo, A. A., 2008.
Clay mineral assemblages and sandstone compositions of the Mesozoic
Longobucco Group, northeastern Calabria: implications for burial history and
diagenetic evolution. International Geology Review, 50, 1116-1131.
Pettijohn, F. J., Potter, P. E., and Siever, R., 1987. Sand and Sandstone, 2nd ed.
Springer, New York. 553 pp.
Pittman, E. D., and Larese, R. E., 1991. Compaction of lithic sands; experimental
results and applications. AAPG Bulletin, 75, 1279-1299.
Pittman, E. D., Larese, R. E., and Heald, M. T., 1992. Clay coats: occurrence and
relevance to preservation of porosity in sandstones. In: Houseknecht, D. W.,
and Pittman, E. D. (Eds), Origin, Diagenesis, and Petrophysics of Clay
Minerals in Sandstones. SEPM Special Publication, 47, 241-264.
Primmer, T. J., Cade, C. A., Evans, J., Gluyas, J. G., Hopkins, M. S., Oxitoby, N.
H., Smalley, P. C., Warren, E. A., and Worden, R. H., 1997. Global patterns
in sandstone diagenesis: their application to reservoir quality prediction for
petroleum exploration. AAPG Menoir, 69, 61-77.
Pritchard, S., Hehir, W., Russell, G., 2004. A review of the status of the
groundwater resources in the Southern Highlands, NSW: Ensuring the
sustainability of the water source Department of Infrastructure, Planning and
Natural Resources, Science Unit, Wollongong.
Purvis, K., 1995. Diagenesis of Lower Jurassic sandstones, Block 211/13 (Penguin
Area), UK northern North Sea. Marine and Petroleum Geology, 12, 219-228.
Raggatt, H. G., 1938. Evolution of the Permo-Triassic basin of east central N.S.W.
D.Sc. thesis, University of Sydney (unpublished).
Rahmani, R. A., Steel, R. J., and Duaiji, A. A., 2003. Concepts and methods of
high-resolution sequence stratigraphy: applications to the Jauf gas reservoir,
Greater Ghawar, Saudi Arabia. GeoFrontier, 1, 15-21.
Ramm, M., 1992. Porosity-depth trends in reservoir sandstones; theoretical models
related to Jurassic sandstones offshore Norway. Marine and Petroleum
Geology, 9, 553-567.
Ramm, M., and Ryseth, A. E., 1996. Reservoir quality and burial diagenesis in the
Statfjord Formation, North Sea. Petroleum Geoscience, 2, 313-324.
Ramm, M., Forsberg, A. W., and Jahren, J. J., 1997. Porosity-depth trends in
deeply buried Upper Jurassic reservoirs in the Norwegian Central Graben;
an example of porosity preservation beneath the normal economic
basement by grain-coating microquartz. AAPG Memoir, 69, 177-199.
Ramm, M., 2000. Reservoir quality and its relationship to facies and provenance in
Middle to Upper Jurassic sequences, northeastern North Sea. Clay
Minerals, 35, 77-94.
385

Ramm, M., and Bjorlykke, K., 1994. Porosity/depth trends in reservoir sandstones;
assessing the quantitative effects of varying pore-pressure, temperature
history and mineralogy, Norwegian Shelf data. Clay Minerals, 29, 475-490.
Reed, J. S., Eriksson, K. A., and Kowalewski, M., 2005. Climatic, depositional and
burial controls on diagenesis of Appalachian Carboniferous sandstones:
qualitative and quantitative methods. Sedimentary Geology, 176, 225-246.
Retallack, G. J., 1977. Triassic Palaeosols in the Upper Narrabeen Group of New
South Wales, Part 1 Fractures of the palaeosols. Journal of the Geological
Society of Australia, 23, 383-400.
Reynolds, S. A., 1988. Depositional development and fluvial architecture of the
Narrabeen Group, Illawarra district, Sydney basin, Australia. MA thesis,
University of Texas at Austin (unpublished).
Rezaee, M. R., and Tingate, P. R., 1997. Origin of quartz cement in the Tirrawarra
Sandstone, southern Cooper Basin, South Australia. Journal of Sedimentary
Research, 67, 168-177.
Rieser, A. B., Neubauer, F., Liu, Y., and Ge, X., 2005. Sandstone provenance of
northwestern sectors of the intracontinental Cenozoic Qaidam Basin,
western China: tectonic vs. climatic control. Sedimentary Geology, 177, 118.
Roche, I., 1997. The Sydney Subgroup in the Southern Sydney Basin. B.Sc.
(Hons) thesis, University of Wollongong, Wollongong.
Roche, I., and Hutton, A., 1998. Distribution of the Sydney Subgroup, southwestern
margin of the Sydney Basin. In: Boyd, R. L., and Winwood-Smith, J. (Eds),
Proceedings of the Thirty-second Newcastle Symposium on Advances in
the Study of the Sydney Basin. University of Newcastle, N.S.W.,
Department of Geology Newcastle, N.S.W., Australia, 32, 119-126.
Rossi, C., Marfil, R., Ramseyer, K., and Permanyer, A., 2001. Facies-related
diagenesis and multiphase siderite cementation and dissolution in the
reservoir sandstones of the Khatatba Formation, Egypt's Western Desert.
Journal of Sedimentary Research, 71, 459-472.
Rossi, C., Kalin, O., Arribas, J., and Tortosa, A., 2002. Diagenesis, provenance
and reservoir quality of Triassic TAGI sandstones from Ourhoud field,
Berkine (Ghadames) Basin, Algeria. Marine and Petroleum Geology, 19,
117-142.
Runkel, A. C., Tipping, R. G., Alexander, E. C., and Alexander, S. C., 2006.
Hydrostratigraphic characterization of intergranular and secondary porosity
in part of the Cambrian sandstone aquifer system of the cratonic interior of
North America: improving predictability of hydrogeologic properties.
Sedimentary Geology, 184, 281-304.
Rust, B. R., and Jones, B. G., 1987. The Hawkesbury Sandstone south of Sydney,
Australia: Triassic analogue for the deposit of a large braided river. Journal
of Sedimentary Petrology, 57, 222-233.
Ryan, K. J., 2005. The geological controls of gas in the Narrabeen Group,
Southern Coalfield, NSW. B.Sc. (Hons) thesis, University of Wollongong,
Wollongong.
386

Ryu, I.-C., 2003. Petrography, diagenesis and provenance of Eocene Tyee Basin
sandstones, southern Oregon Coast Range: new view from sequence
stratigraphy. The Island Arc, 12, 398-410.
Sager, M. L., 2007. Petrologic study of the Murrysville Sandstone in southwestern
Pennsylvania. M.Sc. thesis, West Virginia University.
Saikia, C., Ahmad, A. H. H., and Wasim, S. M., 2011. Facies controlled diagenetic
evolution of the Delhi Group sandstones, Bayana Basin, Rajasthan. Journal
of Geochemical Society of India, 77, 261-268.
Salem, A. M., Ketzer, J. M., Morad, S., Rizk, R. R., and Al-Aasm, I. S., 2005.
Diagenesis and reservoir-quality evolution of incised-valley sandstones;
evidence from the Abu Madi gas reservoirs (upper Miocene), the Nile Delta
basin, Egypt. Journal of Sedimentary Research, 75, 572-584.
Salem, A. M., Morad, S., Mato, L. F., and Al-Aasm, I. S., 2000. Diagenesis and
reservoir-quality evolution of fluvial sandstones during progressive burial
and uplift: evidence from the Upper Jurassic Boipeba Member, Reconcavo
Basin, northeastern Brazil. AAPG Bulletin, 84, 1015-1040.
Sams, M. S., and Martijn, A., 2001. The effect of clay distribution on the elastic
properties of sandstone. Geophysical Prospecting, 49, 128-150.
Schmd, S., Worden, R. H., and Fisher, Q. J., 2003. The origin and regional
distribution of dolomite cement in sandstones from a Triassic dry river
system, Corrib Field, offshore west of Ireland. Journal of Geochemical
Exploration, 78-79, 475-479.
Schmid, S., Worden, R. H., and Fisher, Q. J., 2004. Diagenesis and reservoir
quality of the Sherwood Sandstone (Triassic), Corrib Field, Slyne Basin,
west of Ireland. Marine and Petroleum Geology, 21, 299-315.
Schoen, J. H., 1996. Physical properties of rocks: fundamentals and principles of
petrophysics, Handbook of Geophysical Exploration Seismic Exploration.
18. Elsevier Science, Oxford, UK.
Shah, B. A., and Bandyopadhyay, D. N., 2005. Feldspar alteration and diagenetic
characteristics of the Parsora sandstones, Son Basin, India. Gondwana
Research, 8, 258-265.
Shanmugam, G., 1988. Porosity prediction in sandstones using erosional
unconformities. In: Meshri, I. D., and Ortoleva, P.J. (Eds), Prediction of
Reservoir Quality through Chemical Modeling. AAPG Memoir, 49, 1-23.
Shibaoka, M., and Smyth, M., 1975. Coal petrology and the formation of coal
seams in some Australian sedimentary basins. Economic Geology 70, 14631473.
Siebert, R. M., Moncure, G. K., and Lahann, R. W., 1984. A theory of framework
grain dissolution in sandstones. In: Surdam, R., Mcdonald, D. A. (Eds),
Clastic Diagenesis. Association of Petroleum Geologists Memoir, vol. 37.
American Association of Petroleum Geologists,Tulsa, Oklahoma, 136-175.
Sims, C., 1996. The geology of the Sydney Subgroup Robertson NSW. B.Sc
(Hons) thesis, University of Wollongong, Wollongong.
Slatt, R. M., Jordan, D. W., Dagostino, A., and Gillespie, R. H., 1992. Outcrop
gamma-ray logging to improve understanding of subsurface well log
correlations. In: Hurst, A., Griffiths, C. M., Worthington, P. F. (Eds.),
387

Geological applications of wireline logs II. Geological Society Special


Publications, 65, 3-19.
Slatt, R. M., Stone, C. G., and Weimer, P., 2000. Characterization of slope and
basin facies tracts, Jackfork Group, Arkansas, with applications to
deepwater (Turbidite) reservoir management. GCSSEPM Foundation 20th
annual research conference, deep-water reservoirs of the world, December
3-6, 2000. 940-980.
Smosna, R., and Sager, M. L., 2008. The making of a high-porosity, highpermeability reservoir-the Murrysville Sandstone of Pennsylvania.
Northeastern Geology and Environmental Sciences, 30, 87-100.
Sonel, N., Demirel, I. H., Sari, A., and Buyukutku, A. G., 2009. Petrography,
provenance, and reservoir characteristics of the Upper Cretaceous to
Middle-Upper Eocene sandstones in the Eregli-Ulukisla Basin, Turkey.
Petroleum Science and Technology, 27, 543-556.
Song, F., Hou, J.-G., and Su, N.-N., 2009. Model building for Chang-8 low
permeability sandstone reservoir in the Yanchang formation of the Xifeng oil
field. Mining Science and Technology (China), 19, 245-251.
Spry, A. H., 2000. The Hawkesbury Sandstone: its origins and later life. In
Sandstone City: Sydney's dimension stone and other sandstone
geomaterials: proceedings of a symposium held on 7th July 2000, during the
15th Australian Geological Convention at the University of Technology
Sydney (G. H. McNally and B. J. Franklin, eds). Conference publications,
Springwood, N. S. W.
Standard, J. C., 1961. A new study of the Hawkesbury Sandstone; preliminary
findings. Journal and Proceedings of the Royal Society of New South
Wales, 95, 145-146.
Standard, J. C., 1964. Stratigraphy, structure and petrology of the Hawkesbury
Sandstone. PhD thesis, University of Sydney, Sydney.
Standard, J. C., 1969. Hawkesbury Sandstone. In: Packham, G. H. (Eds), The
Geology of New South Wales. Journal of the Geological Society of Australia,
16, 407-417.
Stonecipher, S. A., 2000. Applied sandstone diagenesis: practical petrographic
solutions for a variety of common exploration, development and production
problems. SEPM Short Course 50.
Storvoll, V., Bjorlykke, K., Karlsen, D., and Saigal, G., 2002. Porosity preservation
in reservoir sandstones due to grain-coating illite: a study of the Jurassic
Garn Formation from the Kristin and Lavrans fields, offshore Mid-Norway.
Marine and Petroleum Geology, 19, 767-781.
Stow, D. A. V., 2005. Sedimentary Rocks in the Field A Colour Guide. Elsevier
Academic Press, 320.
Sun, S., Shu, L., Zeng, Y., Cao, J., and Feng, Z., 2007. Porosity-permeability and
textural heterogeneity of reservoir sandstones from the Lower Cretaceous
Putaohua Member of Yaojia Formation, Weixing Oilfield, Songliao Basin,
northeast China. Marine and Petroleum Geology, 24, 109-127.

388

Sur, K. H., Lee, Y. I., and Hisada, K.-I., 2002. Diagenesis of the Lower Cretaceous
Kanmon Group sandstones, SW Japan. Journal of Asian Earth Sciences,
20, 921-935.
Surdam, R. C., Boese, S. W., and Crossey, L. J., 1984. The chemistry of
secondary porosity. In: Mcdonald, D. A., and Surdam, R. C. (Eds), Clastic
diagenesis. AAPG Memoir, 37, 127-134.
Suzuki, S., Takemura, S., Yumul, G. P., David, S. D., and Asiedu, D. K., 2000.
Composition and provenance of the Upper Cretaceous to Eocene
sandstones in central Palawan, Philippines; constraints on the tectonic
development of Palawan. The Island Arc, 9, 611-626.
Tamar-Agha, M. Y., 2009. The influence of cementation on the reservoir quality of
the Risha Sandstone Member (Upper Ordovician), Risha Gasfield, NE
Jordan. Journal of Petroleum Geology, 32, 193-208.
Tamrakar, N. K., Yokota, S., and Shrestha, S. D., 2007. Relationships among
mechanical, physical and petrographic properties of Siwalik sandstones,
central Nepal Sub-Himalayas. Engineering Geology, 90, 105-123.
Taylor, K. G., Gawthorpe, R. L., Curtis, C. D., Marshall, J. D., and Awwiller, D. N.,
2000. Carbonate cementation in a sequence-stratigraphic framework: Upper
Cretaceous sandstones, Book Cliffs, Utah-Colorado. Journal of Sedimentary
Research, 70, 360-372.
Taylor, K. G., Gawthorpe, R. L., and Van Wagoner, J. C., 1995. Stratigraphic
control on laterally persistent cementation, Book Cliffs, Utah. Journal of
Sedimentary Research, 69, 225-228.
Taylor, T. R., 1990. The influence of calcite dissolution on reservoir porosity in
Miocene sandstones, Picaroon Field, offshore Texas Gulf Coast. Journal of
Sedimentary Research, 60, 322-334.
Taylor, T. R., Stancliffe, R., Macaulay, C. I., and Hathon, L. A., 2004. High
temperature quartz cementation and the timing of hydrocarbon
accumulation in the Jurassic Norphlet Sandstone offshore Gulf of Mexico,
USA. In: Cubitt, J. M., England, W. A., and Larter, S. (Eds), Understanding
Petroleum Reservoirs: Towards an Integrated Reservoir Engineering and
Geochemical Approach. Geological Society (London) Special Publications,
237, 257-278.
Taylor, T. R., Giles, M. R., Hathon, L. A., Diggs, T. N., Braunsdorf, N. R., Birbiglia,
G. V., Kittridge, M. G., Macaulay, C. I., and Espejo, I. S., 2010. Sandstone
diagenesis and reservoir quality prediction: models, myths, and reality.
AAPG Bulletin, 94, 1093-1132.
Taylor, T. R., and Land, L. S., 1996. Association of allochthonous waters and
reservoir enhancement in deeply buried Miocene sandstones: Picaroon
Field, Corsair Trend, offshore Texas. In: Crossey, L. J., Loucks, R., and
Totten, M. W. (Eds), Siliciclastic Diagenesis and Fluid Flow: Concepts and
Applications. SEPM Special Publication, 55, 37-48.
Thomas, A., Balcer, D., Himes, T., Bonnell, L., Jones, J., and O'Mahoney, L., 2004.
Deep Cotton Valley Formation diagenesis and reservoir quality, Viosca Knoll
area, offshore Gulf of Mexico. In, American Association Petroleum
Geologists
Abstract
Volume,
Hedberg
research
conference
389

Austin,Texas,http://www.searchanddiscovery.net/documents/abstracts/hedb
erg2004austin/index.htm.
Thomson, A., 1979. Preservation of porosity in the deep Woodbine/Tuscaloosa
trend, Louisiana. Gulf Coast Association of Geological Societies
Transactions, 30, 396-403.
Tijani, M. N., Nton, M. E., and Kitagawa, R., 2010. Textural and geochemical
characteristics of the Ajali Sandstone, Anambra Basin, SE Nigeria:
implication for its provenance. Comptes Rendus Geosciences, 342, 136150.
Tobin, R. C., Mcclain, T., Lieber, R.B., Ozkan, A., Banfield, L. A., Marchand, A. M.
E., and Mcrae, L. E., 2010. Reservoir quality modeling of tight-gas sands in
Wamsutter field: Integration of diagenesis, petroleum systems, and
production data. AAPG Bulletin, 94, 1229-1266.
Tucker, M. E., 1991. Sedimentary Petrology an Introduction to the Origin of
Sedimentary Rocks. Blackwell Scientific Publications, Oxford, Boston.
Tucker, M. E., 1993. Carbonate diagenesis and sequence stratigraphy.
Sedimentology Review, 1, 51-72.
Umar, M., Friis, H., Khan, A. S., Kassi, A. M., and Kasi, A. K., 2011. The effects of
diagenesis on the reservoir characters in sandstones of the Late Cretaceous
Pab Formation, Kirthar Fold Belt, southern Pakistan. Journal of Asian Earth
Sciences, 40, 622-635.
van De Kamp, P. C., 2010. Arkose, subarkose, quartz sand, and associated muds
derived from felsic plutonic rocks in glacial to tropical humid climates.
Journal of Sedimentary Research, 56, 329-345.
Van Den Bril, K., and Swennen, R., 2008. Sedimentological control on carbonate
cementation in the Luxembourg Sandstone Formation. Geologica Belgica,
12, 3-23.
Van Wagoner, J. C., Mitchum, R.M., Campion, K.M., and Rahmanian, V.D., 1990.
Siliciclastic Sequence Stratigraphy in Well Logs, Cores, and Outcrops.
AAPG, Tulsa, 55 p.
Veevers, J. J., (Ed.), 1984. Phanerozoic Earth History of Australia. Claredon Press.
418 pp.
Veizer, J., 1983. Chemical diagenesis of carbonate: theory and application of trace
element technique. In: Arthur, M. A. (Eds), Stable isotopes in sedimentary
geology. Society of Economic Paleontologists and Mineralogists Short
Course Notes, 10, 3-100.
Vogler, H. A., and Robison, B. A., 1987. Exploration for deep geopressured gas:
Corsair trend, offshore Texas. AAPG Bulletin, 71, 777-787.
von Eynatten, H., and Gaupp, R., 1999. Provenance of Cretaceous synorogenic
sandstones in the Eastern Alps: constraints from framework petrography,
heavy mineral analysis and mineral chemistry. Sedimentary Geology, 124,
81-111.
Vosylius, G., 1998. Reservoir properties of Middle Cambrian rocks. In:
Zdanaviciute, O., Suveizdis, P. (Eds.), Proceedings of the International
Scientific Conference "Prospectives of petroleum exploration in the Baltic
Region". Lithuanian Geological Institute, pp. 43-48.
390

Walderhaug, O., 2000. Modeling quartz cementation and porosity in Middle


Jurassic Brent Group sandstones of the Kvitebjorn field, northern North Sea.
AAPG Bulletin, 84, 1325-1339.
Wanas, H. A., 2008. Calcite-cemented concretions in shallow marine and fluvial
sandstones of the Birket Qarun Formation (Late Eocene), El-Faiyum
depression, Egypt: Field, petrographic and geochemical studies:
implications for formation conditions. Sedimentary Geology, 212, 40-48.
Wanas, H. A., and Abdel-Maguid, N. M., 2006. Petrography and geochemistry of
the Cambro-Ordovician Wajid Sandstone, southwest Saudi Arabia:
Implications for provenance and tectonic setting. Journal of Asian Earth
Sciences, 27, 416-429.
Wang, H.-A., Zhong, J.-H., Zhong, F.-P., Niu, Y.-B., and Wang, P.-J., 2009.
Characteristics and mechanism of low permeability beach-bar sandstone
reservoir of Es4 in Dongying sag. Mining Science and Technology (China),
19, 788-795.
Wang, W. H., 1992. Origin of reddening and secondary porosity in Carboniferous
sandstones, Northern Ireland. In: Parnell, J. (Eds), Basins on the Atlantic
Seabord: Petroleum Geology, Sedimentology and Evolution. Geological
Society Special Publication, 62, 243-254.
Wani, W., and Mondal, M. E. A., 2009. Petrochemistry of sandstones from
Neoproterozoic basins of the Bastar craton, Central Indian Shield:
implications for paleoweathering, provenance and tectonic history. Island
Arc, 18, 352-374.
Ward, C. R., 1971a. Mineralogical changes as marker horizons for stratigraphic
correlation in the Narrabeen Group of the Sydney Basin, N.S.W. Journal
and Proceedings of the Royal Society of New South Wales, 104, 77-88.
Ward, C. R., 1971b. Mesozoic sedimentation and structure in the southern part of
the Sydney basin-the Narrabeen Group. PhD thesis, University of New
South Wales (unpublished), Sydney.
Ward, C. R., 1972. Sedimentation in the Narrabeen Group, southern Sydney
Basin, New South Wales. Journal of the Geological Society of Australia, 19,
393-409.
Ward, C. R., 1980. Notes on the Bulgo Sandstone and the Bald Hill Claystone.
Geological Survey of New South Wales Bulletin, 26, 179-186.
Weaver, C. E., 1989. Clays, Muds, and Shales. Developments in Sedimentology
44. Elsevier Science Publications, pp. 819.
Wilkinson, M., and Haszeldine, R. S., 1996. Aluminium loss during sandstone
diagenesis. Journal of the Geological Society London, 153, 657-660.
Wilkinson, M., Milliken, K. L., and Haszeldine, R. S., 2001. Systematic destruction
of K-feldspar in deeply buried rift and passive margin sandstones. Journal of
the Geological Society, London, 158, 675-683.
Wilkinson, M., Haszeldine, R. S., Ellam, R. M., and Fallick, A., 2004. Hydrocarbon
filling history from diagenetic evidence; Brent Group, UK North Sea. Marine
and Petroleum Geology, 21, 443-455.
Willner, A. P., Ermolaeva, T., Stroink, L., Glasmacher, U. A., Giese, U., Puchkov,
V. N., Kozlov, V. I., and Walter, R., 2001. Contrasting provenance signals in
391

Riphean and Vendian sandstones in the SW Urals (Russia); constraints for


a change from passive to active continental margin conditions in the
Neoproterozoic. Precambrian Research, 110, 215-239.
Wilson, M. D., and Stanton, P. T., 1994. Diagenetic mechanisms of porosity and
permeability reduction and enhancement In: Wilson, M. D. (Eds), Reservoir
Quality Assesment and Prediction in Clastic Rocks. SEPM Special Course,
30, 59-118.
Wolela, A., 2009. Diagenetic evolution of the Ansian-Pliensbachian Adigrat
Sandstone, Blue Nile Basin, Ethiopia. Journal of African Earth Sciences, 56,
29-42.
Wolela, A. M., and Gierlowski-Kordesch, E. H., 2007. Diagenetic history of fluvial
and lacustrine sandstones of the Hartford Basin (Triassic-Jurassic), Newark
Supergroup, USA. Sedimentary Geology, 197, 99-126.
Wong, P. M., 1999. Permeability prediction from well logs using an improved
windowing technique. Journal of Petroleum Geology, 22, 215-226.
Worden, R., and Morad, S., 2000. Quartz cement in oil field sandstones: a review
of the critical problems. In: Worden, R. H., and Morad, S. (Eds), Quartz
Cementation in Sandstones. International association of sedimentologists,
special publication, 29, 1-20.
Worden, R. H., Mayall, M., and Evans, I. J., 2000. The effect of ductile-lithic sand
grains and quartz cement on porosity and permeability in Oligocene and
lower Miocene clastics, South China Sea: prediction of reservoir quality.
AAPG Bulletin, 84, 345-359.
Worden, R. H., and Burley, S., 2003. Sandstone diagenesis: the evolution of sand
to stone. In: Burley, S., Worden, R. (Eds), Sandstone Diagenesis: Recent to
Ancient. International Association of Sedimentologists, Reprint Series, 4, 344.
Worden, R. H., and Morad, S., 2003. Clay minerals in sandstones; controls on
formation, distribution and evolution. In: Worden, R. H., and Morad, S.
(Eds), Clay Mineral Cements in Sandstones. International Association of
Sedimentologists, Special Publication, 34, 3-41.
Worthington, P. F., 2003. Effect of clay content upon some physical properties of
sandstone reservoirs, In: Worden, R. H., and Morad, S. (eds), Clay Mineral
Cements in Sandstones. pp.191-211. IAS Special Publication., 34. Blackwell
Science, Oxford.
Xiaomin, Z., Dakang, Z., Qin, Z., and Li, Z., 2004. Sandstone diagenesis and
porosity evolution of Paleogene in Huimin depression. Petroleum Science,
1, 23-29.
Xu, T., Wang, X., Zhang, Y., Zhao, X., and Bao, Y., 2003. Clay Minerals in
Sedimentary Basins of China. Petroleum Industry Press, Beijing, 65 pp (in
Chinese).
Yang, W., 2007. Transgressive wave ravinement on an epicontinental shelf as
recorded by an Upper Pennsylvanian soil-nodule conglomerate-sandstone
unit, Kansas and Oklahoma, U.S.A. Sedimentary Geology, 197, 189-205.
Yoo, E. K., 1988. The Rocky Glen Ridge and Gilgandra Trough, beneath the Surat
Basin. Geological Survey of New South Wales Quarterly Notes, 72, 17-27.
392

Yoshida, K., and Machiyama, H., 2004. Provenance of Permian sandstones, south
Kitakami Terrane, northeast Japan: implications for Permian arc evolution.
Sedimentary Geology, 166, 185-207.
Zamanzadeh, S. M., Amini, A., and Ghavidel-Syooki, M., 2009. Sequence
stratigraphic controls on early-diagenetic carbonate cementation of shallow
marine clastic sediments (the Devonian Zakeen Formation, southern
Zagros, Iran). Geosciences Journal, 13, 31-57.
Zhang, C. L., Horita, J., Cole, D. R., Zhou, J., Lovley, D. R., and Phelps, T. J.,
2001. Temperature-dependent oxygen and carbon isotope fractionations of
biogenic siderite. Geochimica et Cosmochimica Acta, 65, 2257-2271.
Zhang, J., Qin, L., and Zhang, Z., 2008. Depositional facies, diagenesis and their
impact on the reservoir quality of Silurian sandstones from Tazhong area in
central Tarim Basin, western China. Journal of Asian Earth Sciences, 33,
42-60.
Zhang, J.-L., Jia, Y., and Du, G., 2007. Diagenesis and its effect on reservoir
quality of Silurian sandstones, Tabei area, Tarim Basin, China. Petroleum
Science, 4, 1-13.
Zhang, J.-L., Li, D-Y., and Jiang, Z-Q., 2010. Diagenesis and reservoir quality of
the fourth member sandstones of Shahejie Formation in Huimin depression,
eastern China. J. Cent. South University of Technology, 17, 169-179.
Zhang, L., Bai, G., Luo, X., Ma, X., Chen, M., Wu, M., and Yang, W., 2009.
Diagenetic history of tight sandstones and gas entrapment in the Yulin Gas
Field in the central area of the Ordos Basin, China. Marine and Petroleum
Geology, 26, 974-989.
Zheng, H., Huang, X., and Butcher, K., 2006. Lithostratigraphy, petrography and
facies analysis of the late Cenozoic sediments in the foreland basin of the
west Kunlun. Palaeogeography Palaeoclimatology Palaeoecology, 241, 6178.
Zhigang, W., 2000. A study on the relationship between the diagenesis and
porosity of sandstone reservoir rocks of the Shahejie Formation in Bonan
Field, China. Shiyou Kantan yu Kaifa = Petroleum Exploration and
Development, 27, 37-38.
Zhiqian, G., Bingsong, Y., and Xingyun, L., 2007. The equilibrium between
diagenetic calcites and dolomites and its impact on reservoir quality in the
sandstone reservoir of Kela 2 gas field. Progress in Natural Science, 17,
1051-1058.
Zhou, H. R., Wang, X. L., Liu, Z.R., and Deng, H.W., 2006. Braided river delta
sediments of the Huangshanjie Formation of Upper Triassic in southern
Junggar Basin Journal of Palaeogeography, 8, 187-198 (in Chinese with
Emglish abstract).

393

Volume 2

Appendices

Definition of symbols
Fm

Formation

HBSS
NBG
NPFM
GRFM
BACS
BGSS
SPCS
SBSS
WBCS
CCSS
ICM
BUSM
LDSS
BASM
LRSS
CHSM
UNM2
HGSM
UNM3
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM

Hawkesbury Sandstone
Narrabeen Group
Newport Formation
Garie Formation
Bald Hill Claystone
Bulgo Sandstone
Stanwell Park Claystone
Scarborough Sandstone
Wombarra Claystone
Coalcliff Sandstone
Illawarra Coal Measures
Bulli Coal
Loddon Sandstone
Balgownie Coal Member
Lawrence Sandstone
Cape Horn Coal Member
Unnamed Member Two
Hargrave Coal Member
Unnamed Member Three
Wongawilli Coal
Kembla Sandstone
American Creek Coal Member
Allans Creek Formation
Darkes Forest Sandstone
Bargo Claystone
Tongarra Coal

WTFM

Wilton Formation

Appendix One

Appendix 1-1: Lithofacies codes (modified after Miall, 1977, 1978a, from Dehghani, 1994)
Facies code

Facies

Sedimentary structures

Interpretation

Gms

massive, matrix suported gravel

grading

debris flow deposits

Gm

horizontal bedding, imbrication

longitudinal bars, lag deposits, sieve deposits

crudely bedded

channel floor lag deposit

Gt

massive or crudely bedded gravel


thin layer of gravel at base of
channels
gravel, stratified

trough cross-bedded

Gp

gravel, stratified

planar cross-bedded

minor chanel fills


longitudinal bars, deltaic growths from older bar
remnants
formed due to break down of ironstone bands and
moved along channel floor

Gld

Sm

single line of ironstone clasts at


base of the migrating channels
massive intraformational
conglomerate
conglomerate filled small channels
sand, medium to very coarse, may
be pebbly
sand, medium to very coarse, may
be pebbly
sand, very fine to medium
sand, very fine to very coarse, may
be pebbly
sand, very fine to very coarse, may
be pebbly
inclined units with heterolithic
lithology
erosional scours with intraclasts
sand, fine to very coarse, may be
pebbly
fine to coarse-grained sandstone

massive

longitudinal sandy bars, upper flow regime

Fl

sand, silt, mud deposits

fine lamination, very small ripples

overbank or waning flood

Fsc

silt, mud

laminated to massive

backswamp deposit

Fcf

mud

massive, with freshwater molluscs

backswamp pond deposits

Fm

mud, silt

massive, with dessication cracks

Fib

ironstone bands

coal, carbonaceous mud

plant, mud films

swamp deposits

carbonate

pedogenic features

palaeosol

Gis
Gif
Ge
St
Sp
Sr
Sh
S1
S1h
Se
Ss

crudely bedded

accumulation of ironstone clasts

massive and matrix-supported

scour and film conglomerate

solitary or grouped trough cross-beds


solitary or grouped planar cross-beds
ripple cross-lamination
horizontal lamination, parting or streaming
lineation

3-D dunes (lower flow regime)


2-D dunes (linguid, transverse bars, sand waves),
(lower flow regime)
ripples (lower flow regime)
planar bed (upper flow regime)

low angle (<10) cross-beds

scour films, washed-out dunes, antidunes

low angle cross-bedding

laterally accreted bars

crude cross-bedding

scours fills

broad, shallow scours

scour fills

overbank or drape deposits


chemically precipitated siderite

Appendix 1-2: Facies system for the Sydney Subgroup (Kennedy, 1999).
Kennedy 1999

Gm

Gh

Roche
1997

Diessel
1992

Gm

Gh

Miall 1996

Lithofacies

Sedimentary structures

Interpretation

Gms

matrix supported massive gravel

weak grading

plastic debris flow (high-strength viscous)

Gmg

matrix supported gravel, debris

Inverse to normal grading

pseudoplastic flow (high-strength viscous)

Gcl

clast supported gravel

inverse grading

clast rich debris flow (high strength) or pseudoplastic flow (strength viscous)

Gcm

clast supported massive gravel

inverse grading

clast rich debris flow (high strength) or pseudoplastic flow (strength viscous)

Gh

clast supported, crudely bedded gravel

horizontal bedding, imbrication

longitudinal bedforms lag deposits, sieve deposits

Gt

gravel stratified

trough cross-beds

minor channel fills

Gp

gravel stratified

planar cross-beds

transverse bedforms, deltaic growths from older bar remnants

Sm

Sm

Sm

sandstone, fine to medium grained

massive or faint lamination

sediment gravity flows

Sh

Sh

Sh

sandstone very fine to coarse grained

horizontal lamination, streaming lineation

plane bed flow (critical)

Sl

sandstone very fine to coarse grained

low angle (<150) cross-beds

scour fills, humpback or washed out dunes and antidunes

Ss

sandstone, fine to very coarse

broad, shallow scours

scour fill

St

sandstone fine to very coarse

solitary or grouped trough cross-beds

sinuous-crested and linguoid (3D) dunes

Sp

sandstone, fine to very coarse

solitary or grouped planar cross-beds

transverse and linguoid bedforms (2D) dunes

Sr

sandstone very fine to coarse

ripple cross lamination

ripples (lower flow regime)

Fm

mud, silt

massive, desiccation cracks

overbank, abandoned channel or drape deposits

Fsm

mud, silt

massive

backswamp or abandoned channel deposits

Fr

mud, silt

massive, roots, bioturbation

root bed, incipient soil

Fl

sand, silt, mud

fine lamination, very small ripples

overbank, abandoned channel or waning flood deposits

Sx

Fm

Sx

Fm

Fr
Fl

Fl

Fx

Fx

fine sediments

cross bedded

levee and overbank deposits (proximal)

Ft

Ft

tuffaceous lithologies

any structures

ash fall deposition into clam environments

carbonate

veins

time of dehydration or soil deposit

P
U

P
U

undifferentiated

bright (greater than 90% bright)

horizontal lamination

overbank

Bd

Bd

bright with dull bands (over and up to and including 90% bright)

horizontal lamination

overbank

DB

DB

interbanded dull and bright (over 40% and up to 60% bright)

horizontal lamination

overbank

Db

Db

mainly dull with frequent bright bands over 10% and up to including 40% bright

horizontal lamination

overbank

Dmb

Dmb

dull with minor bright bands (over 1% and up to and including 10% bright

horizontal lamination

overbank

dull (up to and including 1% bright)

horizontal lamination

overbank

Cs

shaly-coal (<50% shale)

horizontal lamination

overbank

Sc

coaly-shale (<50% coal)

horizontal lamination

overbank

CS

shaly-coal (50:50)

horizontal lamination

overbank

Mc

carbonaceous fine lithologies

any structures

any with a nearby carbonaceous source

Appendix 1-3: Sedimentary Environments of the Cumberland Subgroup (Bowman, 1980).


Stratigraphic unit
Erins Vale
Formation

Cumberland
Subgroup

Pheasants
Nest
Formation

Location of good
exposures

Lithology

Grain size

Sorting

Sedimentary
structures

Environment

Waniora Point
Flat Rock

Sandstone
siltstone

Fine to medium

Moderate

Flat bedding,
burrowing

Bay and lagoons

Goolgong Avenue
Cordeaux Road

Coal
claystone
siltstone

Fine

Flat thin bedding

Paludal

Bellambi Point
Towradgi Point

Sandstone

Medium

Moderate

Festoon cross-beds,
overturned trees

Feston cross bedded


zones of fluvial point
bar

OBriens Road

Sandstone

Coarse

Moderate

Poorly bedded

Fluvial channel base

Motorway north

Sandstone
shale

Fine to medium

Moderate

Thin flat bedding

Delta fringe and


littoral

Appendix 1-4: Environment of the Sydney Subgroup (Bowman, 1980).


Formation

Lithology

Color

Sedimentary structures

Environment

Bulli Coal

Coal

Sorting

Black

Flat bedded

Backswamp

Upper Eckersley
Formation claystone

Clay

Mid grey

Ripple cross- bedding

Floodplain

Upper Eckersley
Formation sandstone

Fine to medium
sandstone

Light grey

Cross-bedding

Point bar

Balgownie Coal Member

Coal

Black

Flat bedding

Backswamp

Middle Eckersley
Formation claystone

Clay

Mid grey

Interbedded claystone and sandstone

Floodplain

Lower Eckersley
Formation sandstone

Fine to medium
sandstone

Moderately well
sorted

Light grey

Cross-bedding

Point bar

Wongawilli Coal

Coal

Moderately

Black

Flat bedding

Backswamp

Moderately well
sorted

Kembla Sandstone

Fine sandstone

well sorted

Mid to light grey

Cross-bedding

Point bar

Allans Creek Formation

Shale and Coal

Moderately sorted

Dark grey

Flat bedding

Floodplain

Moderately well
sorted

Mid to light grey

Flat interbedded sandstones


and claystones

Point bar

Dark grey

Flat bedding

Floodplain

Mid to light grey

Flat bedding

Alternating bed facies of


fluvial point bar

Black

Flat bedding

Backswamp

Darkes Forest Sandstone


Bargo Claystone

Shale

Austinmer Sandstone
Member

Very fine to fine


sandstone

Tongarra Coal

Coal

Upper Wilton Formation

Laminate

Moderately sorted

Dark to light grey

Ripple cross - bedding, Laminated

Floodplain

Middle Wilton Formation

Very fine to fine


sandstone

Not Measured

Dark to light

Flat bedding

Point bar

Woonona Coal Member

Coal

Black

Flat bedding

Backswamp

Lower Wilton Formation

Clay to coarse
sandstone

Light grey

Tabular and feston cross -bedding

Point bar

Moderately well
sorted

Moderately well
sorted

Disconformity

Appendix Two

Appendix 2-1: a) Shale unit (428.32-428.62 m) and carbonaceous mudstone


(428.62-428.9 m) are present in the Wilton Formation, b) Grey colour of
medium-grained sandstone in the Wilton Formation (427.03-427.62 m), c)
Carbonaceous shale includes rare bright coal bands in the Tongarra Coal
(421.34-421.44 m), d) Coal units (422.12-422.36 m) occur in the Tongarra
Coal.

Appendix 2-2:, a) Shale unit in the Bargo Claystone (411.5-412.94 m), b)


Siltstone unit contains clayey over base (red arrow) in the Bargo Claystone
(414.7-415.22 m), c) Fine-grained sandstone in the Bargo Claystone
(410.27-410.84 m), d) Medium-grained sandstone in the Darkes Forest
Sandstone (404.7-405.65 m).

Appendix 2-3: a) Fine-grained sandstone includes thin silty lenses


throughout unit in the Darkes Forest Sandstone (403.51-404.7 m), b) Finegrained sandstone occurs in the Darkes Forest Sandstone and contains
muddy wisps (406.65-407.02 m), c) Siltstone unit is present in the Darkes
Forest Sandstone, containing sandy over base (405.65-406.01 m), d) Finegrained sandstone in the lower part of the Darkes Forest Sandstone and
includes silty bands and wisps (402.26-403.51 m).

Appendix 2-4: a) Siltstone unit in the Darkes Forest Sandstone


(397.79-398.06 m), b) Shale unit contains tuffaceous bed near the base
of the Darkes Forest Sandstone (398.93-399 m), c) Black carbonaceous
mudstone in the middle part of the Darkes Forest Sandstone includes
claystone bands over the base (400.84-401.11 m), d) Shale unit has
carbonaceous and band in the American Creek Coal Member (392.54393.01 m).

Appendix 2-5: a) Carbonaceous mudstone is characterised by rare bright


coal bands and clayey band in the American Creek Coal Member (396.66397 m), b) Fine-grained sandstone includes generally carbonaceous
laminae bands in the Kembla Sandstone (679.94-681.16 m), c) Finegrained sandstone is recorded in the upper part of the Kembla Sandstone,
including muddy laminae (678.095-678.23 m), d) Mudstone facies in the
upper part of the Kembla Sandstone (678.23-678.46 m).

Appendix 2-6: a) Carbonaceous shale exists in the Wongawilli Coal


(675.7-657.81 m), b) Black coal is observed in the Wongawilli Coal
(673.75-673.84 m), c) Tuff unit occurs in the Wongawilli Coal (669.49670.06 m), d) Greyish brown tuff in the Wongawilli Coal (669.43669.46 m).

10

Appendix 2-7: a) Shale unit in the Unnamed Member Three (373.65373.9 m), b) Shale is grey in the Unnamed Member Two (660.77661.9 m), c) Fine-grained sandstone in channel facies are present in
upper part of the Unnamed Two Member, containing phases of
mudstone (644.38-645.8 m), d) Fine-grained sandstone in the Unnamed
Two Member (646.23-646.8 m).

11

Appendix 2-8: a) Carbonaceous mudstone facies in the Unnamed Member


Two (654.13-655.18m ), b) Tuff is distinguished in the Unnamed Member
Two (656.79-656.88 m), c) Coal unit in the Cape Horn Coal Member
(643.12-643.245 m), d) Medium-grained sandstone is described in the
Lawrence Sandstone (636.62-639.02 m).

12

Appendix 2-9: a) Fine-grained sandstone is presented in the Lawrence


Sandstone, including carbonaceous laminae (635.86-636.62 m), b)
Medium-grained sandstone in the middle part of the Loddon Sandstone
(630.31-631.75 m), c) Medium-grained sandstone in the Loddon
Sandstone, including rare carbonaceous bands lenses (629.53-630.31 m),
d) Fine-grained sandstone is deposited in the Loddon Sandstone and is
dominated by carbonaceous laminae bands (628.67-629.53 m).

13

Appendix 2-10: a) Brownish black mudstone unit occurs in the upper


the Loddon Sandstone (627.69-628.49 m), b) Carbonaceous mudstone
at the top of the Loddon Sandstone contains coaly lenses (627.54627.67 m), c) Coal unit in the Bulli Coal (625.74-626.15 m), d) Coarsegrained sandstone contains rare pebbles (red arrow) in the Coalcliff
Sandstone (612.08-613.31 m).

14

Appendix 2-11: a) Medium-grained sandstone in the Coalcliff


Sandstone, including minor coaly wisps (597.04-598.60 m), b) Finegrained sandstone contains intraclasts of mudstone in the Coalcliff
Sandstone (592.8-595.55 m), c) Medium-grained sandstone is present in
the Coalcliff Sandstone and has mudstone clasts (615.51-616.27 m), d)
Fine-grained sandstone occurs in floodplain facies of the Coalcliff
Sandstone (599.58-600.10 m).

15

Appendix 2-12: a) Mudstone unit in the Coalcliff Sandstone (600.42600.77 m), b) Shale unit in the Wombarra Claystone (569.07-590.08 m),
c) Fine-grained sandstone in the Wombarra Claystone (557.28559.19 m), d) Coarse-grained sandstone in the Scarborough Sandstone
contains granules and pebbles (544.8-545.05 m).

16

Appendix 2-13: a) Medium-grained sandstone near the top of the


Scarborough Sandstone (514.31-520.68 m), b) Medium-grained
sandstone occurs in the Scarborough Sandstone (523.08-523.35 m),
c) Fine-grained sandstone at the base of the Scarborough Sandstone
(553.14-554.34 m), d) Fine-grained sandstone in the Scarborough
Sandstone (553.04-553.14 m).

17

Appendix 2-14: a) Siltstone unit in the Scarborough Sandstone


(526.01-526.28 m), b) Shale unit is silty in places in the
Scarborough Sandstone (529.92-531.45 m), c) Shale unit in the
Stanwell Park Claystone (504.3-505.39 m), d) Fine-grained
sandstone has minor mudstone band in the Stanwell Park
Claystone (509.14-513.04 m).

18

Appendix 2-15: a) Coarse-grained sandstone exists in the Bulgo


Sandstone (395.3-397.01 m), b) Medium-grained sandstone
contains clayey pebbles in the Bulgo Sandstone (401.11402.31 m), c) Fine-grained sandstone in the Bulgo Sandstone
(411.85-413.98 m), d) Coarse-grained sandstone in the Bulgo
Sandstone, including carbonaceous band (red arrow) near base.
(416.33-417.13 m).

19

Appendix 2-16: a) Medium-grained sandstone in the Bulgo


Sandstone (438.61-438.86 m), b) Fine-grained sandstone is
recorded in the Bulgo Sandstone (485.37-486.14 m), c) Shale
unit in the Bulgo Sandstone (392.03-392.65 m), d) Claystone in
the Bald Hill Claystone (268.40-272.29 m).

20

Appendix 2-17: a) Siltstone unit in the Bald Hill Claystone


(253.60-253.92 m), b) Claystone in the Garie Formation
(263.50-267.20 m), c) Fine-grained is characterised by silty
laminae in the Newport Formation (259.46-261.01 m), d) Finegrained sandstone includes carbonaceous laminae and
carbonaceous wisps in the Newport Formation (255.60256.31 m),

21

Appendix 2-18: a) Siltstone in the Newport Formation


(252.11-252.84 m), b) Mudstone unit at the top of the Newport
Formation (246.90-247.59 m), c) Conglomerate unit has sandy
matrix and pebbly band in the Hawkesbury Sandstone (178.61180.20 m), d) Coarse-grained sandstone includes coarse
granular bands in the Hawkesbury Sandstone (227.39230.38 m).

22

Appendix 2-19: a) Coarse-grained sandstone contains


phases of pebbles (red arrow) in the Hawkesbury Sandstone
(178.12-178.61 m), b) Medium-grained sandstone has
mudstone clasts (red arrow) in the Hawkesbury Sandstone
(185.50-194.46 m), c) Medium-grained sandstone in the
Hawkesbury Sandstone (218.39-219.37 m), d) Whitish grey
fine-grained sandstone in the Hawkesbury Sandstone
(41.168-42.02 m), e) Ripple bedded siltstone in the
Hawkesbury Sandstone (42.98-43.59 m).
23

Appendix Three

EDEN 127

EAW 18a

EDEN 115

EAW 156

EDEN 124

EAW 30

EDEN 125

EDEN 126

EAW 42

HBSS

SPCS

LDSS

WWCO

BGCS

NPFM

SBSS

BASM

KBSS

TGSM

WBCS

LRSS

ACSM

WTFM

BACS

CCSS

UNM2

ACFM

BGSS

BUSM
CHSM
HGSM

UNM3

DFSS

GRFM

0.7 cm = 1000 m
SE

Appendix 3: Correlation between units from all studied wells using the base of the Hawkesbury Sandstone as a datum surface.
24

NW

1 cm = 26.7 m

Appendix Four

Appendix Five

Appendix 5-1: Rock characteristics of samples in the Illawarra Coal


Measures
Samples

Texture

Sorting

LDSS - EAW 156 (14)


LDSS - EAW 156 (15)
LDSS - EAW 30 (30)
LDSS - EAW 30 (31)
LDSS - EAW 30 (32)
LDSS - EDEN 124 (13)
LDSS - EDEN 124 (14)
LDSS - EDEN 124 (15)
LDSS - EDEN 124 (16)
LDSS - EDEN 125 (15)
LDSS - EDEN 125 (16)
LDSS - EDEN 125 (17)
LDSS - EDEN 126 (16)
LDSS - EDEN 126 (17)
LDSS - EDEN 126 (18)
LDSS - EDEN 127 (15)
LDSS - EDEN 127 (16)
LDSS - EDEN 127 (17)
LDSS - EDEN 127 (18)
LRSS - Surface 1
LRSS - Surface 3
LRSS - Surface 4
LRSS - EAW 156 (18)
LRSS - EAW 156 (19)
LRSS - EAW 156 (20)
LRSS - EAW 30 (33)
LRSS - EAW 30 (34)
LRSS - EAW 30 (35)
LRSS - EDEN 124 (18)
LRSS - EDEN 124 (19)
LRSS - EDEN 125 (37)
LRSS - EDEN 125 (19)
LRSS - EDEN 126 (19)
LRSS - EDEN 126 (20)
LRSS - EDEN 126 (21)
LRSS - EDEN 127 (19)
LRSS - EDEN 127 (20)
UNM2 - EAW 30 (51)
UNM3 - EAW 156 (22)
UNM3 - EDEN 126 (22)
UNM3 - EDEN 127 (21)
WWCO - EDEN 124 (22)
KBSS - Surface 1
KBSS - Surface 3
KBSS - Surface 5
KBSS - Surface 6
KBSS - Surface 7

F to M
M
M
C
M
F to M
C
M to C
C
F to M
M
M to C
F to M
M to C
F to M
M
C
C
M to C
M
F
M to C
F
M
M to C
Cz
F to M
VF
F
F
M
M
F
VF
F
Cz
F
VF
F to M
VF
VF
F to M
C
F to M
M to C
C
M to C

0.7
0.6
0.6
0.5
0.7
0.6
0.6
0.7
0.6
0.6
0.5
0.5
0.7
0.6
0.6
0.6
0.7
0.6
0.5
0.6
0.7
0.6
0.8
0.6
0.6
1.1
0.6
0.7
0.6
0.5
0.5
0.5
0.7
0.9
0.9
0.9
0.7
0.8
0.6
0.8
0.7
0.6
0.9
0.5
0.6
0.8
0.6

17

Roundness
F
RF

SR
R
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR to R
SR
SR to R
SR
SR
SR
R
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR to R
SR to R
SR
SR
SR
SR
SR
SR to R
R
SR
SR
SR
SR

A
A
A
A
A
A
A
A
A
A
A
A
A
SA
A
A
A
A
A
SA
A
A
A
A
A
A
A
A
A
A
A
A
SA
A
A
A
A
A
A
A
A
A
SA
A
A
SA
A

R
R
SR
R
R
R
R
R
SR
SR
SR to R
SR to R
SR
R
R
SR
R
R
SR
SR
SR
R
R
R
R
SR
SR
SR
R
SR
R
SR to R
R
SR
R
SR to R
SR
SR
R
R
SR
SR
R
SR
SR
SR
SR

Type of
rock
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
St
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
St
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss

Appendix 5-1 (Cont.): Rock characteristics of samples in the Illawarra


Coal Measures
Samples

Texture

Sorting

KBSS - EAW 156 (25)


KBSS - EAW 156 (26)
KBSS - EAW 156 (27)
KBSS - EDEN 124 (23)
KBSS - EDEN 124 (24)
KBSS - EDEN 124 (25)
KBSS - EDEN 125 (21)
KBSS - EDEN 125 (22)
KBSS - EDEN 126 (23)
KBSS - EDEN 126 (24)
KBSS - EDEN 126 (25)
KBSS - EDEN 127 (22)
KBSS - EDEN 127 (23)
KBSS - EDEN 127 (24)
ACFM - Surface 1
ACFM - Surface 3
ACFM - EDEN 124 (28)
ACFM - EDEN 126 (28)
DFSS - EDEN 124 (30)
DFSS - EDEN 125 (27)
DFSS - EDEN 125 (28)
DFSS - EDEN 126 (36)
DFSS - EDEN 126 (29)
DFSS - EDEN 127 (29)
DFSS - EDEN 127 (30)
BGCS - EDEN 127 (33)
WTFM - EDEN 124 (35)
WTFM - EDEN 126 (34)
WTFM - EDEN 127 (35)

VF
F
M
M to C
M to C
C
F
C
F to M
M to C
M
F
M
F to M
Cz
F to M
F
F
VF
F
F
F to M
F to M
VF
F to M
Cz
VF
Cz
VF

0.7
0.7
0.7
0.7
0.6
0.7
0.5
0.7
0.6
0.7
0.7
0.5
0.6
0.5
0.9
0.6
0.7
0.8
0.6
0.6
0.6
0.6
0.5
0.6
0.6
1
0.6
0.5
0.5

Roundness
F
RF

SR
SR to R
SR
SR
SR
SR to R
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR to R
SR
SR
SR
SR
SR to R
R
SR

A
A
A
A
A
A
A
A
A
SA
SA
A
A
A
A
A
A
A
A
A
A
SA
A
A
A
A
A
SA
A

SR
R
R
R
R
R
R
R
R
R
R
SR
SR
SR
SR
R
SR
R
SR
SR
SR
R
SR to R
SR
R
SR
SR
R
SR

Type of
rock
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
St
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
St
Ss
St
Ss

Q = Quartz, F = Feldspar, RF = Rock fragment, VF = Very fine, F = Fine, M


= Medium, C = Coarse, VC = Very coarse, Cz = Coarse siltstone, SR = Subrounded, R = Rounded, SA = Sub-angular, A = Angular, SS = Sandstone, St
= Siltstone.

17

Appendix 5-2: Rock characteristics of samples in the Narrabeen Group


Samples

Texture

Sorting

NPFM - EAW 42 (2)


NPFM - EAW 42 (4)
NPFM - EAW 42 (6)
BACS - EAW 42 (14)
BGSS - EAW 42 (18)
BGSS - EAW 42 (19)
BGSS - EAW 42 (21)
BGSS - EAW 42 (22)
BGSS - EAW 42 (23)
BGSS - EAW 42 (24)
BGSS - EAW 42 (25)
BGSS - EAW 42 (27)
BGSS - EAW 42 (29)
BGSS - EAW 42 (30)
BGSS - EAW 42 (31)
BGSS - EAW 42 (32)
SPCS - EAW 42 (34)
SPCS - EAW 42 (35)
SBSS - EAW 42 (37)
SBSS - EAW 42 (38)
SBSS - EAW 42 (39)
SBSS - EAW 42 (42)
SBSS - EAW 42 (+42)
SBSS - EAW 42 (43)
SBSS - EAW 42 (44)
SBSS - EAW 42 (45)
SBSS - EAW 42 (+45)
SBSS - EAW 30 (1)
SBSS - EAW 30 (2)
SBSS - EAW 30 (4)
SBSS - EAW 30 (5)
SBSS - EAW 30 (6)
SBSS - EAW 30 (7)
SBSS - EDEN 125 (1)
SBSS - EDEN 125 (2)
SBSS - EDEN 125 (3)
SBSS - EDEN 125 (4)
WBCS - EAW 156 (1)
WBCS - EAW 156 (5)
WBCS - EAW 156 (6)
WBCS - EAW 42 (47)
WBCS - EAW 42 (48)
WBCS - EAW 42 (+48)
WBCS - EAW 42 (49)

F to M
F
F
M to C
C
M
M
F to M
M
M
M to C
M
M
C
C
C
M
C
M
C
F to M
C
F
M
C
M
C
C
F to M
C
F
F to M
M to C
C
C
VC
M to C
C
F
M to C
VF
C
F to M
VF

0.5
0.7
0.6
0.6
0.5
0.5
0.5
0.7
0.6
0.5
0.5
0.7
0.6
0.5
0.7
0.6
0.6
0.8
0.5
0.7
0.6
0.9
0.5
0.8
0.8
0.6
1
0.7
0.5
0.6
0.7
0.6
0.4
0.6
0.8
0.7
0.5
0.9
0.9
0.9
0.7
0.8
0.5
0.6

17

Roundness
F
RF

SR
SR
SR to R
SR to R
R
SR
SR
SR
SR to R
R
SR to R
SR
SR
R
SR to R
SR to R
SR
SR
SR
SR to R
SR
SR to R
SR
SR
SR
SR to R
SR
SR
SR
SR
R
SR to R
SR to R
SR to R
SR
SR
SR
SR
SR
R
R
SR to R
SR to R
SR to R

A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
A
SA
SA
A

A
A
A
A
A
A
A
A
A

SR
SR to R
SR to R
SR to R
R
SR
SR to R
SR
SR to R
R
SR to R
SR
SR
R
SR to R
SR to R
SR to R
SR to R
SR
R
SR to R
R
R
SR to R
SR
SR to R
SR
SR
SR
SR
SR
SR
SR
R
SR
SR to R
R
SR to R
R
R
R
R
SR to R
SR to R

Type of
rock
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss

Appendix 5-2 (Cont.): Rock characteristics of samples in the Narrabeen


Group
Samples

Texture

Sorting

WBCS - EAW 30 (9)


WBCS - EAW 30 (10)
WBCS - EAW 30 (11)
WBCS - EAW 30 (12)
WBCS - EAW 30 (14)
WBCS - EAW 30 (15)
WBCS - EAW 30 (16)
WBCS - EAW 30 (17)
WBCS - EAW 30 (18)
WBCS - EAW 30 (20)
WBCS - EAW 30 (21)
WBCS - EDEN 124 (2)
WBCS - EDEN 124 (3)
WBCS - EDEN 124 (4)
WBCS - EDEN 124 (5)
WBCS - EDEN 124 (6)
WBCS - EDEN 124 (7)
WBCS - EDEN 125 (6)
WBCS - EDEN 125 (7)
WBCS - EDEN 125 (8)
WBCS - EDEN 125 (10)
WBCS - EDEN 126 (38)
WBCS - EDEN 126 (3)
WBCS - EDEN 126 (4)
WBCS - EDEN 126 (6)
WBCS - EDEN 126 (7)
WBCS - EDEN 126 (8)
WBCS - EDEN 126 (10)
WBCS - EDEN 126 (12)
WBCS - EDEN 127 (2)
WBCS - EDEN 127 (3)
WBCS - EDEN 127 (4)
WBCS - EDEN 127 (5)
WBCS - EDEN 127 (7)
WBCS - EDEN 127 (8)
WBCS - EDEN 127 (10)
CCSS - Surface 1
CCSS - Surface 2
CCSS - Surface 3
CCSS - EAW 156 (12)
CCSS - EAW 156 (13)
CCSS - EAW 30 (22)
CCSS - EAW 30 (23)
CCSS - EAW 30 (24)
CCSS - EAW 30 (25)

Cz
F to M
C
M
M
M
M
C
C
C
Cz
VF
C
M to C
C
M to C
M to C
F to M
M
C
M
F
M
VC
C
C
F to M
VF
F
F
M to C
F to M
C
M to C
M
Cz
C
VC
M
F to M
M
C
M
C
C

1.1
0.6
0.8
0.5
0.5
0.6
0.7
0.7
0.9
0.7
0.8
0.7
0.6
0.6
0.9
0.5
0.4
0.5
0.5
0.8
0.5
0.8
0.7
0.7
0.6
0.7
0.6
0.7
0.7
0.6
0.6
0.6
0.6
1.1
0.8
0.7
0.7
1.2
0.5
0.7
0.6
0.8
0.6
0.7
0.7

17

Roundness
F
RF

SR to R
SR
SR
SR to R
R
SR to R
SR
SR
SR
SR
R
SR
SR
SR
SR to R
SR to R
SR
SR to R
SR
SR
SR to R
R
SR to R
SR
SR
SR
SR to R
R
R
R
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR to R
SR
SR
SR

A
A
SA
A
SA
A
A
A
A
A
A
A
A
A
A
SA
A
A
A
A
A
A
A
A
A
A
A
SA
A
A
A
A
A
A
A
SA
SA
A
A

SR
R
R
R
SR
SR to R
SR
SR to R
R
SR to R
SR
SR
SR
SR
SR to R
SR to R
SR
R
SR
SR to R
R
R
R
R
SR
R
R
R
R
SR
SR
R
SR
SR
SR
SR
SR
SR
SR
R
R
SR
R
R

Type of
rock
St
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
St
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
St
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss

Appendix 5-2 (Cont.): Rock characteristics of samples in the Narrabeen


Group
Samples

Texture

Sorting

CCSS - EAW 30 (26)


CCSS - EAW 30 (54)
CCSS - EDEN 124 (9)
CCSS - EDEN 124 (10)
CCSS - EDEN 124 (11)
CCSS - EDEN 125 (11)
CCSS - EDEN 125 (12)
CCSS - EDEN 125 (14)
CCSS - EDEN 127 (11)
CCSS - EDEN 127 (12)
CCSS - EDEN 127 (13)

M
M
F to M
VF
F to M
VC
VF
VF
M
F to M
F to M

0.6
0.6
0.7
0.6
0.9
0.7
0.8
0.8
0.5
0.6
0.6

Roundness
F
RF

SR
R
SR
SR to R
SR
SR to R
SR to R
SR to R
SR
R
SR

A
A
A
A
A

A
A
A

R
R
R
SR
SR to R
SR to R
SR
SR
SR
R
SR

Type
of rock
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss

Q = Quartz, F = Feldspar, RF = Rock fragment, VF = Very fine, F = Fine, M


= Medium, C = Coarse, VC = Very coarse, Cz = Coarse siltstone, SR = Subrounded, R = Rounded, SA = Sub-angular, A = Angular, SS = Sandstone, St
= Siltstone.

17

Appendix 5-3: Rock characteristics of samples in the Hawkesbury


Sandstone
Samples

Texture

Sorting

HBSS - Surface 6
HBSS - Surface 7
HBSS - Surface 8
HBSS - Surface 9
HBSS - Surface 10
HBSS - Surface 11
HBSS - Surface 12
HBSS - EAW 18 a (1)
HBSS - EAW 18 a (2)
HBSS - EAW 18 a (3)
HBSS - EAW 18 a (4)
HBSS - EAW 18 a (5)
HBSS - EAW 18 a (7)
HBSS - EAW 18 a (8)
HBSS - EAW 18 a (9)
HBSS - EAW 18 a (10)
HBSS - EDEN 115 (1)
HBSS - EDEN 115 (2)
HBSS - EDEN 115 (3)
HBSS - EDEN 115 (4)
HBSS - EDEN 115 (5)
HBSS - EDEN 115 (6)
HBSS - EDEN 115 (7)
HBSS - EDEN 115 (8)
HBSS - EDEN 115 (9)
HBSS - EDEN 115 (10)
HBSS - EDEN 115 (11)
HBSS - EDEN 115 (12)
HBSS - EDEN 115 (13)
HBSS - EDEN 115 (14)
HBSS - EDEN 115 (15)
HBSS - EDEN 115 (16)

M to C
M
M
M
M to C
M to C
C
F
C
M
C
M
C
Cz
M
M
C
M
C
M
F
C
Cz
C
M to C
M
Cz
M
M to C
C
C
M

0.5
0.8
0.5
0.7
0.6
0.6
0.6
0.7
0.5
0.8
0.5
0.6
0.6
1.3
0.6
0.6
0.7
0.6
0.9
0.7
0.7
0.7
0.7
0.4
0.7
1
0.9
0.9
0.6
0.9
0.7
0.7

Roundness
Q
F
SR
SR
SR
SR
R
SR
SR
SR
SR to R
R
SR
SR
SR to R
SR to R
SR
SR
SR to R
SR to R
R
SR
SR to R
SR
R
SR
SR
SR to R
R
SR
R
SR
SR to R
SR to R

SA
A
A

A
A
A

A
A

A
A

RF
SR
SR
SR
SR
R
R
SR
R
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
SR
R
SR
SR
SR
SR
SR
SR
SR

Type of
rock
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
St
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
St
Ss
Ss
Ss
St
Ss
Ss
Ss
Ss
Ss

Q = Quartz, F = Feldspar, RF = Rock fragment, VF = Very fine, F = Fine, M


= Medium, C = Coarse, VC = Very coarse, Cz = Coarse siltstone, SR = Subrounded, R = Rounded, SA = Sub-angular, A = Angular, SS = Sandstone, St
= Siltstone.

17

Appendix 5-4: Classification of the Illawarra Coal Measures


according to Folk (1968) and Dickinson (1985)
Samples
LDSS - EAW 156 (14)
LDSS - EAW 156 (15)
LDSS - EAW 30 (30)
LDSS - EAW 30 (31)
LDSS - EAW 30 (32)
LDSS - EDEN 124 (13)
LDSS - EDEN 124 (14)
LDSS - EDEN 124 (15)
LDSS - EDEN 124 (16)
LDSS - EDEN 125 (15)
LDSS - EDEN 125 (16)
LDSS - EDEN 125 (17)
LDSS - EDEN 126 (16)
LDSS - EDEN 126 (17)
LDSS - EDEN 126 (18)
LDSS - EDEN 127 (15)
LDSS - EDEN 127 (16)
LDSS - EDEN 127 (17)
LDSS - EDEN 127 (18)
LRSS - Surface 1
LRSS - Surface 3
LRSS - Surface 4
LRSS - EAW 156 (18)
LRSS - EAW 156 (19)
LRSS - EAW 156 (20)
LRSS - EAW 30 (33)
LRSS - EAW 30 (34)
LRSS - EAW 30 (35)
LRSS - EDEN 124 (18)
LRSS - EDEN 124 (19)
LRSS - EDEN 125 (37)
LRSS - EDEN 125 (19)
LRSS - EDEN 126 (19)
LRSS - EDEN 126 (20)
LRSS - EDEN 126 (21)
LRSS - EDEN 127 (19)
LRSS - EDEN 127 (20)
UNM2 - EAW 30 (51)
UNM3 - EAW 156 (22)
UNM3 - EDEN 126 (22)
UNM3 - EDEN 127 (21)
WWCO - EDEN 124 (22)
KBSS - Surface 1
KBSS - Surface 3
KBSS - Surface 5
KBSS - Surface 6

QFR
Q

3.8
2.7
3.2
3.4
2.0
2.6
3.0
2.1
1.8
1.5
1.9
2.3
1.5
3.6
3.7
2.7
2.6
3.0
2.7
2.6
2.0
0.2
3.2
4.4
3.2
4.5
3.9
4.7
2.2
3.4
2.6
2.7
3.6
4.8
5.3
6.9
5.1
2.8
2.1
10.5
3.3
1.9
0.9
1.6
1.0
0.9

19.3
19.0
33.0
28.5
29.8
12.9
16.6
13.8
11.9
26.1
24.3
22.7
17.4
19.9
21.5
26.3
13.1
14.9
14.2
24.4
32.7
18.8
38.8
17.2
19.4
66.9
49.3
73.0
14.4
28.5
28.2
25.0
25.6
61.4
63.9
79.9
40.3
52.8
19.9
58.2
43.9
27.6
7.6
24.5
11.9
1.3

76.9
78.3
63.8
68.2
68.2
84.6
80.4
84.1
86.4
72.4
73.8
75.0
81.0
76.5
74.8
71.0
84.3
82.0
83.1
72.9
65.3
80.9
58.0
78.3
77.4
28.6
46.8
22.3
83.3
68.1
69.2
72.3
70.8
33.8
30.8
13.2
54.6
44.4
78.0
31.4
52.8
70.6
91.5
73.9
87.0
97.8

11

FQmLt
Qm

Lt

4.1
2.8
3.4
3.7
2.2
2.6
3.1
2.1
1.8
1.6
2.1
2.4
1.6
4.0
3.8
2.9
2.8
3.3
2.9
2.9
2.1
0.2
3.5
4.7
3.4
4.6
4.6
5.2
2.3
3.5
2.7
2.9
3.8
5.3
5.7
7.2
5.6
3.1
2.2
11.6
3.6
1.9
0.9
1.8
1.0
0.9

13.7
14.8
27.3
20.9
21.8
10.7
12.2
11.9
9.6
23.9
14.9
19.1
14.0
9.7
17.7
18.5
6.3
8.4
7.8
17.9
29.0
15.6
32.8
12.3
14.5
65.9
41.0
70.5
10.8
25.3
24.8
20.8
21.8
57.4
61.1
79.1
34.0
47.4
16.2
53.8
38.3
24.9
6.7
19.5
10.8
1.3

82.2
82.4
69.2
75.4
76.0
86.7
84.7
85.9
88.6
74.6
83.0
78.5
84.4
86.2
78.5
78.6
90.9
88.3
89.3
79.3
68.9
84.2
63.7
83.1
82.2
29.5
54.4
24.4
86.9
71.2
72.5
76.4
74.4
37.4
33.1
13.7
60.4
49.5
81.6
34.7
58.1
73.2
92.3
78.8
88.2
97.8

Appendix 5-4 (Cont.): Classification of the Illawarra Coal Measures


according to Folk (1968) and Dickinson (1985)
Samples
KBSS - Surface 7
KBSS - EAW 156 (25)
KBSS - EAW 156 (26)
KBSS - EAW 156 (27)
KBSS - EDEN 124 (23)
KBSS - EDEN 124 (24)
KBSS - EDEN 124 (25)
KBSS - EDEN 125 (21)
KBSS - EDEN 125 (22)
KBSS - EDEN 126 (23)
KBSS - EDEN 126 (24)
KBSS - EDEN 126 (25)
KBSS - EDEN 127 (22)
KBSS - EDEN 127 (23)
KBSS - EDEN 127 (24)
ACFM - Surface 1
ACFM - Surface 3
ACFM - EDEN 124 (28)
ACFM - EDEN 126 (28)
DFSS - EDEN 124 (30)
DFSS - EDEN 125 (27)
DFSS - EDEN 125 (28)
DFSS - EDEN 126 (36)
DFSS - EDEN 126 (29)
DFSS - EDEN 127 (29)
DFSS - EDEN 127 (30)
BGCS - EDEN 127 (33)
WTFM - EDEN 124 (35)
WTFM - EDEN 126 (34)
WTFM - EDEN 127 (35)

F
1.7
4.1
0.7
2.1
1.8
1.6
1.6
2.9
2.5
3.8
2.4
2.7
2.7
1.9
3.2
1.3
1.8
4.5
3.3
4.4
2.5
5.6
4.5
3.8
2.8
3.3
0.9
2.3
4.7
1.8

QFR
Q
9.8
27.0
24.3
15.9
13.7
6.6
9.2
65.5
9.5
30.0
12.9
24.7
33.8
12.1
32.3
85.4
19.3
20.6
19.7
16.9
35.2
25.5
28.3
28.4
55.1
31.6
79.7
20.6
62.4
42.8

88.6
69.0
75.0
82.1
84.6
91.8
89.1
31.6
88.0
66.1
84.8
72.5
63.5
86.0
64.5
13.2
78.9
74.9
77.0
78.7
62.3
69.0
67.2
67.8
42.1
65.1
19.4
77.1
32.9
55.4

1.7
4.4
0.7
2.1
1.9
1.6
1.7
3.2
2.6
4.4
2.4
2.9
2.8
2.0
3.4
1.6
1.9
4.9
3.3
4.6
2.8
5.8
5.2
4.4
3.4
3.5
0.9
2.4
5.2
1.9

FQmLt
Qm
8.5
20.7
21.9
12.0
8.8
6.6
8.0
61.7
6.3
19.1
9.9
19.6
30.4
9.5
28.1
82.3
15.1
14.1
18.5
11.6
28.9
22.4
18.2
17.7
45.9
28.3
79.5
17.3
58.8
40.5

Lt
89.9
74.9
77.4
85.8
89.3
91.8
90.3
35.1
91.1
76.5
87.6
77.5
66.8
88.5
68.6
16.1
83.0
81.0
78.1
83.8
68.3
71.8
76.6
77.9
50.7
68.2
19.5
80.3
36.1
57.6

Q = Quartz, F = Feldspar, R = Rock fragment + chert, Qm = Monocrystalline


quartz, Lt = Rock fragment + chert.

17

Appendix 5-5: Classification of the Narrabeen Group according to


Folk (1968) and Dickinson (1985)
Samples
NPFM - EAW 42 (2)
NPFM - EAW 42 (4)
NPFM - EAW 42 (6)
BACS - EAW 42 (14)
BGSS - EAW 42 (18)
BGSS - EAW 42 (19)
BGSS - EAW 42 (21)
BGSS - EAW 42 (22)
BGSS - EAW 42 (23)
BGSS - EAW 42 (24)
BGSS - EAW 42 (25)
BGSS - EAW 42 (27)
BGSS - EAW 42 (29)
BGSS - EAW 42 (30)
BGSS - EAW 42 (31)
BGSS - EAW 42 (32)
SPCS - EAW 42 (34)
SPCS - EAW 42 (35)
SBSS - EAW 42 (37)
SBSS - EAW 42 (38)
SBSS - EAW 42 (39)
SBSS - EAW 42 (42)
SBSS - EAW 42 (+42)
SBSS - EAW 42 (43)
SBSS - EAW 42 (44)
SBSS - EAW 42 (45)
SBSS - EAW 42 (+45)
SBSS - EAW 30 (1)
SBSS - EAW 30 (2)
SBSS - EAW 30 (4)
SBSS - EAW 30 (5)
SBSS - EAW 30 (6)
SBSS - EAW 30 (7)
SBSS - EDEN 125 (1)
SBSS - EDEN 125 (2)
SBSS - EDEN 125 (3)
SBSS - EDEN 125 (4)
WBCS - EAW 156 (1)
WBCS - EAW 156 (5)
WBCS - EAW 156 (6)
WBCS - EAW 42 (47)
WBCS - EAW 42 (48)
WBCS - EAW 42 (+48)
WBCS - EAW 42 (49)
WBCS - EAW 30 (9)
WBCS - EAW 30 (10)

F
0.9
0.3
0.7
1.7
1.2
0.4
0.7
2.3
1.3
0.6
0.9
2.0
0.9
2.9
1.3
0.4
3.7
3.8
3.8
2.1
4.3
3.8
4.3
3.9
2.1
3.4
3.6
1.0
0.5
3.5
1.7
0.2
0.3
0.0
0.0
0.7
0.7
4.2
1.9
2.9
2.8
2.1
0.9
2.5
0.0
1.6

QFR
Q
97.9
85.5
83.5
53.2
93.3
86.2
87.6
71.9
89.2
71.8
86.8
78.8
86.3
85.0
88.8
86.4
35.5
31.3
49.0
48.4
71.4
25.1
55.1
23.6
52.0
28.6
57.6
44.8
42.8
43.9
65.4
67.5
35.2
49.2
58.2
37.3
45.6
18.7
29.3
24.0
93.8
68.6
77.1
75.8
98.0
41.3

17

1.1
14.2
15.8
45.1
5.5
13.4
11.7
25.9
9.4
27.6
12.3
19.2
12.8
12.1
9.9
13.1
60.8
64.9
47.3
49.6
24.3
71.1
40.5
72.5
45.9
67.9
38.8
54.3
56.8
52.6
32.9
32.3
64.5
50.8
41.8
62.0
53.8
77.2
68.8
73.0
3.5
29.3
22.0
21.7
2.0
57.1

0.9
0.3
0.7
2.8
1.6
0.5
0.8
2.5
1.5
0.6
1.0
2.3
1.0
3.2
1.5
0.5
4.3
4.5
4.5
2.5
5.1
4.0
5.0
4.1
2.3
3.9
3.9
1.0
0.5
3.8
1.7
0.2
0.3
0.0
0.0
0.7
0.7
4.4
2.0
3.3
2.9
2.4
1.1
2.9
0.0
1.7

FQmLt
Qm
97.9
84.4
82.7
22.7
91.4
84.1
85.7
68.9
88.1
70.5
85.6
75.0
84.2
83.5
87.6
84.4
24.6
17.5
39.3
38.2
66.0
21.4
48.2
19.4
46.5
17.9
54.5
39.5
38.5
38.9
64.2
66.4
31.3
45.4
53.6
30.5
41.7
14.0
25.8
16.3
93.5
63.8
73.1
71.7
98.0
37.2

Lt
1.2
15.3
16.6
74.5
7.0
15.5
13.5
28.5
10.4
28.9
13.4
22.7
14.8
13.4
10.9
15.1
71.1
78.0
56.2
59.3
28.9
74.6
46.8
76.5
51.2
78.1
41.7
59.5
61.0
57.4
34.1
33.4
68.4
54.6
46.4
68.8
57.6
81.6
72.2
80.5
3.6
33.7
25.8
25.4
2.0
61.1

Appendix 5-5 (Cont.): Classification of the Narrabeen Group


according to Folk (1968) and Dickinson (1985)
Samples
WBCS - EAW 30 (11)
WBCS - EAW 30 (12)
WBCS - EAW 30 (14)
WBCS - EAW 30 (15)
WBCS - EAW 30 (16)
WBCS - EAW 30 (17)
WBCS - EAW 30 (18)
WBCS - EAW 30 (20)
WBCS - EAW 30 (21)
WBCS - EDEN 124 (2)
WBCS - EDEN 124 (3)
WBCS - EDEN 124 (4)
WBCS - EDEN 124 (5)
WBCS - EDEN 124 (6)
WBCS - EDEN 124 (7)
WBCS - EDEN 125 (6)
WBCS - EDEN 125 (7)
WBCS - EDEN 125 (8)
WBCS - EDEN 125 (10)
WBCS - EDEN 126 (38)
WBCS - EDEN 126 (3)
WBCS - EDEN 126 (4)
WBCS - EDEN 126 (6)
WBCS - EDEN 126 (7)
WBCS - EDEN 126 (8)
WBCS - EDEN 126 (10)
WBCS - EDEN 126 (12)
WBCS - EDEN 127 (2)
WBCS - EDEN 127 (3)
WBCS - EDEN 127 (4)
WBCS - EDEN 127 (5)
WBCS - EDEN 127 (7)
WBCS - EDEN 127 (8)
WBCS - EDEN 127 (10)
CCSS - Surface 1
CCSS - Surface 2
CCSS - Surface 3
CCSS - EAW 156 (12)
CCSS - EAW 156 (13)
CCSS - EAW 30 (22)
CCSS - EAW 30 (23)
CCSS - EAW 30 (24)
CCSS - EAW 30 (25)
CCSS - EAW 30 (26)
CCSS - EAW 30 (54)
CCSS - EDEN 124 (9)

F
0.9
0.7
0.6
0.9
1.3
1.5
1.0
0.0
1.6
0.0
0.6
0.0
0.8
0.6
0.0
0.8
0.0
0.9
0.7
1.3
1.5
1.1
1.3
0.9
1.5
1.6
2.1
0.8
0.3
0.7
0.3
1.6
2.0
4.5
1.4
0.9
1.2
2.6
0.6
1.5
0.8
1.2
0.8
1.0
0.9
0.3

QFR
Q
44.0
57.6
65.5
64.7
33.8
35.4
34.4
26.8
89.1
50.4
46.5
49.3
19.8
31.1
35.2
56.9
44.4
24.9
42.1
51.1
54.1
9.8
11.3
21.6
40.8
66.1
56.8
36.4
34.6
32.5
38.7
34.2
29.0
56.3
19.5
6.3
29.3
45.6
45.5
45.2
33.7
30.1
30.1
49.2
54.4
36.2

78

55.1
41.7
33.9
34.3
64.9
63.1
64.6
73.2
9.4
49.6
53.0
50.7
79.4
68.4
64.8
42.3
55.6
74.2
57.2
47.6
44.4
89.1
87.4
77.5
57.6
32.3
41.2
62.9
65.1
66.8
61.1
64.2
69.0
39.2
79.1
92.8
69.5
51.8
53.9
53.3
65.5
68.7
69.0
49.7
44.7
63.5

1.0
0.7
0.7
1.1
1.4
1.6
1.1
0.0
1.6
0.0
0.6
0.0
0.8
0.6
0.0
0.9
0.0
1.0
0.7
1.4
1.7
1.1
1.3
0.9
1.7
1.7
2.3
0.9
0.3
0.8
0.3
1.7
2.1
5.4
1.4
0.9
1.2
3.1
0.7
1.7
0.9
1.2
0.9
1.1
1.0
0.3

FQmLt
Qm
39.5
55.2
61.7
60.6
28.7
30.9
30.7
25.9
88.9
41.0
41.6
43.8
18.2
23.1
25.7
51.2
37.7
20.4
36.5
45.3
46.6
6.7
7.9
15.5
35.5
62.6
52.8
29.5
30.2
28.4
31.4
30.1
22.4
47.9
17.4
5.5
27.9
36.8
37.5
39.9
26.9
27.3
21.8
45.7
50.2
30.4

Lt
59.5
44.0
37.7
38.3
69.9
67.5
68.3
74.1
9.5
59.0
57.8
56.2
81.0
76.3
74.3
47.9
62.3
78.6
62.8
53.3
51.7
92.2
90.8
83.6
62.8
35.6
45.0
69.6
69.5
70.8
68.3
68.2
75.5
46.7
81.2
93.6
70.9
60.1
61.8
58.5
72.2
71.5
77.2
53.3
48.9
69.3

Appendix 5-5 (Cont.): Classification of the Narrabeen Group


according to Folk (1968) and Dickinson (1985)
Samples
CCSS - EDEN 124 (10)
CCSS - EDEN 124 (11)
CCSS - EDEN 125 (11)
CCSS - EDEN 125 (12)
CCSS - EDEN 125 (14)
CCSS - EDEN 127 (11)
CCSS - EDEN 127 (12)
CCSS - EDEN 127 (13)

F
0.6
0.0
0.5
0.0
0.0
1.8
1.2
1.4

QFR
Q
39.4
43.9
23.1
66.0
96.0
28.6
29.9
27.5

59.9
56.1
76.4
34.0
4.0
69.5
68.8
71.1

0.7
0.0
0.5
0.0
0.0
1.9
1.3
1.5

FQmLt
Qm
32.9
37.8
17.1
64.1
95.9
24.3
24.4
21.6

Lt
66.4
62.2
82.4
35.9
4.1
73.8
74.2
76.8

Q = Quartz, F = Feldspar, R = Rock fragment + chert, Qm =


Monocrystalline quartz, Lt = Rock fragment + chert.

77

Appendix 5-6: Classification of the Hawkesbury Sandstone according


to Folk (1968) and Dickinson (1985)
Samples
HBSS - Surface 6
HBSS - Surface 7
HBSS - Surface 8
HBSS - Surface 9
HBSS - Surface 10
HBSS - Surface 11
HBSS - Surface 12
HBSS - EAW 18 a (1)
HBSS - EAW 18 a (2)
HBSS - EAW 18 a (3)
HBSS - EAW 18 a (4)
HBSS - EAW 18 a (5)
HBSS - EAW 18 a (7)
HBSS - EAW 18 a (8)
HBSS - EAW 18 a (9)
HBSS - EAW 18 a (10)
HBSS - EDEN 115 (1)
HBSS - EDEN 115 (2)
HBSS - EDEN 115 (3)
HBSS - EDEN 115 (4)
HBSS - EDEN 115 (5)
HBSS - EDEN 115 (6)
HBSS - EDEN 115 (7)
HBSS - EDEN 115 (8)
HBSS - EDEN 115 (9)
HBSS - EDEN 115 (10)
HBSS - EDEN 115 (11)
HBSS - EDEN 115 (12)
HBSS - EDEN 115 (13)
HBSS - EDEN 115 (14)
HBSS - EDEN 115 (15)
HBSS - EDEN 115 (16)

F
0.0
0.3
0.0
0.0
0.7
0.2
0.4
0.0
0.0
0.0
0.5
0.0
0.6
1.1
0.0
0.0
0.5
0.0
0.0
0.0
0.0
0.0
2.0
0.7
0.0
0.0
0.8
0.3
0.0
0.0
0.0
0.0

QFR
Q
85.7
90.6
88.5
73.1
96.0
80.7
89.9
77.9
94.3
74.3
91.0
91.5
90.8
88.7
94.6
88.9
96.9
97.6
90.1
91.2
79.7
94.6
98.0
91.1
95.2
94.6
80.6
97.1
93.9
93.1
93.1
99.4

14.3
9.1
11.5
26.9
3.3
19.1
9.7
22.1
5.7
25.7
8.6
8.5
8.7
10.2
5.4
11.1
2.6
2.4
9.9
8.8
20.3
5.4
0.0
8.2
4.8
5.4
18.6
2.6
6.1
6.9
6.9
0.6

0.0
0.3
0.0
0.0
0.8
0.3
0.4
0.0
0.0
0.0
0.5
0.0
0.6
1.1
0.0
0.0
0.6
0.0
0.0
0.0
0.0
0.0
2.0
0.7
0.0
0.0
0.8
0.3
0.0
0.0
0.0
0.0

FQmLt
Qm
85.3
90.4
88.1
72.7
95.9
80.1
89.5
76.6
94.1
72.8
90.7
91.3
90.6
88.4
94.0
88.1
96.5
97.3
89.7
90.8
79.4
94.5
98.0
90.7
95.1
94.5
80.3
97.0
93.6
92.7
93.0
99.4

Lt
14.7
9.2
11.9
27.3
3.3
19.7
10.0
23.4
5.9
27.2
8.8
8.7
8.9
10.4
6.0
11.9
2.9
2.7
10.3
9.2
20.6
5.5
0.0
8.5
4.9
5.5
18.8
2.6
6.4
7.3
7.0
0.6

Q = Quartz, F = Feldspar, R = Rock fragment + chert, Qm = Monocrystalline


quartz, Lt = Rock fragment + chert.

77

Appendix 5-7: Point count results of samples from the Illawarra Coal Measures

77

Appendix 5-7 (Cont.): Point count results of samples from the Illawarra Coal Measures

So = Sorting, Q = Quartz, Qm = Monocrystalline quartz, Qpu = Plutonic quartz, Qv = Volcanic quartz, Qp = Polycrystalline quartz, TQ = Total quartz, F =
Feldspar, Plag = Plagioclase, K-f = K-feldspar, TF = Total feldspar, RF = Rock fragment, SRF = Sedimentary rock fragment, VRF = Volcanic rock fragment,
MRF = Metamorphic rock fragment, Sa = Sand rock fragment, Si = Silt fragment, Shr = Shale fragment, Cf = Carbonate fragment, Vi = Vitric fragment, Fe =
Felsitic fragment, Mi = Microlithic fragment, Lath = Lathwork fragment, Ch = Chert, TRF = Total rock fragment, Chl = Chlorite, Mu = Muscovite, HML = Heavy
minerals, Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, THML = Total heavy minerals, C = Cement, Qo = Quartz overgrowth, I = Iron oxide, Py =
Pyrite, Carb = Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, TC = Total cement, Ma = Matrix,
TPo = Total thin section porosity, GSZ = Grain size.

77

Appendix 5-8: Point count results of samples from the Narrabeen Group

77

Appendix 5-8 (Cont.): Point count results of samples from the Narrabeen
Group

So = Sorting, Q = Quartz, Qm = Monocrystalline quartz, Qpu = Plutonic quartz, Qv = Volcanic quartz, Qp = Polycrystalline quartz, TQ = Total quartz, F = Feldspar, Plag = Plagioclase, K-f = Kfeldspar, TF = Total feldspar, RF = Rock fragment, SRF = Sedimentary rock fragment, VRF = Volcanic rock fragment, MRF = Metamorphic rock fragment, Sa = Sand rock fragment, Si = Silt
fragment, Shr = Shale fragment, Cf = Carbonate fragment, Vi = Vitric fragment, Fe = Felsitic fragment, Mi = Microlithic fragment, Lath = Lathwork fragment, Ch = Chert, TRF = Total rock
fragment, Chl = Chlorite, Mu = Muscovite, HML = Heavy minerals, Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, THML = Total heavy minerals, C = Cement, Qo = Quartz overgrowth,
I = Iron oxide, Py = Pyrite, Carb = Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, TC = Total cement, Ma = Matrix, TPo = Total thin
section porosity, GSZ = Grain size.

77

Appendix 5-9: Point count results of samples from the Hawkesbury Sandstone

So = Sorting, Q = Quartz, Qm = Monocrystalline quartz, Qpu = Plutonic quartz, Qv = Volcanic quartz, Qp = Polycrystalline quartz, TQ = Total quartz, F =
Feldspar, Plag = Plagioclase, K-f = K-feldspar, TF = Total feldspar, RF = Rock fragment, SRF = Sedimentary rock fragment, VRF = Volcanic rock fragment, MRF
= Metamorphic rock fragment, Sa = Sand rock fragment, Si = Silt fragment, Shr = Shale fragment, Cf = Carbonate fragment, Vi = Vitric fragment, Fe = Felsitic
fragment, Mi = Microlithic fragment, Lath = Lathwork fragment, Ch = Chert, TRF = Total rock fragment, Chl = Chlorite, Mu = Muscovite, HML = Heavy minerals,
Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, THML = Total heavy minerals, C = Cement, Qo = Quartz overgrowth, I = Iron oxide, Py = Pyrite, Carb =
Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, TC = Total cement, Ma = Matrix, TPo = Total thin
section porosity, GSZ = Grain size.

71

Appendix 5-10: Average of components from point count results of samples from the Illawarra Coal Measures

Appendix 5-11: Average of components from point count results of samples from the Narrabeen Group

Appendix 5-12: Average of components from point count results of samples from the Hawkesbury Sandstone

Q = Quartz, Qm = Monocrystalline quartz, Qp = Polycrystalline quartz, TQ = Total quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar, TF = Total feldspar, RF = Rock fragment, SRF
= Sedimentary rock fragment, VRF = Volcanic rock fragment, MRF = Metamorphic rock fragment, TRF = Total rock fragment, Ch = Chert, Chl = Chlorite, Mu = Muscovite, HML = Heavy
minerals, Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, THML = Total heavy minerals, C = Cement, Qo = Quartz overgrowth, I = Iron oxide, Py = Pyrite, Carb = Carbonate
cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, TC = Total cement, Ma = Matrix, TPo = Total thin section porosity, GSZ = Grain size.

77

Appendix 5-13: Average of point count results of samples from the southern Sydney Basin.
Fm

HBSS
NPFM
BACS
BGSS
SPCS
SSBS
WBCS
CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS
WTFM

32
3
1
12
2
19
43
20
19
18
1
3
1
19
4
7
1
3

RF

Ch

Qm

Qp

TQ

Plag

K-f

TF

SRF

VRF

MRF

TRF

58.9
63.8
8.2
53.3
9.7
23.9
21.9
19.7
9.4
15.4
15.3
10.2
14.2
8.9
10.2
9.4
17.1
12.8

2.6
3.8
23.5
8.8
8.6
5.6
5.1
4.7
4.3
3.4
3.7
2.5
2.1
2.5
2.7
4
0.2
2

61.5
67.6
31.7
62.1
18.3
29.5
27
24.4
13.7
18.8
19
12.8
16.3
11.4
12.9
13.4
17.3
14.8

0.03
0.3
0.6
0.4
1
0.6
0.4
0.4
1.4
1.4
0.9
1.3
0.8
1.1
1.3
1.6
0.2
0.8

0.1
0.2
0.4
0.5
1.1
0.7
0.3
0.2
0.4
0.3
0.1
0.2
0.3
0.3
0.2
0.2
0
0.2

0.1
0.5
1
0.9
2.1
1.3
0.7
0.7
1.8
1.7
1
1.5
1.1
1.4
1.5
1.8
0.2
1.1

1.4
1.4
6
1
5.7
3.3
3.4
3
3.9
1.2
0
1.6
3.5
1.6
0.7
1.1
0
1.1

0
5.9
20.9
9.4
28.9
28.9
31.9
38.8
48.6
32.7
16
21.8
38.2
52.8
38.1
29.1
4.2
23.5

0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

1.4
7.3
26.9
10.4
34.6
32.2
35.3
41.8
52.4
33.8
16
23.5
41.7
54.4
38.7
30.2
4.2
24.7

5.1
1.4
15.2
2.9
14.4
20.9
28.2
34.7
42.5
29.2
3
19.8
37
45.7
35.8
25.9
4.2
16.7

Chl

0.04
0.03
0.6
0.03
0.2
0.1
0.1
0.1
0.2
0.1
0
0
0
0.04
0.03
0
0
0.07

0.3
0.9
0.4
0
0.2
0.3
0.7
0.5
1.1
0.8
1.1
1
0
0.2
0.8
0.7
0.2
0.6

HML

Mu
Ho

THML

0.1
1.2
1.3
1
1.2
1.1
0.9
0.8
0.7
0.7
0.3
0.6
0.7
0.7
0.7
1
0.2
0.5

0.04
0.2
0
0.1
0
0.01
0.002
0
0
0.01
0
0
0
0.005
0
0
0
0

0.03
0
0.1
0
0
0.01
0.002
0.01
0.04
0.03
0
0
0
0.01
0
0
0
0

0.3
0.1
0.8
0.3
0.3
0.2
0.2
0.3
0.2
0.1
0
0.07
0.6
0.3
0.3
0.2
0
0.2

0.5
1.6
2.2
1.3
1.5
1.4
1.1
1
1
0.8
0.3
0.7
1.3
1
1.1
1.2
0.2
0.7

Qo

Py

6.3
1.5
0.2
3.4
2.6
2.0
1.4
1.1
0.6
0.5
0.4
0.3
0.8
0.3
0.4
0.2
0.4
0.4

0.5
0.1
0.4
0
0.1
0.1
0.1
0.03
0.1
0.1
0
0
0
0
5.2
0
0
0.03

0.03
0.1
0.1
0
0.1
0.05
0.02
0.03
0.04
0.03
0
0.03
0
0.02
0.05
0.01
0
0.03

Ca

Carb
A

Tcarb

3
5.4
19.5
4.3
4.3
9.1
6.1
3.4
1.6
3
5.8
6.7
2.5
1.1
0.2
9.1
16.2
3

0.1
0
0.2
0
0.6
0.3
0.7
2.3
0.9
1.6
0.3
1.3
0
4.3
2.6
10.6
0
6.4

0.6
1.4
0.6
0.6
9.8
1.1
3.5
3.5
2.5
6.5
12.4
7.6
2
5.9
5.7
16.1
17.4
6.5

0.3
0.3
0.2
0.1
3.8
0.9
2.3
1.6
1.2
2.7
6.2
4.4
0.3
1.7
0.9
3.7
4.9
4.1

4
7.1
20.5
5
18.5
11.4
12.7
10.8
6.3
13.8
24.7
20
4.8
13
9.4
39.4
38.5
20

Q = Quartz, Qm = Monocrystalline quartz, Qp = Polycrystalline quartz, TQ = Total quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar,
TF = Total feldspar, RF = Rock fragment, SRF = Sedimentary rock fragment, VRF = Volcanic rock fragment, MRF = Metamorphic rock
fragment, TRF = Total rock fragment, Ch = Chert, Chl = Chlorite, Mu = Muscovite, HML = Heavy minerals, Ho, Hornblende, R = Rutile, Z =
Zircon, T = Tourmaline, THML = Total heavy minerals, C = Cement, Qo = Quartz overgrowth, I = Iron oxide, Py = Pyrite, Carb = Carbonate
cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, TC = Total cement, Ma = Matrix, TPo =
Total thin section porosity, GSZ = Grain size.

77

TC
10.8
8.9
21.2
8.3
21.3
13.5
14.1
11.9
7
14.4
25.1
20.3
5.6
13.4
15.1
39.7
38.9
20.5

Ma

TPo

GSZ

12
12.4
14.5
5
15.4
15.3
17.4
15
20.2
28.1
37.5
40.3
34
17.1
29.9
13
39
37.7

8.2
0.8
1.5
12
6.8
6.5
3.5
4.5
2.6
1.4
0
0
0
1.1
0
0
0
0

359.3
164
498
422.9
446
450.6
404.8
392.1
464.2
199.5
61.5
73.6
195
617
79
132.7
55.7
60.2

Appendix 5-14: Average of point count results of subsurface samples from the southern Sydney Basin.

Fm

HBSS
NPFM
GRFM
BACS
BGSS
SPCS
SSBS
WBCS
CCSS
BUSM
LDSS
BASM
LRSS
CHSM
UNM2
HGSM
UNM3
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

Av
thick
138.4
13.9
2.4
21.8
135.1
14.4
36.3
30.8
15.4
3
7.5
1.9
6.8
0.4
12.9
0.5
5.8
10.5
6.2
5.9
5.5
5.7
11.4
6.4
5.2

Base
of
unit

Av C
thick

365.7
351.8
349.4
328
192.5
178
141.8
111.0
95.6
92.6
85.1
83.2
76.4
76
63.1
62.6
57
46.3
40.1
34.2
28.7
23
11.6
5.2
0

434.9
358.8
350.6
338.5
260.1
185.3
160
126.4
103.3
94.1
88.9
84.2
79.8
76.2
69.6
62.9
59.7
51.6
43.2
37.2
31.5
25.9
17.3
8.4
2.6

RF

Ch

Chl

HML

Mu

Ma

TPo

GSZ

11.8
8.9

12.7
12.4

7.5
0.8

343.2
164

20.5
5
18.5
11.4
12.7
11.9

21.2
8.3
21.3
13.5
14.1
13.1

14.5
5
15.4
15.3
17.4
16.8

1.5
12
6.8
6.5
3.5
5.1

498
422.9
446
450.6
404.8
352.1

1.2

6.3

20.2

2.6

464.2

7.5

3.2

15.7

16.3

31.3

1.7

182.6

0.3

12.4

6.2

24.7

25.1

37.5

61.53

6.7
2.5
1.6

1.3
0
2.2

7.6
2
7.6

4.4
0.3
1.8

20
4.8
13.2

20.3
5.6
13.6

40.3
34
20

0
0
1.3

73.6
195
477.7

0.05
0.01
0

0.3
9.1
16.2

4.9
10.6
0

11.3
16
17.4

1.8
3.7
4.9

18.3
39.4
38.5

18.9
39.7
38.9

23.3
13
39

0
0
0

64.9
132.7
55.7

0.03

6.4

6.5

4.1

20

20.5

37.7

60.2

Qo

Py

0.4
1.6

6.4
1.5

0.3
0.1

0.8
0.3
0.3
0.2
0.2
0.3

2.2
1.3
1.5
1.4
1.1
1

0.2
3.4
2.6
2.0
1.4
1.1

0.04

0.2

0.01

0.1

0.3

1
0
0.3

0.6
0.7
0.7

0
0
0.007

0
0
0

0.4
0.7
0.2

0.7
1
0.2

0.07

0.6

0.5

Qm

Qp

TQ

Plag

K-f

TF

SRV

VRF

MRF

TRF

Ho

THML

25
3

58.6
63.8

2.8
3.8

61.5
67.6

0.02
0.3

0.1
0.2

0.1
0.5

1.4
1.4

0
5.9

0
0

1.4
7.3

4.3
1.4

0.04
0.03

0.3
0.9

0.1
1.2

0.04
0.2

0.04
0

0.2
0.1

1
12
2
19
43
17

8.2
53.3
9.7
23.9
21.9
20.7

23.5
8.8
8.6
5.6
5.1
5.2

31.7
62.1
18.3
29.5
27
25.9

0.6
0.4
1
0.6
0.4
0.4

0.4
0.5
1.1
0.7
0.3
0.2

1
0.9
2.1
1.3
0.7
0.6

6
1
5.7
3.3
3.4
3.5

20.9
9.4
28.9
28.9
31.9
33.2

0
0
0
0
0
0

26.9
10.4
34.6
32.2
35.3
36.7

15.2
2.9
14.4
20.9
28.2
29.1

0.6
0.03
0.2
0.1
0.1
0.1

0.4
0
0.2
0.3
0.7
0.6

1.3
1
1.2
1.1
0.9
0.8

0
0.1
0
0.01
0.002
0

0.1
0
0
0.01
0.002
0.01

19

9.4

4.3

13.7

1.4

0.4

1.8

3.9

48.6

52.4

42.5

0.2

1.1

0.7

15

15.2

3.2

18.4

1.6

0.2

1.8

1.1

27.5

28.6

23.9

0.1

0.9

0.7

15.3

3.7

19

0.9

0.1

16

16

1.1

3
1
14

10.2
14.2
9.8

2.5
2.1
3

12.8
16.3
12.8

1.3
0.8
1.2

0.2
0.3
0.3

1.5
1.1
1.5

1.6
3.5
2

21.8
38.2
47.3

0
0
0

23.5
41.7
49.4

19.8
37
39.5

0
0
0.05

2
7
1

9.3
9.4
17.1

2
4
0.2

11.3
13.4
17.3

1.8
1.6
0.2

0.3
0.2
0

2.1
1.8
0.2

1
1.1
0

42
29.1
4.2

0
0
0

43
30.2
4.2

37.4
25.9
4.2

12.8

14.8

0.8

0.2

1.1

1.1

23.5

24.7

16.7

Ca

Tcarb

0.03
0.1

3.9
5.4

0.1
0

0.7
1.4

0.4
0.3

5
7.1

0.4
0
0.1
0.1
0.1
0.03

0.1
0
0.1
0.05
0.02
0.02

19.5
4.3
4.3
9.1
6.1
3.8

0.2
0
0.6
0.3
0.7
2.2

0.6
0.6
9.8
1.1
3.5
4.1

0.2
0.1
3.8
0.9
2.3
1.8

0.6

0.1

0.04

1.6

0.9

2.5

0.8

0.5

0.1

0.03

3.5

1.6

0.3

0.4

5.8

0
0
0.01

0.07
0.6
0.3

0.7
1.3
1

0.3
0.8
0.4

0
0
0

0.03
0
0.02

0
0
0

0
0
0

0.4
0.2
0

1.1
1.2
0.2

0.5
0.3
0.4

0.1
0
0

0.2

0.7

0.4

Q = Quartz, Qm = Monocrystalline quartz, Qp = Polycrystalline quartz, TQ = Total quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar,
TF = Total feldspar, RF = Rock fragment, SRF = Sedimentary rock fragment, VRF = Volcanic rock fragment, MRF = Metamorphic rock
fragment, TRF = Total rock fragment, Ch = Chert, Chl = Chlorite, Mu = Muscovite, HML = Heavy minerals, Ho, Hornblende, R = Rutile, Z =
Zircon, T = Tourmaline, THML = Total heavy minerals, C = Cement, Qo = Quartz overgrowth, I = Iron oxide, Py = Pyrite, Carb = Carbonate
cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, TC = Total cement, Ma = Matrix, TPo =
Total thin section porosity, GSZ = Grain size.

78

Carb
S

TC

Appendix 5-15: X-ray diffraction results of fine-grained samples from the Illawarra Coal Measures.
Samples

LDSS - EAW 156 (14)


LDSS - EAW 156 (15)
LDSS - EAW 30 (30)
LDSS - EAW 30 (32)
LDSS - EDEN 124 (13)
LDSS - EDEN 125 (15)
LDSS - EDEN 125 (16)
LDSS - EDEN 126 (16)
LDSS - EDEN 126 (18)
LDSS - EDEN 127 (15)
BASM - EAW 156 (16)
BASM - EAW 156 (17)
LRSS - Surface 1
LRSS - Surface 2
LRSS - Surface 3
LRSS - Surface 5
LRSS - EAW 156 (18)
LRSS - EAW 156 (19)
LRSS - EAW 30 (33)
LRSS - EAW 30 (34)
LRSS - EAW 30 (35)
LRSS - EDEN 124 (17)
LRSS - EDEN 124 (18)
LRSS - EDEN 124 (19)
LRSS - EDEN 124 (20)
LRSS - EDEN 125 (18)
LRSS - EDEN 125 (37)
LRSS - EDEN 125 (19)
LRSS - EDEN 126 (19)
LRSS - EDEN 126 (20)
LRSS - EDEN 126 (21)
LRSS - EDEN 127 (19)
LRSS - EDEN 127 (20)
UNM2 - EAW 156 (21)
UNM2 - EAW 30 (36)
UNM2 - EAW 30 (51)
UNM2 - EAW 30 (37)
UNM3 - EAW 156 (22)
UNM3 - EAW 30 (39)
UNM3 - EAW 30 (55)
UNM3 - EAW 30 (42)
UNM3 - EAW 30 (43)
UNM3 - EDEN 124 (21)
UNM3 - EDEN 125 (20)
UNM3 - EDEN 126 (22)
UNM3 - EDEN 127 (21)

Height
above
base
87.3
85.4
89
86.5
94.2
91.1
88.1
93
87.5
95.1
87.4
85.4

82
79.4
80.5
79.5
77.4
81
78.9
77.9
76.8
85
82.9
79.2
81.6
80
77.1
78.6
77.4
64.8
71.2
68
66.2
60.4
65.9
61.9
59.9
58.1
59.2
58.6
61.6
59

F
Q

32.1
49.8
42.2
41.4
42.6
25
52.3
40.4
39.6
54
3.4
33.8
49.7
34.8
75.8
34.1
36.6
30.7
34
41.5
33.7
40.2
33.2
45.3
39.6
38.3
46
46
37.8
35.6
34
38.2
33.6
38.9
77.4
33.5
35.3
28.9
44.6
48
40.9
51.3
41.2
42
34.7
29.5

Plag

Cl
K-f

Ab

Lab

Orth

Mic

18.8
8.6
13
14.6
5.6
1.9
3.7
3.2
13.8
12.3
1.6
7.8
11.6
8.5
0
1.1
18
16.9
14.6
26.7
15.3
7.4
9.7
14.8
6.4
10.3
12
13.3
11.9
16.1
13
15.7
13.2
13.1
0
17.6
9.1
3.3
5.3
5.6
4.8
2.2
5.1
7.1
19.7
14.2

0
3.4
2.3
0
0.5
0.6
2
1.5
2.7
0.8
0.5
2.1
0.9
1.9
0.8
1.3
3
3
2.5
0
1.9
0.8
0
0
1.6
1.3
0.7
0
5
2.8
3
0
1.5
3.3
1.3
2.5
3.1
3.5
3.5
4.1
2.7
3
1.1
0.9
1.5
0

0
0
0
0
0
0
0
0
0
0
1.7
0
0
0
6.5
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
4
0
0
0
0
0

1.8
2
2.8
2.5
2.7
3.3
2.7
4.2
3.3
2.8
1.1
3.1
2.1
3.5
4.5
3.4
2.5
2.7
2.2
1.6
1.9
2.9
0.6
1.1
3.7
2.7
2.2
2.6
3.3
2.6
1.3
2.6
1.3
3.3
2.3
2.5
3.5
3.2
5.3
5.3
4.1
6.3
2.8
3.4
2.4
1.3

HML

TF

ill

Ka

Di

Chl

Mix

TCl

20.6
14
18.1
17.1
8.8
5.8
8.4
8.9
19.8
15.9
4.9
13
14.6
13.9
11.8
5.8
23.5
22.6
19.3
28.3
19.1
11.1
10.3
15.9
11.7
14.3
14.9
15.9
20.2
21.5
17.3
18.3
16
19.7
3.6
22.6
15.7
10
14.1
15
15.6
11.5
9
11.4
23.6
15.5

4.6
5.7
9.4
10.6
8
7.2
5.8
8.7
6.4
6.1
7.7
8.7
4
5.1
0
6.6
5.6
3.7
7.2
5.2
7.6
4.8
2.9
3.6
5.1
7.1
5.7
6.5
4.7
5.5
5.5
4.5
2.5
6.1
3.7
3.1
7.1
2
5.7
4.6
3.1
5.3
3.6
5.6
4.7
2.1

8.2
2.8
0.3
0.2
0.4
26.4
0.6
0.2
0.8
0.4
39.2
2.8
12.9
6.9
1
19.3
3.4
0.6
8.4
0.5
8.3
11.2
8.6
8.6
1
5.4
1.1
2
2.3
4.3
1.4
12.6
10.2
4.1
0.2
9.8
2
19.5
1.5
1
2.4
0.6
4.1
1.3
7.2
13.8

1.9
0
0
0
0
2
0.4
1.4
0.5
0
0
0
0.2
0.1
0
0.7
0
0
0
0
0
1.1
0
1
1
1
0.9
0.6
0.5
0.2
0
0.7
0
1
0
0.6
0.2
1.2
0.9
0.1
2
1.9
0.9
0
0
0

1.9
0
1.1
1.3
0.6
2.9
0
1.4
0.9
0
0
0.9
0.3
0.8
0
1.9
0.3
0
0
0.7
0.4
0
0
0
0.3
0.1
0.9
0.3
0
0.3
0
0
0
0.3
0
0
0.2
0
0.4
0.1
0.6
0
0
0.7
0
0

27
18.5
17.6
20
26
13.3
10.8
24.9
22.3
15
40.5
26.9
11.4
23.6
0.2
17.7
23.3
14.9
22.2
16.6
20.6
17.2
9.9
14.4
23.2
19.3
15.6
15.5
21.1
21.5
14.7
13.7
7.7
19.5
8.3
15.1
24.4
31
18.1
15.8
9.2
13.9
21.8
22.5
19.4
12.1

43.6
27
28.4
32.1
35
51.8
17.6
36.6
30.9
21.5
87.4
39.3
28.8
36.5
1.2
46.2
32.6
19.2
37.8
23
36.9
34.3
21.4
27.6
30.6
32.9
24.2
24.9
28.6
31.8
21.6
31.5
20.4
31
12.2
28.6
33.9
53.7
26.6
21.6
17.3
21.7
30.4
30.1
31.3
28

77

Mu

0
0
4.6
2.3
7.1
8.2
2.7
9.2
3.7
1.3
0
7.4
0
7.3
2.7
6.9
1.4
1.4
3.2
0
4.2
7.7
1.4
1.3
9
7.5
3.2
2.4
5
3
2.9
3.5
1
4.3
0
2.4
5.8
0.5
5.8
5.7
8
6.8
5.2
8.9
3.4
2

Carb

He

Ho

Py

THML

Ca

Tcarb

Rock
type

0
0.2
0.5
0.6
0.5
0.6
0.2
0
0.2
0.2
0
0
0.5
0.3
0
0.3
0.1
0.2
0.4
0
0.7
0.1
0
0
0.1
0.5
0.2
0.3
0.2
0
0.4
0.2
0
0.1
0
0.2
0
0
0
0
0
0
0
0
0.3
0

1.4
2.3
2.7
2.3
3
3.6
4.5
3.2
3.4
4.4
0.7
2.7
1.9
3.3
3.7
3.4
2.7
1.5
1.9
1.5
2.5
3.9
2.2
2.5
4.6
2.9
3.8
4.3
3.6
4.1
2.7
3.9
2
2.1
3.4
1.1
3.6
3.2
6.3
5.9
4
5.8
5.5
3.7
2.1
2.3

0
0
0
0
0
0
0
0
0
0
0.7
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

0
0
0.3
0.2
0.1
0.6
0
0.5
0.1
0
0.4
0.4
0
0.2
2.3
0.3
0.1
0.1
0.1
0
0.2
0.2
0
0
0.3
0.1
0
0.1
0.1
0.2
0
0.1
0
0.1
0
0
0.3
0.3
0.5
0.4
0.9
0.7
0
0.3
0
0

1.4
0.6
0.8
0
0.6
2
0
0.3
0.9
0
0.2
1.2
2.1
2.6
0
1.5
0.4
0.8
0.2
0
0.7
0.2
0
0
1.7
1.2
0.9
0.8
0.6
0.5
0
0.7
0.4
1.3
0
0
0.9
0.9
1.1
0.8
0.2
0.6
0.4
1.1
0.2
0.1

0
0.2
0.2
0.2
0.2
0.6
0.1
0.3
0.2
0
0.8
0.1
0
0.1
0
0.5
0.1
0
0.3
0
0.2
0.1
0
0
0.1
0.3
0.1
0.1
0
0
0.2
0.2
0
0
0
0
0
0.3
0.2
0
0.1
0
0
0.2
0.1
0.2

2.8
3.3
4.5
3.3
4.4
7.4
4.8
4.3
4.8
4.6
2.8
4.4
4.5
6.5
6
6
3.4
2.6
2.9
1.5
4.3
4.5
2.2
2.5
6.8
5
5
5.6
4.5
4.8
3.3
5.1
2.4
3.6
3.4
1.3
4.8
4.7
8.1
7.1
5.2
7.1
5.9
5.3
2.7
2.6

0.3
0.7
0.9
1.8
1.4
0.7
0.6
0
0.6
2.5
0.7
1.7
0.2
0.5
0
0.6
2.1
0.9
1.5
1.4
1.5
2
0.1
2.5
1.9
1.6
0.8
1.3
1
2.7
0.4
2.5
0.3
2.1
2
2.7
4.1
1.7
0.5
1.8
7
1.4
6.1
2.1
1
1.5

0.3
0.8
0.7
1.6
0.7
0.6
0.3
0.6
0.6
0.2
0.3
0.3
1.1
0.4
0
0.4
0.2
0.2
1.1
0.8
0.3
0.2
1
0.5
0.4
0.4
0.8
1
0.2
0.6
5.2
0.2
0.2
0.3
0
0.1
0.2
0.1
0.3
0
0.1
0
0
0.2
0.6
0.3

0.3
2.7
0.6
0.4
0
0.3
7.5
0
0
0
0.3
0.1
1.1
0.1
0
0
0
12.1
0.2
3.2
0
0
22.4
3.4
0
0
4.2
2.4
1.5
0
11.1
0.3
16.5
0.1
0.8
5.9
0.2
0
0
0.4
2.8
0
1
0
2
14

0
1.7
0
0
0
0.2
5.8
0
0
0
0.2
0
0
0
2.5
0
0.2
10.3
0
0.3
0
0
8
1
0
0
0.9
0.5
1.2
0
4.2
0.4
9.6
0
0.6
2.9
0
0.4
0
0.4
3.1
0.2
1.2
0
0.7
6.6

0.9
5.9
2.2
3.8
2.1
1.8
14.2
0.6
1.2
2.7
1.5
2.1
2.4
1
2.5
1
2.5
23.5
2.8
5.7
1.8
2.2
31.5
7.4
2.3
2
6.7
5.2
3.9
3.3
20.9
3.4
26.6
2.5
3.4
11.6
4.5
2.2
0.8
2.6
13
1.6
8.3
2.3
4.3
22.4

Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Ss
Sh
Sh
Ss
Sh
Ss
Sh
Ss
Ss
St
Ss
Ss
Sh
Ss
Ss
Sh
St
Ss
Ss
Ss
Ss
Ss
St
Ss
St
St
Ss
Sh
Ss
Sh
Sh
St
Sh
Sh
St
Ss
Ss

Appendix 5-15 (cont.): X-ray diffraction results of fine-grained samples from the Illawarra Coal Measures.

Samples

WWCO - EDEN 124 (22)


KBSS - Surface 2
KBSS - Surface 3
KBSS - Surface 4
KBSS - Surface 9
KBSS - EAW 156 (24)
KBSS - EAW 156 (25)
KBSS - EAW 156 (26)
KBSS - EAW 156 (27)
KBSS - EAW 30 (52)
KBSS - EAW 30 (50)
KBSS - EDEN 125 (21)
KBSS - EDEN 126 (23)
KBSS - EDEN 126 (25)
KBSS - EDEN 127 (22)
KBSS - EDEN 127 (23)
KBSS - EDEN 127 (24)
ACSM - EDEN 125 (24)
ACSM - EDEN 126 (26)
ACSM - EDEN 127 (25)
ACSM - EDEN 127 (27)
ACFM - Surface 1
ACFM - Surface 2
ACFM - Surface 3
ACFM - EDEN 124 (28)
ACFM - EDEN 124 (29)
ACFM - EDEN 125 (25)
ACFM - EDEN 125 (26)
ACFM - EDEN 126 (27)
ACFM - EDEN 126 (28)
ACFM - EDEN 127 (28)
DFSS - EDEN 124 (30)
DFSS - EDEN 124 (31)
DFSS - EDEN 125 (27)
DFSS - EDEN 125 (28)
DFSS - EDEN 126 (36)
DFSS - EDEN 126 (29)
DFSS - EDEN 127 (29)
DFSS - EDEN 127 (30)
BGCS - EDEN 124 (32)
BGCS - EDEN 124 (33)
BGCS - EDEN 125 (29)
BGCS - EDEN 125 (30)
BGCS - EDEN 125 (31)
BGCS - EDEN 125 (32)
BGCS - EDEN 125 (33)
BGCS - EDEN 126 (30)
BGCS - EDEN 126 (31)
BGCS - EDEN 126 (32)
BGCS - EDEN 127 (31)

Height
above
base
46.4

45
43.7
42.1
40.2
43.7
42.5
46.2
45.6
42.3
47
43.6
41
37.3
37.3
39.3
35.5

33.3
31.6
34
30
33.3
28.9
30.5
28.9
24.4
25.8
23.9
25.5
23.7
30.3
24.6
18.5
14.9
24.2
22.1
20.2
16.9
14.2
21.5
16.6
13.6
22.1

F
Q

51.6
45.2
42.9
47.1
58.3
31.3
33.2
52.8
46
30.9
52.5
46.7
43.5
36
32
49.9
39.3
41.5
44.2
31.2
24
43.2
50.6
47.8
41
38.6
49.4
42.1
13.8
45.1
27.4
30.5
42.4
50.2
34.6
43.7
43.4
27.9
46.7
13.3
41.3
36.9
30.5
36.1
26.2
34.4
30.7
7.9
37
42

Plag

Cl
K-f

Ab

Lab

Orth

Mic

5.8
0.8
2.6
0.4
2.7
10.5
18.2
3.1
9.7
15.1
13.6
7.8
8.5
10
9.4
11
12
0.8
16.3
3.1
9.1
10.9
3.4
6.1
11.3
11.4
4.7
10.3
2.8
11.9
3.7
12.4
10.3
9.8
16.3
16.5
12.9
12
16.2
2.1
9
8.5
13.8
7.9
0.9
9
9.4
10
11.4
12

0.4
1.9
1.4
2.3
3.5
0.9
0
0
1.4
2.9
2.8
3.1
2.2
1.8
1.6
0.6
1.7
1.1
1.6
1.2
1.5
1.6
2.4
2.6
1.3
1.7
1.8
2.7
0.4
2.1
0.4
2
0.3
1.5
1.4
1.3
2.4
3.2
0.9
0
1.3
2.5
1.9
0.4
1.3
1.8
1.5
0
1.5
1.4

0
0
0
0
0
0
0
0
0
3.1
0
0
0.6
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1.1
0
0.5
0
0
0
0
0
0
0
0
1.4
0
0
0
0
0
0
0
1.3
0
0

2.4
5.1
1.8
3.9
5.9
2.9
1.5
0.7
1.6
4.1
2.3
4.7
3.1
3.1
1.4
1.6
2.3
2.8
1.4
2.7
2.4
2.7
4.3
2.7
1.4
3.1
1.7
3.3
1.9
2.5
2.9
1.1
2.8
3.8
1.8
2.1
2.2
0.6
1.5
0.2
3.1
3.5
1.7
2.6
3.2
3.9
3.5
0.9
3.5
3.1

HML

TF

ill

Ka

Di

Chl

Mix

TCl

8.6
7.8
5.8
6.6
12.1
14.3
19.7
3.8
12.7
25.2
18.7
15.6
14.4
14.9
12.4
13.2
16
4.7
19.3
7
13
15.2
10.1
11.4
14
16.2
8.2
16.3
6.2
16.5
7.5
15.5
13.4
15.1
19.5
19.9
17.5
15.8
18.6
3.7
13.4
14.5
17.4
10.9
5.4
14.7
14.4
12.2
16.4
16.5

3.5
3.4
2.1
3.7
2.7
4.9
4.2
0.6
4
8.9
8.8
2.5
4
3.7
5.4
2.7
3.9
1
3.5
2.5
4.6
2.7
4.2
4.6
4.3
6.3
2.5
5.9
4.9
3.5
2.6
5.7
7
4.3
3.7
4.2
2
4.9
3.8
7.3
6.4
8.5
5.1
8.9
9.6
9.8
8.6
9.1
7.2
6.3

11.9
11.2
23
12.9
5.5
8.4
16
29.4
10.4
0.2
0.2
5.9
5.1
15.8
13.5
18.8
11.5
32.6
4.4
26.4
17.1
20
7.3
13.7
2.1
0.5
21.5
0.2
40.2
6.5
25
0.3
0.2
6.7
0.2
4
9.9
1.3
2.1
41.8
1
0.9
0.5
0.2
17.5
0.2
1.2
29.5
1.1
0.2

0.7
1.1
1.3
1.6
1
0
0
0.1
1.2
0.4
0
3.3
1.8
0
0
0
0
2.2
1.5
2.2
0
0
0.8
0.2
0.4
0.6
0.7
0.5
0
2.1
0
0
0.1
0.8
0.4
0.5
0.1
0.8
0
0
0
0.7
0.1
0.3
1.5
0.8
0
0
0
0.5

0.3
0.3
0
0
0
1.3
0.8
0
0.1
1.3
0
1.1
0.8
0
0
0
0
0
0.4
0
0
0.5
0.5
0
0
1.5
0
1.2
0.4
0.1
0
0
0.9
0
0.2
0
0
0
0
0
0
0.9
0.5
1.3
0
0.8
2.3
0
2.6
1.2

16.4
19
7.8
11.8
7.8
26.1
19.2
9.8
14.8
23.3
14.2
13.1
15.2
21.7
13.6
6.9
21.7
9.8
12
22.7
29
6.8
13.7
9.5
9.8
25.2
8.7
18.5
23.3
15.9
22.2
13.1
20.3
12
12.7
15
9.7
8.8
12.2
25.9
23.9
21.6
14.6
21.7
28.7
23
27.2
37.4
21
19.1

32.8
35
34.2
30
17
40.7
40.2
39.9
30.5
34.1
23.2
25.9
26.9
41.2
32.5
28.4
37.1
45.6
21.8
53.8
50.7
30
26.5
28
16.6
34.1
33.4
26.3
68.8
28.1
49.8
19.1
28.5
23.8
17.2
23.7
21.7
15.8
18.1
75
31.3
32.6
20.8
32.4
57.3
34.6
39.3
76
31.9
27.3

77

Mu

0
4.2
0
4.3
4.3
5.2
1.2
0
0
5.4
2.8
1
2.5
1.4
1.4
0
2.6
0
0
0.5
4.2
4.1
5
4.3
1.2
6.6
0
7.6
7.2
0
7.7
2.2
9.3
4.3
0.9
0
0
0.9
2.2
0.4
7.7
8.7
1.7
10.6
4.1
8.9
11.1
0
9.5
9.4

Carb

He

Ho

Py

THML

Ca

Tcarb

Rock
type

0
0
0
0
0
0.1
0
0.1
0.1
0
0.3
0
0.1
0
0
0
0
0
0
0.2
0
0.3
0
1.1
0
0.1
0
0.1
0.3
0.3
0
0
0.1
0
0.2
0
0
0
0
0
0.1
0.3
0
0
0
0
0.4
0
0.2
0.1

2.9
5.7
3.9
5.8
6.3
1.2
2.7
0.1
2.6
1.2
0.8
3.6
3.8
4
1.8
2.6
2.9
3.3
1.5
3.8
3.1
4.3
6.5
4.6
3
1.7
4.6
4.6
0.6
3.2
2.3
4.6
1.9
4.3
2.7
2.9
3.7
4.8
2.6
0
3
3
3.8
3.5
3
4.1
1.3
0
1.8
2.1

0
0
0
0
0
0
0
0.2
0
0
0.2
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0.4
0
0

0
0.3
0
0.3
0.2
0.2
0.1
0
0
0.3
0
0.2
0.2
0.2
0
0
0
0.1
0
0.1
0.3
0
0.2
0
0
0.7
0
0.3
0.3
0
0.2
0
0.2
0.3
0
0
0
0.1
0
0.2
0.2
0.3
0
0.5
0.6
0.5
0.5
0.4
0.3
0.3

2.4
1.6
1.6
2.3
1.7
1.5
0.9
1.7
1.4
1.1
0
6.7
1.2
0.8
0.3
0
0.1
1.1
1.6
1.2
0.3
2.1
1.1
1.7
0
1.2
0
1.4
1.1
3.1
0
0
0.6
1.6
0.9
0.5
0.6
0.7
0
0
0
1.7
1.3
1
0
1.2
0.8
0
0.8
1.2

0
0.2
0
0.3
0.1
0.4
0
0.4
0.2
0.2
0
0
0.3
0.1
0.2
0
0.1
0.1
0.1
0.4
0.5
0.4
0
0.4
0
0.3
0
0.1
0.7
0.1
0.2
0
0
0
0
0
0
0.5
0
0.9
0.1
0.2
0
0.3
0.2
0.2
0.5
0.3
0.3
0

5.3
7.8
5.5
8.7
8.3
3.4
3.7
2.5
4.3
2.8
1.3
10.5
5.6
5.1
2.3
2.6
3.1
4.6
3.2
5.7
4.2
7.1
7.8
7.8
3
4
4.6
6.5
3
6.7
2.7
4.6
2.8
6.2
3.8
3.4
4.3
6.1
2.6
1.1
3.4
5.5
5.1
5.3
3.8
6
3.5
1.1
3.4
3.7

0.9
0
0
0.2
0
3.4
1.6
0.5
1.4
1.1
1.5
0.3
0.6
1.2
1.4
0.3
1.3
1.3
1.3
1.2
3.9
0
0
0.2
0.2
0.1
2.2
0.9
0.3
0.4
4.8
1.3
3.4
0.4
1
1.2
0.2
1.8
0.8
1.4
2.4
1.5
2
4.2
3.2
0.9
0.3
0.9
1.2
0.8

0
0
10
1.8
0
0.2
0.3
0
0.4
0.5
0
0
0.3
0.2
0.4
0
0.4
0
0.2
0.1
0
0
0
0.3
6.8
0.4
0
0.3
0.4
0.1
0
17.2
0.2
0
19.5
0
1.8
1.2
0.2
1.6
0.5
0.3
15.5
0.4
0
0.5
0.5
1.6
0.5
0.3

0.7
0
1.1
0.9
0
0.1
0.1
0.3
3.3
0
0
0
4.7
0
14.5
5
0.2
0.1
6
0.1
0
0.3
0
0.1
14.9
0
1.7
0
0.1
2.4
0
9
0
0
2.7
6.5
7.9
25.9
8.3
2.6
0
0
5.6
0.1
0
0
0.2
0.2
0.1
0

0.1
0
0.5
0.4
0
1.4
0
0.2
1.4
0
0
0
1.5
0
3.1
0.6
0
2.2
4
0.4
0
0.1
0
0.1
2.3
0
0.5
0
0.2
0.7
0.1
0.6
0
0
0.8
1.6
3.2
4.6
2.5
0.9
0
0
1.4
0
0
0
0
0.1
0
0

1.7
0
11.6
3.3
0
5.1
2
1
6.5
1.6
1.5
0.3
7.1
1.4
19.4
5.9
1.9
3.6
11.5
1.8
3.9
0.4
0
0.7
24.2
0.5
4.4
1.2
1
3.6
4.9
28.1
3.6
0.4
24
9.3
13.1
33.5
11.8
6.5
2.9
1.8
24.5
4.7
3.2
1.4
1
2.8
1.8
1.1

Ss
Sh
Ss
Sh
St
Sh
Ss
Ss
Ss
St
St
Ss
Ss
Ss
Ss
Ss
Ss
St
Sh
St
Sh
St
St
Ss
Ss
Sh
St
St
Sh
Ss
Sh
Ss
Sh
Ss
Ss
Ss
Ss
Ss
Ss
Sh
Sh
Sh
St
Sh
Sh
Sh
Sh
Sh
St
Sh

Appendix 5-15 (cont.): X-ray diffraction results of fine-grained samples from the Illawarra Coal Measures.
Samples

BGCS - EDEN 127 (32)


BGCS - EDEN 127 (33)
TGSM - EDEN 124 (34)
TGSM - EDEN 125 (34)
TGSM - EDEN 125 (38)
TGSM - EDEN 126 (33)
TGSM - EDEN 127 (34)
WTFM - EDEN 124 (35)
WTFM - EDEN 124 (36)
WTFM - EDEN 125 (35)
WTFM - EDEN 126 (34)
WTFM - EDEN 126 (37)
WTFM - EDEN 126 (35)
WTFM - EDEN 127 (35)
WTFM - EDEN 127 (36)

Height
above
base

F
Q

Plag
Ab

Cl
K-f

Lab

Orth

Mic

HML

TF

ill

Ka

Di

Chl

Mix

TCl

Mu

He

Carb

Ho

Py

THML

Ca

Tcarb

Rock
type

17.9

3.2

0.1

1.1

1.4

68.7

24.8

94.9

0.2

0.2

0.3

0.3

0.6

Sh

13.3
8.8
9.1
6.9
8.3
9.4
6.1
2.5
1
4.1
2.5
0.005
4.3
1.6

49.6
36.5
34.3
39.9
30.4
37.9
34.7
3.9
2.7
39.9
34.3
1.7
40.2
33.1

11.1
1.8
3.4
6
5.1
4.6
7.1
0
1.4
6.8
5.6
0
9.4
4.7

1.6
0.6
1.5
1.7
2.1
0.5
0.4
2.5
0.6
2
1.7
4.6
0
1

0
0
0
0
0
0
0
1
1.4
0
0
2
0
0

1.8
3.1
2.9
2.8
3.7
3.1
1.7
0.1
0.9
2.8
4.6
1.5
0.7
3.8

14.5
5.5
7.8
10.5
10.9
8.2
9.2
3.6
4.3
11.6
11.9
8.1
10.1
9.5

4.1
5
7.9
7.2
7.7
5
0.6
0.4
7.1
3
4.3
7
0.7
4

4.1
12.9
8.9
5.2
12.7
16.7
29.3
13.8
44.9
15.9
12.2
9.1
27.5
14.9

0
0
0.5
1.4
0.8
0.6
0
2
1.9
1.6
1
2.6
0
1.5

0
0
0.2
0
0.8
0.3
0
0
0
0.7
1.8
0
0
0.9

16.7
28.6
29.3
22.5
23.2
22.2
12.9
7.3
36.8
12.7
17
24.7
8.3
16.8

24.9
46.5
46.8
36.3
45.2
44.8
42.8
23.5
90.7
33.9
36.3
43.4
36.5
38.1

0.8
5.7
4.5
5.4
5.9
0.3
0
0
0
2.8
8.9
1.5
0
7.8

0
0.2
0
0
0.3
0.3
0
0.1
0
0.3
0
0
0
0

0.7
3.1
3.1
3.2
2.3
2.3
1.5
2.3
0
4
3.7
2.6
0.2
2.7

0
0
0
0
0
0
0
0
0.5
0
0
0
0
0

0
0.1
0.2
0.2
0.3
0
0
0
0.4
0
0.4
0.6
0
0.5

0
0.7
1.2
0.2
1.8
1
1.3
1.8
0
1.3
1.2
2.1
0
1

0
0.2
0.3
0.3
0.5
0.5
0
0
0.6
0.3
0.2
0
0
0.3

0.7
4.3
4.8
3.9
5.2
4.1
2.8
4.2
1.5
5.9
5.5
5.3
0.2
4.5

4
0.5
1.4
3.3
0.8
1.9
0.5
14.9
0.6
2.4
3
18.3
0
5.8

0
1
0.4
0.7
1.2
1.6
2.4
41.5
0.1
2.3
0.1
18.5
4.2
0.3

4.3
0
0
0
0.4
0.9
4.6
5.4
0.1
0.7
0
2
5.8
0.5

1.2
0
0
0
0
0.3
3
3
0
0.5
0
1.2
3
0.4

9.5
1.5
1.8
4
2.4
4.7
10.5
64.8
0.8
5.9
3.1
40
13
7

St
Sh
Sh
Sh
St
Sh
Ss
Sh
Sh
St
Sh
Sh
Ss
Sh

Q = Quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar, Ab = Albite, Lab = Labradorite, Orth = Orthoclase, Mic = Microcline, TF = Total
feldspar, Cl = Clay minerals, ill = illite, Ka = Kaolinite, Di = Dickite, Chl = Chlorite, Mix = Mixed layer illite/smectite, TCl = Total clay minerals, Mu =
Muscovite, HML = Heavy minerals, He = Hematite, Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, Py = Pyrite, THML = Total heavy
minerals, Carb = Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, Ss = Sandstone, St =
Siltstone, Sh = Shale.

77

Appendix 5-16: X-ray diffraction results of fine-grained samples from the Narrabeen Group.
F

Samples

Height
above
base

NPFM - EAW 42 (2)


NPFM - EAW 42 (3)
NPFM - EAW 42 (4)
NPFM - EAW 42 (5)
NPFM - EAW 42 (6)
GRFM - EAW 42 (7)
GRFM - EAW 42 (8)
BACS - EAW 42 (9)
BACS - EAW 42 (10)
BACS - EAW 42 (11)
BACS - EAW 42 (12)
BACS - EAW 42 (13)
BACS - EAW 42 (15)
BACS - EAW 42 (16)
BACS - EAW 42 (17)
BGSS - EAW 42 (19)
BGSS - EAW 42 (20)
BGSS - EAW 42 (21)
BGSS - EAW 42 (22)
BGSS - EAW 42 (23)
BGSS - EAW 42 (24)
BGSS - EAW 42 (26)
BGSS - EAW 42 (27)
BGSS - EAW 42 (28)
BGSS - EAW 42 (29)
BGSS - EAW 42 (33)
SPCS - EAW 42 (34)
SPCS - EAW 42 (36)
SBSS - EAW 42 (37)
SBSS - EAW 42 (39)
SBSS - EAW 42 (40)
SBSS - EAW 42 (41)
SBSS - EAW 42 (+42)
SBSS - EAW 42 (43)
SBSS - EAW 42 (45)
SBSS - EAW 42 (46)
SBSS - EAW 30 (2)
SBSS - EAW 30 (3)
SBSS - EAW 30 (5)
SBSS - EAW 30 (6)
SBSS - EAW 30 (8)
SBSS - EDEN 124 (1)
WBCS - EAW 156 (2)
WBCS - EAW 156 (3)
WBCS - EAW 156 (4)
WBCS - EAW 156 (5)

365.2
363.2
361.2
357.4
354
352.2
351.5
356.8
353.5
349.6
346.6
343.6
337.3
334.6
330.8
392.1
389.9
385.5
379.4
330.7
323.3
312.8
309.6
249
246.6
194.6
183.5
178.1
179.4
174
172.6
168.6
159
156.8
148.1
143
155.8
153
149.3
147
141.7
143.3
134.6
132.6
128
124.6

83.2
32.8
65.9
38.1
53.5
0.6
0.6
0.7
0.8
0.8
0.5
7.3
0.7
13.2
17.3
78.6
12.9
80
71.8
82.9
78.8
12.5
73.7
15.6
82.6
39.1
53.3
37.7
55.2
59.4
25.3
34.9
47.2
54.4
57
29.6
57.8
26.2
49.8
60.4
29.3
33
21
29
30.3
49.8

Plag
Ab
Lab
0
0.5
0.2
0.6
1.2
1.6
2.1
2.9
3.1
2.7
2.5
2.9
4.2
2
1
0
0
0
0
0
0
0
0
0.7
0
0.9
0.2
1.2
0.2
0
0.9
0.8
0.9
0
0.2
0.9
0
1.2
1
0
1.3
1.3
7.9
3.8
0.7
8.1

0.8
0.2
0.7
0
0.3
0.6
1.2
0.8
1.3
1.9
1.1
0.7
0.6
0.6
0
1.2
0
1.1
1.8
1
1
0
1.7
0
0.8
0.6
2.4
0.9
1.8
2.4
1.7
1.3
1.5
2.3
1.8
1
2.2
0.7
1.2
2.2
1.5
1.1
3.5
2.4
1.6
0

Cl

K-f
Orth
Mic
0
0
0
0
0
0
0
0.1
0.3
0.6
0
0
0.2
0.1
1.5
0
1.2
0
0
0
0
1.8
0
0.4
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
1.2
0
0
0

0.9
3
0
2
0.4
1
3.2
3.5
3.5
3.6
3
2.9
3.6
3
0.5
0.3
2
0.4
2.2
1.8
0.4
2.2
1.7
3.1
0.9
2.8
2.2
3.8
3
3.4
4.1
3.4
2.5
4.1
2
3.5
2.7
3.6
3.2
3.3
3.8
3.4
0.4
2.3
3.9
1.5

HML

TF

ill

Ka

Di

Chl

Mix

TCl

1.7
3.7
0.9
2.6
1.9
3.2
6.5
7.3
8.2
8.8
6.6
6.5
8.6
5.7
3
1.5
3.2
1.5
4
2.8
1.4
4
3.4
4.2
1.7
4.3
4.8
5.9
5
5.8
6.7
5.5
4.9
6.4
4
5.4
4.9
5.5
5.4
5.5
6.6
5.8
13
8.5
6.2
9.6

1.1
2.5
3.7
2
2.9
1.4
3.2
2.9
4.6
3.6
4.1
2.9
4.3
3.8
0
2.6
0
2.6
3.1
1.1
1.7
0
2.8
6.4
2.5
8.6
6
9.3
5.7
5.3
10.2
9.8
7.2
4.8
5.9
11.9
5.4
8.1
6.5
5.4
9.7
5.8
0.8
7.9
9.2
0

2.8
22.7
9
24.3
27.4
61.3
56.7
43.1
51.7
41.6
34.8
34.6
53.4
46.3
39.7
2.2
11.3
3
4.1
0.8
0.2
11.7
5.9
14.9
2.3
4.9
1.8
4.6
2.3
0.9
2.5
3.2
5.3
2.4
8.8
3.2
0.2
12.2
0.6
0.2
7.9
13.6
5.9
16.2
5.6
27.4

1.8
0
0.1
0.4
0.4
0.8
2.6
5.7
3.6
5.8
4.6
4.1
4.3
1.4
0
2.7
0
2
3.2
1.5
4.1
0
3.9
0.2
0.3
1.2
1.1
0.9
1.8
1.6
0.2
0.6
1.2
1
1.1
0
1.4
1.3
0.9
0.9
1.6
1.2
1.6
0.3
0.9
0

0
0.2
0
0
0.3
1.5
2.9
4.2
5.5
5.5
2.3
2.6
5.7
1.9
1.2
0
0
0
0
0
0.7
0
0
0.4
0
0
0.4
0.7
0.1
1.2
1.1
0.2
0.2
0.4
0
0.4
1.3
0.8
1.2
1.1
2.3
0.7
7.3
3.6
3.6
0

3.1
10.9
6.1
9.1
2.5
4.1
5
5.8
3.1
3.6
6.9
6.3
4
7.6
6.7
5
23.6
3.1
5.9
3
4.1
23.7
3.6
21.6
4.2
19.7
13.2
18.8
13.9
11
27.8
22.4
19.6
12.9
9.9
26.5
12.7
23
16.5
13.4
22.6
16.4
12.9
20.7
25
9.2

8.8
36.3
18.9
35.8
33.5
69.1
70.4
61.7
68.5
60.1
52.7
50.5
71.7
61
47.6
12.5
34.9
10.7
16.3
6.4
10.8
35.4
16.2
43.5
9.3
34.4
22.5
34.3
23.8
20
41.8
36.2
33.5
21.5
25.7
42
21
45.4
25.7
21
44.1
37.7
28.5
48.7
44.3
36.6

77

Mu

0
18.6
4.2
14.3
2.1
1.6
3.3
3.4
2.6
2.7
4.4
5.9
1.9
5.8
22
0
44.1
0
0
0
0
40.2
0
28.2
0
11.1
0
10.4
1.2
0.9
11.4
8.8
1.8
0.8
0
12.1
0
10.6
6.4
1.1
9.1
10.1
6.7
3.9
9.5
0

Carb

He

Ho

Py

THML

Ca

Tcarb

Rock
type

0
0.2
0.3
0.6
0.2
0.8
1.4
15.2
5.8
15
18.1
15.7
4.4
1.1
0.4
0
0.4
0
0
0
0
0.4
0
0.4
0
0.5
0.4
0.4
0.3
0.3
1.4
0.2
0.4
0.1
0.3
0.5
0.2
0.3
0.5
0.3
0.5
0.3
0
0.4
0
0

3.6
1.6
2.1
1.8
2.4
3.8
3.5
1.4
1.8
1.2
4.6
2.9
0.4
3.1
0
3.4
0
4
4
4.8
2.8
0
4
1.1
3.1
4
5
4.9
4.5
5.8
3.5
5.6
4.8
6
4.5
3.7
4.7
2.9
3
3.9
3.5
4.7
2.6
3.4
4.3
3.3

0.4
0
0.5
0.2
0.4
0.4
0.2
0.2
0.1
0.6
0
0.2
0
0
0.4
0.6
0.6
0.4
0
0
0.5
0.3
0
0.2
0.3
0.1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0.2
0
0

0
0.5
0
0.2
0
0.8
1.1
0.9
1.3
1
1.3
1.1
1.2
0.8
0.5
0
0.9
0
0
0
0
0.9
0
0.9
0
0.5
0
0.6
0
0
0.7
0.7
0.1
0
0
0.6
0
0.5
0
0
0.8
0.4
0.2
0.5
0.7
0

0.6
1.5
0
1.5
0.8
3
4.9
5.6
5.6
6.5
5.6
5
5.9
4.1
1.3
0.4
0
0
1.5
1
1
0
0
1.8
0
0.9
0.8
1.6
2
2.4
1.9
1.4
1
0.8
0
0.9
1.5
2.1
1.1
1.7
2.7
2.7
1.4
1.7
2.6
0.1

0
0.5
0.3
0.5
0.2
0.6
0.6
1.6
1.3
1.8
2.7
2.6
1.3
0.3
0.7
0
0.8
0
0
0
0
0.7
0.1
0.9
0
0.5
0.2
0.6
0.4
0.1
1.4
0.8
0.4
0.2
0.2
0.8
0
0.4
0.3
0.1
0.6
0.3
0
0.3
0.6
0

4.6
4.3
3.2
4.8
4
9.4
11.7
24.9
15.9
26.1
32.3
27.5
13.2
9.4
3.3
4.4
2.7
4.4
5.5
5.8
4.3
2.3
4.1
5.3
3.4
6.5
6.4
8.1
7.2
8.6
8.9
8.7
6.7
7.1
5
6.5
6.4
6.2
4.9
6
8.1
8.4
4.2
6.5
8.2
3.4

0.3
3.5
6.4
3.2
4.7
14.9
6.5
0.4
2.7
0.2
1.9
0.5
2.8
4.3
5.1
2.8
0
3.4
2.1
2
4.4
3.5
1.7
0.9
2.5
3.3
3.3
2.3
3.8
3.6
4.1
4.6
4.4
4.4
8.2
2.6
9.2
5.6
7
4.9
2
4.3
26
3
1
0.4

0
0.2
0
0.2
0
0.3
0.5
0.3
0.5
0.4
0.5
0.4
0.5
0.2
0.4
0
0.7
0
0
0
0
0.7
0
1
0
0.7
0.4
0.6
0.2
0
0.9
0.8
0.4
0.3
0
0.8
0
0.5
0.4
0.5
0.6
0.5
0
0.3
0.5
0

1.2
0.4
0.3
0.6
0.3
0.5
0.5
1.2
0.8
0.7
0.8
1.1
0.6
0.4
0.9
0.2
0.7
0
0.3
0.1
0.3
0.6
0.7
1
0.3
0.6
5.8
0.7
2.3
0.9
0.9
0.5
1
2.5
0
1
0.4
0
0.2
0.1
0.2
0
0
0.1
0
0.1

0.2
0.2
0.2
0.4
0
0.4
0
0.1
0
0.2
0.3
0.3
0
0
0.4
0
0.8
0
0
0
0
0.8
0.2
0.3
0.2
0
3.5
0
1.3
0.8
0
0
0.1
2.6
0.1
0
0.3
0
0.2
0.5
0
0.2
0.6
0
0
0.1

1.7
4.3
6.9
4.4
5
16.1
7.5
2
4
1.5
3.5
2.3
3.9
4.9
6.8
3
2.2
3.4
2.4
2.1
4.7
5.6
2.6
3.2
3
4.6
13
3.6
7.6
5.3
5.9
5.9
5.9
9.8
8.3
4.4
9.9
6.1
7.8
6
2.8
5
26.6
3.4
1.5
0.6

Ss
Sh
Ss
St
Ss
Sh
Sh
Sh
Sh
Sh
Sh
Sh
Sh
St
Sh
Ss
Sh
Ss
Ss
Ss
Ss
Sh
Ss
Sh
Ss
Sh
Ss
Sh
Ss
Ss
Sh
Sh
Ss
Ss
Ss
Sh
Ss
Sh
Ss
Ss
Sh
St
Sh
Sh
Sh
Ss

Appendix 5-16 (cont.): X-ray diffraction results of fine-grained samples from the Narrabeen Group.
Samples

WBCS - EAW 156 (7)


WBCS - EAW 156 (8)
WBCS - EAW 156 (9)
WBCS - EAW 156 (10)
WBCS - EAW 42 (47)
WBCS - EAW 42 (+48)
WBCS - EAW 42 (49)
WBCS - EAW 30 (9)
WBCS - EAW 30 (10)
WBCS - EAW 30 (12)
WBCS - EAW 30 (13)
WBCS - EAW 30 (14)
WBCS - EAW 30 (15)
WBCS - EAW 30 (16)
WBCS - EAW 30 (19)
WBCS - EAW 30 (21)
WBCS - EDEN 124 (2)
WBCS - EDEN 125 (5)
WBCS - EDEN 125 (36)
WBCS - EDEN 125 (6)
WBCS - EDEN 125 (7)
WBCS - EDEN 125 (9)
WBCS - EDEN 125 (10)
WBCS - EDEN 126 (1)
WBCS - EDEN 126 (2)
WBCS - EDEN 126 (38)
WBCS - EDEN 126 (3)
WBCS - EDEN 126 (5)
WBCS - EDEN 126 (8)
WBCS - EDEN 126 (9)
WBCS - EDEN 126 (10)
WBCS - EDEN 126 (11)
WBCS - EDEN 126 (12)
WBCS - EDEN 127 (1)
WBCS - EDEN 127 (2)
WBCS - EDEN 127 (4)
WBCS - EDEN 127 (6)
WBCS - EDEN 127 (8)
WBCS - EDEN 127 (9)
WBCS - EDEN 127 (10)
CCSS - Surface 3
CCSS - EAW 156 (11)
CCSS - EAW 156 (12)
CCSS - EAW 156 (13)
CCSS - EAW 30 (23)
CCSS - EAW 30 (26)
CCSS - EAW 30 (54)

Height
above
base
120.5
115.6
112.1
111.2
144.4
139.6
138.6
150.2
147.3
143.6
140.6
138.1
134
129.4
121.9
115.9
126.1
129.4
126.8
122.7
120.7
116
113.9
144.2
140.1
137.8
136.1
131.6
124.1
121.6
119.2
115.1
113.8
137.1
134.2
128.9
124.7
118.8
115.7
112.4
100.8
97.4
96.3
112.8
102.2
101.1

F
Q

27.5
26.9
0.4
30.1
55.9
66.1
45.5
36.6
55.3
68.5
46.6
76.3
71.4
60.6
40
41.1
56.4
35.6
52
55.6
67.5
37.3
61.4
23.6
40.5
62
61.6
39
51
34.4
36.3
34.9
58.4
28.2
53.8
30.2
40
47.5
34.6
32.4
59.2
47.6
32.6
57.8
53.2
70.8
67.1

Plag

Cl
K-f

Ab

Lab

Orth

Mic

6.8
2.7
2.8
0
0.3
0.2
1.3
0
0.1
0
0
0
0
0.4
0.9
0.2
0.1
0.9
0.2
0.2
0
0.7
0
0.8
0
0
0
0.2
2.1
0.2
4.9
0.5
0.9
1.1
1.1
3
1.9
5.9
1.4
6.4
4.7
0.8
22.2
4.3
0.5
0
0

1.6
1.3
3.9
2.6
0.9
1
0.6
0.6
2.5
1.8
1.2
1.3
2.1
2.3
0.9
1.1
1.2
0.9
1
2
1.3
0.8
1.9
1.3
1.1
1.8
2
0.2
0.8
2.4
0.5
1.5
2.2
1.4
1.3
1.3
0.5
1
1.1
1.2
0.6
0.9
0
0.4
1.9
2.2
1.7

0
0
0.8
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

3.1
3.7
4.9
3.3
2
0.4
2.1
3
4
3
3.2
1.2
2.4
3.1
2.5
5.3
1.6
3.4
2.1
3
1.4
3
2.6
3.4
3
1.9
2.3
2.9
2.3
3.2
2.3
3.6
1.6
3.7
2.7
2
2.8
2.3
3.5
1.3
1
1.7
1.6
0.8
3
2.7
2.4

HML

TF

ill

Ka

Di

Chl

Mix

TCl

11.5
7.7
12.4
5.9
3.2
1.6
4
3.6
6.6
4.8
4.4
2.5
4.5
5.8
4.3
6.6
2.9
5.2
3.3
5.2
2.7
4.5
4.5
5.5
4.1
3.7
4.3
3.3
5.2
5.8
7.7
5.6
4.7
6.2
5.1
6.3
5.2
9.2
6
8.9
6.3
3.4
23.8
5.5
5.4
4.9
4.1

6.6
7.1
8.4
7.4
3.5
4.2
4.3
2.2
6
5.1
2.5
5.4
5
7
10
4.5
4.6
9.1
4.4
4.1
3.7
5.7
4.9
8.6
1.6
4.9
5
1.8
3.8
6.9
1.6
5.6
5.6
9.4
4.8
2.4
2.4
7.9
6.6
4.2
2.9
2.3
3.1
1.7
9.7
6.7
5.5

17.8
10.8
37.3
5.9
11.6
9.3
17.7
3
3.2
0.2
0.2
0.4
0.2
0.2
0.3
0.2
8.5
6.6
7
4.1
2.9
11.2
3.1
2.8
8.4
2.4
1.6
4.4
11.4
1.5
0.3
3
0.2
4.1
9.6
0.2
6.3
4.2
0.2
4.2
18.2
26.6
11.5
24.4
0.2
0.2
0.2

0.8
0.7
3.3
2.7
0.6
0.5
0.7
1.4
1.5
1.5
1.4
0.2
1.8
1.9
0.6
1.4
0.9
1.5
0.7
1.8
1.1
1.7
2.1
1.2
2
2.4
2.5
0.8
1
1.3
2.7
1.8
0.8
0.9
1
2.5
0.7
0.7
1.1
1.2
0
1.6
0
1.2
1
0.9
1.5

1.7
2.2
1.2
2.2
0
0
0
2.4
0.9
1.1
2.2
0
0.7
1.5
1.9
2.5
0.1
1.6
0
0
0
1.6
0
2
1.9
0.1
0.3
1.3
0
1.6
2.4
1.3
0
1.1
0.3
0.5
0.5
0.1
2.5
0.5
0
0
1.5
0
1.3
0.4
1.4

19.8
28.3
4.3
23.6
11.8
9.2
13.3
21.2
14.6
12.3
16
7.7
7.8
11.6
19.5
19.1
14.7
20.9
16.7
14.7
10.5
16.9
11.9
29.6
18.2
13.2
10.1
18.8
17.1
20.2
19.4
26.2
14.9
27.8
15.5
9.4
17.3
18.9
25.4
12.3
7.5
10.6
16
7
14.9
11.1
13.7

46.7
49.1
54.5
41.8
27.5
23.2
36
30.2
26.2
20.2
22.3
13.7
15.5
22.2
32.3
27.7
28.8
39.7
28.8
24.7
18.2
37.1
22
44.2
32.1
23
19.5
27.1
33.3
31.5
26.4
37.9
21.5
43.3
31.2
15
27.2
31.8
35.8
22.4
28.6
41.1
32.1
34.3
27.1
19.3
22.3

77

Mu

6.3
7.4
6.4
9.9
5.5
0
4.5
16
0
0
16.8
0
0
0.4
9.8
16.7
1.2
10.5
8.9
1.8
0
11.1
0
7.6
12.6
0
0
23.3
0
10.4
7.2
15.3
1.3
7.8
2.9
0.8
13.9
4.7
14.9
3.7
2.9
3.8
7.2
0
2.9
0.1
1.4

Carb

He

Ho

Py

THML

Ca

Tcarb

Rock
type

0.6
0.2
0.9
0
0.1
0.1
0.4
0.2
0
0
0
0.2
0.1
0.2
0.5
0.2
0
0.2
0.4
0.2
0
0.2
0.1
0.5
0.1
0.1
0.1
0.5
0
0.3
0
0.3
0
0.2
0.1
0
0.3
0.4
0
0
0.3
0
0
0
0.4
0.2
0

2.4
2.1
4
2.8
4
3.9
4.1
0.5
5.8
4
3.6
3.8
5.9
5
2.1
1.6
3.5
3.1
2.8
5
2.5
3
4.5
2.3
2
4.3
5.1
0
4.3
4.9
2.6
2
2.9
4.4
2.7
3.1
0.7
3.6
2.3
3.6
1.8
1.1
0.7
0.7
4.9
4
3.9

0
0
0
0
0.2
0.4
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0.2
0
0
0
0

0.3
0.4
2.6
0.5
0
0
0
0.2
0
0
0.1
0
0
0
0.4
0.4
0
0.3
0
0
0
0.3
0
0.4
0
0
0
0.2
0
0.4
0
0.4
0
0.6
0
0
0.1
0
0.6
0
0
0
0.2
0
0.2
0
0

1.5
1.9
7
1.9
0
0
0.3
0.6
1.5
0.3
1.3
0
1.6
1.8
0.4
1.1
1.9
0.8
0.1
0.4
0
1.1
1.6
2.5
2.7
3.1
1.7
1.1
2.1
1
1.9
1.6
0
1.6
1.2
1
0.3
0.6
0.8
0.6
0
1.5
0
0.9
2.3
0.4
0.8

0.4
0.5
0.8
0.4
0.2
0.1
0.3
0.3
0
0
0.2
0
0
0.1
0.4
0.4
0
0.3
0.3
0.3
0
0.2
0.1
0.7
0.3
0.1
0
0.4
0
0.4
0
0.7
0
0.6
0.2
0
0.3
0.9
0.5
0.2
0.3
0.3
0.2
0.1
0.3
0
0

5.2
5.1
15.3
5.6
4.5
4.5
5.1
1.8
7.3
4.3
5.2
4
7.6
7.1
3.8
3.7
5.4
4.7
3.6
5.9
2.5
4.8
6.3
6.4
5.1
7.6
6.9
2.2
6.4
7
4.5
5
2.9
7.4
4.2
4.1
1.7
5.5
4.2
4.4
2.4
2.9
1.3
1.7
8.1
4.6
4.7

2.4
3.4
9.7
6.2
2.7
3.5
3.7
11.4
4.6
1.9
4.2
1.1
1
2.9
8.9
3.3
5.2
3.8
3.2
4.2
0.5
4.9
1.5
11.8
5
2.7
3.3
4.7
2.7
10
16.3
0.5
3.2
6.6
2.6
2.5
11.6
0.9
4
1.1
0.5
0.3
2.9
0.7
2.6
0.3
0.4

0.3
0.3
0.3
0.2
0.1
0
0.3
0.2
0
0
0.3
0
0
0.2
0.7
0.5
0
0.5
0.2
0.2
0
0.2
0.2
0.6
0.2
0
0.2
0.1
0
0.5
0.3
0.5
0.2
0.5
0.1
39.7
0.1
0.2
0.4
0.6
0
0.1
0.1
0
0.5
0
0

0.1
0.1
1
0.1
0.5
0.8
0.7
0
0
0
0
1
0
0.6
0.2
0.2
0
0
0
0.7
3.2
0.1
2.7
0
0
0.1
2.3
0
1.4
0.1
0.5
0.1
5.6
0
0
1
0
0.2
0
21.3
0
0.4
0
0
0.2
0
0

0
0
0
0.2
0.1
0.3
0.2
0.2
0
0.3
0.2
1.4
0
0.2
0
0.2
0.1
0
0
1.7
5.4
0
1.4
0.3
0.4
0.9
1.9
0.3
0
0.3
0.8
0.2
2.2
0
0.1
0.4
0.3
0
0.1
5.2
0.1
0.4
0
0
0
0
0

2.8
3.8
11
6.7
3.4
4.6
4.9
11.8
4.6
2.2
4.7
3.5
1
3.9
9.8
4.2
5.3
4.3
3.4
6.8
9.1
5.2
5.8
12.7
5.6
3.7
7.7
5.1
4.1
10.9
17.9
1.3
11.2
7.1
2.8
43.6
12
1.3
4.5
28.2
0.6
1.2
3
0.7
3.3
0.3
0.4

Sh
Sh
Sh
Sh
Ss
Ss
Ss
St
Ss
Ss
Sh
Ss
Ss
Ss
Sh
St
Ss
Sh
St
Ss
Ss
St
Ss
Sh
St
Ss
Ss
St
Ss
Sh
Ss
Sh
Ss
Sh
Ss
Ss
St
Ss
Sh
St
Ss
St
Ss
Ss
Ss
Ss
Ss

Appendix 5-16 (cont.): X-ray diffraction results of fine-grained samples from the Narrabeen Group.
Samples

CCSS - EAW 30 (27)


CCSS - EAW 30 (28)
CCSS - EDEN 124 (8)
CCSS - EDEN 124 (9)
CCSS - EDEN 124 (10)
CCSS - EDEN 124 (11)
CCSS - EDEN 124 (12)
CCSS - EDEN 125 (12)
CCSS - EDEN 125 (13)
CCSS - EDEN 125 (14)
CCSS - EDEN 127 (11)
CCSS - EDEN 127 (12)
CCSS - EDEN 127 (13)
CCSS - EDEN 127 (14)

Height
above
base

98.1
97.7
113
110
105.6
103.1
99.2
101.2
100
96.8
105
103
100.7
97.8

34.8
31.8
39.2
53.6
55.5
60.3
30.7
41
36.1
53.2
48.2
56.8
50
36

Plag

Cl
K-f

Ab

Lab

Orth

Mic

0.6
0.7
1.5
0.9
3.9
0.1
0
0.7
0.7
0.1
5.8
5.6
5.1
0.8

0
0
2.3
1.2
0.9
0.9
0
0.7
0.5
0.9
1.5
1.1
1.3
0

0
0
0
0
0
0
0
0
0
0
0
0
0
0

3.8
3.6
3.1
2.6
1.4
0.7
3.1
1.7
3
1.3
2.3
1.5
2.1
2.8

HML

TF

ill

Ka

Di

Chl

Mix

TCl

4.4
4.3
6.9
4.7
6.2
1.7
3.1
3.1
4.2
2.3
9.6
8.2
8.5
3.6

1.7
2.3
4.2
4.1
4.6
3
1.4
2.5
4.5
1.9
5.4
4.6
3.8
2.9

17.6
10.1
6
8.6
6.3
12.1
10.1
8.7
10.3
18.1
3
3.4
1.3
12.2

0.5
0.4
1.7
1.8
1.8
1
0
0.9
1.4
1.5
1.1
0
1
0

2.4
2.2
0.4
0
0
0
1.4
0
1.4
0.4
0
0
0
1.3

15.5
18.1
20.3
14.8
12.7
8.9
27.4
11.4
20.2
11.4
12
12.9
15
19.1

37.7
33.1
32.6
29.3
25.4
25
40.3
23.5
37.8
33.3
21.5
20.9
21.1
35.5

Mu

18.5
25.3
6
1.2
2.1
0.7
22.4
3.1
16.8
3.8
0
0.7
1.8
20.7

Carb

He

Ho

Py

THML

Ca

Tcarb

Rock
type

0.2
0.5
0.1
0.1
0.2
0
0
0
0.2
0
0
0
0
0

0
0
3.2
3.3
3.5
2.2
0.4
2.7
1
2
3.8
3.8
4.4
0.3

0
0
0
0
0.3
0
0
0
0
0
0
0
0
0

0.4
0.5
0.1
0
0
0
0.2
0
0.3
0
0
0
0
0.3

1.2
0.5
1.9
1.7
1.3
0
0.6
0.7
1.5
1.6
0
0
0.5
0.3

0.5
0.7
0.2
0.1
0
0.1
0.4
0
0.5
0.3
0.1
0
0
0.6

2.3
2.2
5.5
5.2
5.3
2.3
1.6
3.4
3.5
3.9
3.9
3.8
4.9
1.5

1.9
2.2
9.3
1.9
0.8
0.7
1.5
3.4
1.1
3.4
1.5
5
12.6
1.9

0.2
0.5
0.2
1.1
0.7
0.8
0.2
15.2
0.4
0
1.2

0.1
0.3
0
1.6
3.4
5.7
0
5.4
0.1
0
12.3
3.2
0.3
0.1

0.1
0.3
0.3
1.4
0.6
2.8
0.2
1.9
0
0.1
1.8
1.4
0.5
0.2

2.3
3.3
9.8
6
5.5
10
1.9
25.9
1.6
3.5
16.8
9.6
13.7
2.7

St
St
St
Ss
Ss
Ss
St
Ss
St
Ss
Ss
Ss
Ss
St

Q = Quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar, Ab = Albite, Lab = Labradorite, Orth = Orthoclase, Mic = Microcline, TF = Total
feldspar, Cl = Clay minerals, ill = illite, Ka = Kaolinite, Di = Dickite, Chl = Chlorite, Mix = Mixed layer illite/smectite, TCl = Total clay minerals, Mu =
Muscovite, HML = Heavy minerals, He = Hematite, Ho, Horenblende, R = Rutile, Z = Zircon, T = Tourmaline, Py = Pyrite, THML = Total heavy
minerals, Carb = Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, Ss = Sandstone, St =
Siltstone, Sh = Shale.

77

0.3
0.5

Appendix 5-17: X-ray diffraction results of fine-grained samples from the Hawkesbury Sandstone.
Samples

HBSS - Surface 1
HBSS - Surface 3
HBSS - Surface 4
HBSS - Surface 5
HBSS - Surface 7
HBSS - Surface 8
HBSS - Surface 9
HBSS - EAW 18 a (1)
HBSS - EAW 18 a (3)
HBSS - EAW 18 a (5)
HBSS - EAW 18 a (8)
HBSS - EAW 18 a (9)
HBSS - EAW 18 a (10)
HBSS - EDEN 115 (2)
HBSS - EDEN 115 (4)
HBSS - EDEN 115 (5)
HBSS - EDEN 115 (7)
HBSS - EDEN 115 (10)
HBSS - EDEN 115 (11)
HBSS - EDEN 115 (12)
HBSS - EDEN 115 (16)

Height
above
base

399.9
386.4
380.9
372.5
370.8
366.1
442.4
431.8
427.5
418.8
403.5
398
394.9
367.7

F
Q

47.9
29.1
38.5
42.6
74.1
66.3
68.9
56.2
57.5
64.9
27.8
70.8
72.3
71.6
58
53.8
36.2
78
33.2
68.9
66.4

Plag

Cl
K-f

Ab

Lab

Orth

Mic

0.2
0.3
1
0.2
0.6
0
0
0.3
0.2
0
0.9
0
0
0
0.3
0
1
0
1.3
0
0.2

1.3
0
2.2
1.4
0.4
0.6
0
0.4
0
0
0
0
0
0
0
0
0.1
0
0
0.7
0

0
0
0
0
0
0
0
0
0
0
0.6
0
0
0
0
0
0
0
0
0
0

1.4
3.5
4.6
4.1
1.9
0
2.1
0.5
0.7
0
1.8
0
0
0
0
0
1.9
0
1.9
0.9
0

HML

TF

ill

Ka

Di

Chl

Mix

TCl

2.9
3.8
7.8
5.7
2.9
0.6
2.1
1.2
0.9
0
3.3
0
0
0
0.3
0
3
0
3.2
1.6
0.2

5.2
2.6
6.5
4.5
2.1
2.3
0.1
0.2
0
0
2
0.7
0
0.2
0
0
2.6
0.6
1
0
0

7.7
14.5
11
15.1
7.7
8
20.3
21.1
28.5
25.5
29.3
17.3
21.2
22.2
30.9
39
26.4
13
34.9
9.3
20

3.1
0.2
3.2
2.3
1.9
7.5
1.3
1.9
2
7.5
0
2.3
1.8
5.6
4.9
0.7
1
5
0
5.9
4

2.8
1.5
2.1
0
0
0.7
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

18.5
22.3
9.1
8.9
8.7
8.5
4
4.1
4.7
0.2
17.8
5.3
3
0.1
4.7
6.3
13.2
0.2
13.7
8.2
5.8

37.3
41.1
31.9
30.8
20.4
27
25.7
27.3
35.2
33.2
49.1
25.6
26
28.1
40.5
46
43.2
18.8
49.6
23.4
29.8

Mu

3.4
21
11.7
12.6
0
0
0
0
0
0
12
0
0
0
0
0
9.4
0
6.2
0
0

Carb

He

Ho

Py

THML

Ca

Tcarb

Rock
type

0.1
0.2
0.4
0
0
0.1
0
0
0
0
0.5
0
0
0
0
0
0.5
0
0.2
0
0

2.1
0.7
4.2
5.1
2.2
0.5
1.3
2.8
1.7
0
0
0
0
0
0.5
0
0
0
0.6
0.6
0

0
0
0
0
0
0.4
0.5
0
0
0.3
0
0.5
0
0
0
0
0
0.4
0
0
0.3

0
0.6
0.5
0.5
0
0
0
0
0
0
0.1
0
0
0
0
0
0
0.2
0
0
0

5.3
0.8
3.9
2.3
0
4.8
1.5
0.1
0
0
0.8
0
0
0
0
0
1
0
1.5
4
0.6

0.2
0.5
0.5
0
0
0.2
0
0.1
0
0.2
0.9
0.3
0
0.2
0.1
0.1
0.7
0.2
0.7
0
0.2

7.7
2.8
9.5
7.9
2.2
6
3.3
3
1.7
0.5
2.3
0.8
0
0.2
0.6
0.1
2.2
0.8
3
4.6
1.1

0.3
1.3
0.1
0.1
0
0.1
0
0.9
4.7
1.2
4.6
2.7
1.7
0
0
0
5.3
2.4
4
1.3
2.3

0.4
0.5
0.4
0.3
0.3
0
0
0
0
0
0.4
0
0
0
0
0
0.2
0
0.2
0
0

0
0.2
0
0
0.1
0
0
8.1
0
0
0
0
0
0
0.6
0
0.3
0
0.2
0
0

0.1
0.2
0.1
0
0
0
0
3.3
0
0.2
0.5
0.1
0
0.1
0
0.1
0.2
0
0.4
0.2
0.2

0.8
2.2
0.6
0.4
0.4
0.1
0
12.3
4.7
1.4
5.5
2.8
1.7
0.1
0.6
0.1
6
2.4
4.8
1.5
2.5

St
Sh
Sh
Sh
Ss
Ss
Ss
Ss
Ss
Ss
St
Ss
Ss
Ss
Ss
Ss
St
Ss
St
Ss
Ss

Q = Quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar, Ab = Albite, Lab = Labradorite, Orth = Orthoclase, Mic = Microcline, TF = Total
feldspar, Cl = Clay minerals, ill = illite, Ka = Kaolinite, Di = Dickite, Chl = Chlorite, Mix = Mixed layer illite/smectite, TCl = Total clay minerals, Mu =
Muscovite, HML = Heavy minerals, He = Hematite, Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, Py = Pyrite, THML = Total heavy
minerals, Carb = Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, Ss = Sandstone, St =
Siltstone, Sh = Shale.

71

Appendix 5-18: Average of X-ray diffraction results of fine-grained in the Illawarra Coal Measures.
Samples

Medium Sandstone
Fine Sandstone
Siltstone
Shale
Total Sandstone
Tuff
Coal
Carbonate cement

27
24
21
39
51
9
1
1

42.7
38.8
42.7
31.2
40.9
36.6
11.8
3.3

Cl

Plag

K-f

TF

ill

Ka

Di

Chl

Mix

TCl

12.2
13.8
10.5
7.2
12.9
5.2
11
2.5

2.5
2.2
3.4
3.2
2.4
3.9
17.4
1.6

14.7
16
13.8
10.4
15.3
9
28.4
4.1

5.1
3.7
4.9
5.9
4.4
3.8
11
0

7.5
9.1
8.7
13.1
8.2
12.5
25.4
0.3

0.6
0.4
0.7
0.7
0.5
1.7
3.8
2.1

0.5
0.1
0.5
0.5
0.3
0.1
1
0

16.7
13.7
15.6
22.8
15.3
23.9
4.9
4.3

30.4
27
30.4
42.9
28.8
42
46.1
6.7

Mu
2.2
1.8
4.1
5.4
2
0.3
2.8
1.1

HML

Carb

He

Ho

Py

THML

Ca

Tcarb

0.2
0.1
0.1
0.1
0.2
0.04
2.2
0

3.1
2.7
3.3
2.9
2.9
3.9
0
2.9

0
0.008
0.01
0.04
0.004
0.1
0
0

0.1
0.2
0.2
0.3
0.1
0.3
0.5
0

0.8
0.8
0.9
0.9
0.8
0.9
6.2
2.7

0.1
0.1
0.2
0.3
0.1
0.2
0.6
0

4.4
3.9
4.7
4.5
4.1
5.5
9.5
5.6

0.9
1.1
1.8
2.7
1
1.5
1.4
0.4

0.9
2.6
1.1
1.9
1.7
0.2
0
78

2.7
6.5
0.8
0.6
4.5
2.5
0
0.3

1.2
2.3
0.5
0.4
1.7
2.4
0
0.5

5.6
12.4
4.2
5.5
8.8
6.6
1.4
79.2

Appendix 5-19: Average of X-ray diffraction results of fine-grained in the Narrabeen Group.
Samples

Medium Sandstone
Fine Sandstone
Siltstone
Shale
Total Sandstone
Basalt

39
15
18
32
54
3

61.8
52.3
36.6
22
59.2
4.7

Cl

Plag

K-f

TF

ill

Ka

Di

Chl

Mix

TCl

3.1
2.6
1.7
3
3
36.4

2.1
1.7
2.9
3.3
2
8.3

5.1
4.4
4.6
6.3
4.9
44.7

4.4
3.9
3.2
6
4.3
0.8

4.2
10.2
12.3
18.1
5.8
6.3

1.5
1.1
1
1.6
1.4
1.1

0.4
0.3
1.2
2.1
0.4
2.8

10.5
12.8
16.9
17.3
11.1
9.5

20.9
28.3
34.7
45.2
22.9
20.5

Mu
0.8
3.1
13.9
11.3
1.4
13.5

HML

Carb

He

Ho

Py

THML

Ca

Tcarb

0.1
0.2
0.3
2.5
0.1
0.5

4
3.2
1.7
2.8
3.7
0.1

0.1
0.1
0.01
0.1
0.1
2.1

0.01
0.007
0.2
0.7
0.009
0

0.8
1
1.3
2.5
0.9
0.1

0.1
0.2
0.4
0.8
0.1
0.6

5.1
4.6
3.9
9.3
5
3.4

3
4.5
4.2
4.8
3.4
10.3

1.2
1.2
0.3
0.5
1.2
0.5

1.3
1.2
1.3
0.4
1.3
1.6

0.8
0.5
0.5
0.2
0.7
0.7

6.3
7.4
6.3
5.9
6.6
13.2

Appendix 5-20: Average of X-ray diffraction results of fine-grained in the Hawkesbury Sandstone.
Samples

Medium Sandstone
Fine Sandstone
Siltstone
Shale
Total Sandstone
Siderite

12
2
4
3
14
1

68.1
55
36.3
36.7
66.3
6.2

Cl

Plag

K-f

TF

ill

Ka

Di

Chl

Mix

TCl

0.3
0.4
1.2
1.7
0.3
1.3

0.5
0.3
1.9
4.1
0.4
2.7

0.7
0.6
3.1
5.8
0.7
4

0.5
0.1
2.7
4.5
0.4
0.7

18.7
30.1
24.6
13.5
20.3
1.1

4.1
1.3
1
1.9
3.7
2.9

0.1
0
0.7
1.2
0.1
0.7

4.5
5.2
15.8
13.4
4.6
5.9

27.8
36.7
44.8
34.6
29.1
11.3

Mu
0
0
7.8
15.1
0
8.1

HML

Carb

He

Ho

Py

THML

Ca

Tcarb

0.008
0
0.3
0.2
0.007
0.5

0.6
1.4
0.7
3.3
0.7
0.2

0.2
0
0
0
0.2
0

0.02
0
0.03
0.5
0.01
0.3

0.9
0.1
2.2
2.3
0.8
1.9

0.1
0.1
0.6
0.3
0.1
0.5

1.8
1.6
3.8
6.7
1.8
3.4

1.4
0.5
3.6
0.5
1.2
65.5

0.03
0
0.3
0.4
0.02
0.5

0.1
4.1
0.1
0.1
0.6
0

0.1
1.7
0.3
0.1
0.3
1

1.5
6.2
4.3
1.1
2.2
67

Q = Quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar, Ab = Albite, Lab = Labradorite, Orth = Orthoclase, Mic = Microcline, TF = Total
feldspar, Cl = Clay minerals, ill = illite, Ka = Kaolinite, Di = Dickite, Chl = Chlorite, Mix = Mixed layer illite/smectite, TCl = Total clay minerals, Mu =
Muscovite, HML = Heavy minerals, He = Hematite, Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, Py = Pyrite, THML = Total heavy
minerals, Carb = Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement.

77

Appendix 5-21: Average of XRD results of samples from the southern Sydney Basin
Fm

HBSS
NPFM
GRFM
BACS
BGSS
SPCS
SSBS
WBCS
CCSS
LDSS
BASM
LRSS
UNM2
UNM3
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

21
5
2
8
11
2
14
44
21
10
2
21
4
9
1
16
4
10
8
13
5
8

56.3
54.7
0.6
5.2
57.1
45.5
44.3
44.7
48.4
41.9
18.6
39.9
46.3
40.1
51.6
43
35.2
39.9
39.9
29.9
35.8
23.8

Cl

Plag

K-f

TF

ill

Ka

Di

Chl

Mix

TCl

0.6
0.9
2.8
3.5
1
2.4
2.2
3.1
3.7
10.9
6
13.7
12.5
9.7
6.2
10.2
8.7
9.4
14.9
9.3
5.5
6

1.2
1.3
2.1
3.3
1.9
3
3.3
2.7
2.2
2.8
3
2.8
2.9
4.2
2.4
3.1
2.3
2.8
2
2.6
3.1
2.6

1.9
2.2
4.9
6.8
2.9
5.4
5.5
5.8
5.9
13.7
9
16.5
15.4
14
8.6
13.3
11
12.2
16.9
11.9
8.6
8.5

1.5
2.4
2.3
3.3
2.9
7.7
7.3
5.1
3.8
7.3
8.2
4.9
5
4.1
3.5
4.1
2.9
4.2
4.5
7.1
6.6
3.4

20.1
17.2
59
43.2
5.6
3.2
4.5
6.3
10
4
21
6.2
4
5.7
11.9
11.7
20.1
13.7
3.1
12.8
11.3
21

3
0.5
1.7
3.7
1.7
1
1.1
1.3
0.9
0.6
0
0.4
0.5
0.8
0.7
0.7
1.5
0.5
0.3
0.3
0.7
1.3

0.3
0.1
2.2
3.6
0.1
0.6
0.8
1.2
0.7
1
0.5
0.3
0.1
0.2
0.3
0.4
0.1
0.4
0.1
0.7
0.3
0.4

8
6.3
4.6
5.5
10.7
16
17.8
16.5
14.3
19.5
33.7
16.4
16.8
18.2
16.4
15.4
18.4
15.4
13
23.5
25.2
17.1

32.9
26.7
69.8
59.2
20.9
28.4
31.4
30.4
29.6
32.5
63.4
28.2
26.4
29
32.8
32.3
43
34.2
21
44.5
43.9
43.2

Mu
3.6
7.8
2.5
6.1
11.2
5.2
5.3
6.2
6.7
3.9
3.7
3.6
3.1
5.1
0
2.3
1.2
4.4
2.5
5.6
4.4
2.6

HML

Carb

He

Ho

Py

THML

Ca

Tcarb

0.1
0.3
1.1
9.5
0.2
0.4
0.4
0.2
0.1
0.3
0
0.2
0.1
0.03
0
0.04
0.05
0.2
0.04
0.1
0.2
0.1

1.1
2.3
3.7
1.9
2.8
5
4.4
3.3
2.3
3.1
1.7
3
2.6
4.3
2.9
3.1
2.9
3.5
3.4
2
2.8
2.1

0.1
0.3
0.3
0.2
0.3
0
0
0.02
0.02
0
0.4
0
0
0
0
0.03
0
0
0
0.03
0
0.1

0.1
0.1
1
1
0.3
0.3
0.3
0.2
0.1
0.2
0.4
0.2
0.1
0.3
0
0.1
0.1
0.2
0.1
0.3
0.2
0.2

1.3
0.9
4
5
0.6
1.2
1.6
1.3
0.8
0.7
0.7
0.7
0.6
0.6
2.4
1.4
1.1
1.2
0.6
0.6
1
1.1

0.2
0.3
0.6
1.5
0.3
0.4
0.4
0.3
0.2
0.2
0.5
0.1
0
0.1
0
0.2
0.3
0.2
0.1
0.2
0.4
0.2

2.9
4.2
10.6
19.1
4.4
7.3
7.1
5.2
3.6
4.4
3.6
4.3
3.3
5.4
5.3
4.8
4.4
5.3
4.2
3.3
4.5
3.7

1.6
3.6
10.7
2.2
2.4
2.8
4.9
4.8
2.6
1
1.2
1.2
2.7
2.6
0.9
0.9
1.9
0.9
1.3
1.8
1.6
5.7

0.1
0.1
0.4
0.4
0.3
0.5
0.4
1.1
1.1
0.6
0.3
0.7
0.2
0.2
0
0.9
0.1
0.8
5
1.7
1
8.7

0.5
0.6
0.5
0.8
0.4
3.3
0.7
1.1
1.6
1.2
0.2
3.7
1.8
2.2
0.7
1.9
1.6
2
7.5
1
0.3
2.4

0.3
0.2
0.2
0.2
0.2
1.8
0.4
0.6
0.6
0.8
0.1
1.9
0.9
1.4
0.1
0.6
1.7
0.4
1.7
0.3
0.1
1.4

2.4
4.5
11.8
3.6
3.3
8.3
6.5
7.6
5.8
3.5
1.8
7.6
5.5
6.4
1.7
4.3
5.2
4.1
15.5
4.8
2.9
18.1

Q = Quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar, Ab = Albite, Lab = Labradorite, Orth = Orthoclase, Mic = Microcline, TF = Total
feldspar, Cl = Clay minerals, ill = illite, Ka = Kaolinite, Di = Dickite, Chl = Chlorite, Mix = Mixed layer illite/smectite, TCl = Total clay minerals,
Mu = Muscovite, HML = Heavy minerals, He = Hematite, Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, Py = Pyrite, THML = Total
heavy minerals, Carb = Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement.

77

Appendix 5-22: Average of XRD results of subsurface samples from the southern Sydney Basin
Fm
HBSS
NPFM
GRFM
BACS
BGSS
SPCS
SBSS
WBCS
CCSS
BUSM
LDSS
BASM
LRSS
CHSM
UNM2
HGSM
UNM3
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

Av
thick

Base
of
unit

Av C
thick

138.4
13.9
2.4
21.8
135.1
14.4
36.3
30.8
15.4
3
7.5
1.9
6.8
0.4
12.9
0.5
5.8
10.5
6.2
5.9
5.5
5.7
11.4
6.4
5.2

365.7
351.8
349.4
327.6
192.5
178.1
141.8
111
95.6
92.6
85.1
83.2
76.4
76
63.1
62.6
56.8
46.3
40.1
34.2
28.7
23
11.6
5.2
0

434.9
358.8
350.6
338.5
260.1
185.3
160
126.4
103.3
94.1
88.9
84.2
79.8
76.2
69.6
62.9
59.7
51.6
43.2
37.2
31.5
25.9
17.3
8.4
2.6

Cl

HML

Mu

Plag

K-f

TF

ill

Ka

Di

Chl

Mix

TCl

Carb

He

Ho

Py

THML

Ca

Tcarb

14
5
2
8
11
2
14
44
20

58.3
54.7
0.6
5.2
57.1
45.5
44.3
44.7
47.8

0.4
0.9
2.8
3.5
1
2.4
2.2
3.1
3.6

0.6
1.3
2.1
3.3
1.9
3
3.3
2.7
2.3

1
2.2
4.9
6.8
2.9
5.4
5.5
5.8
5.9

0.5
2.4
2.3
3.3
2.9
7.7
7.3
5.1
3.8

24.2
17.2
59
43.2
5.6
3.2
4.5
6.3
9.5

3
0.5
1.7
3.7
1.7
1
1.1
1.3
1

0
0.1
2.2
3.6
0.1
0.6
0.8
1.2
0.7

6.2
6.3
4.6
5.5
10.7
16
17.8
16.5
14.7

34
26.7
69.8
59.2
20.9
28.4
31.4
30.4
29.7

2
7.8
2.5
6.1
11.2
5.2
5.3
6.2
6.9

0.1
0.3
1.1
9.5
0.2
0.4
0.4
0.2
0.1

0.4
2.3
3.7
1.9
2.8
5
4.4
3.3
2.3

0.1
0.3
0.3
0.2
0.3
0
0
0.02
0

0.02
0.1
1
1
0.3
0.3
0.3
0.2
0.1

0.6
0.9
4
5
0.6
1.2
1.6
1.3
0.9

0.3
0.3
0.6
1.5
0.3
0.4
0.4
0.3
0.2

1.5
4.2
10.6
19.1
4.4
7.3
7.1
5.2
3.6

2.2
3.6
10.7
2.2
2.4
2.8
4.9
4.8
2.7

0.1
0.1
0.4
0.4
0.3
0.5
0.4
1.1
1.2

0.7
0.6
0.5
0.8
0.4
3.3
0.7
1.1
1.7

0.4
0.2
0.2
0.2
0.2
1.8
0.4
0.6
0.6

3.3
4.5
11.8
3.6
3.3
8.3
6.5
7.6
6.1

10
2
17

41.9
18.6
37.9

10.9
6
15.4

2.8
3
2.2

13.7
9
17.7

7.3
8.2
5.2

4
21
5.3

0.6
0
0.4

1
0.5
0.2

19.5
33.7
17.1

32.5
63.4
28.2

3.9
3.7
3.4

0.3
0
0.2

3.1
1.7
3

0
0.4
0

0.2
0.4
0.1

0.7
0.7
0.5

0.2
0.5
0.1

4.4
3.6
3.9

1
1.2
1.4

0.6
0.3
0.8

1.2
0.2
4.5

0.8
0.1
2.2

3.5
1.8
8.9

46.3

12.5

2.9

15.4

0.5

0.1

16.8

26.4

3.1

0.1

2.6

0.1

0.6

3.3

2.7

0.2

1.8

0.9

5.5

9
1
12
4
10
8
13
5
8

40.1
51.6
41.2
35.2
36.8
39.9
29.9
35.8
23.8

9.7
6.2
12.3
8.7
9.5
14.9
9.3
5.5
6

4.2
2.4
2.8
2.3
2.6
2
2.6
3.1
2.6

14
8.6
15.1
11
12.1
16.9
11.9
8.6
8.5

4.1
3.5
4.5
2.9
4.3
4.5
7.1
6.6
3.4

5.7
11.9
11.3
20.1
13.7
3.1
12.8
11.3
21

0.8
0.7
0.6
1.5
0.6
0.3
0.3
0.7
1.3

0.2
0.3
0.5
0.1
0.5
0.1
0.7
0.3
0.4

18.2
16.4
16.6
18.4
17.7
13
23.5
25.2
17.1

29
32.8
33.4
43
36.7
21
44.5
43.9
43.2

5.1
0
2
1.2
4.3
2.5
5.6
4.4
2.6

0.03
0
0.1
0.05
0.1
0.04
0.1
0.2
0.1

4.3
2.9
2.3
2.9
2.9
3.4
2
2.8
2.1

0
0
0.03
0
0
0
0.03
0
0.1

0.3
0
0.1
0.1
0.2
0.1
0.3
0.2
0.2

0.6
2.4
1.3
1.1
1
0.6
0.6
1
1.1

0.1
0
0.2
0.3
0.2
0.1
0.2
0.4
0.2

5.4
5.3
3.9
4.4
4.4
4.2
3.3
4.5
3.7

2.6
0.9
1.2
1.9
1.3
1.3
1.8
1.6
5.7

0.2
0
0.2
0.1
1.1
5
1.7
1
8.7

2.2
0.7
2.4
1.6
2.7
7.5
1
0.3
2.4

1.4
0.1
0.7
1.7
0.5
1.7
0.3
0.1
1.4

6.4
1.7
4.5
5.2
5.7
15.5
4.8
2.9
18.1

Q = Quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar, Ab = Albite, Lab = Labradorite, Orth = Orthoclase, Mic = Microcline, TF = Total
feldspar, Cl = Clay minerals, ill = illite, Ka = Kaolinite, Di = Dickite, Chl = Chlorite, Mix = Mixed layer illite/smectite, TCl = Total clay minerals, Mu =
Muscovite, HML = Heavy minerals, He = Hematite, Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, Py = Pyrite, THML = Total heavy
minerals, Carb = Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement.

788

Appendix 5-23: X-ray diffraction results of other samples from the southern Sydney Basin.
Plag
Samples

HBSS - Surface 2
CCSS - EDEN 126 (13)
CCSS - EDEN 126 (14)
CCSS - EDEN 126 (15)
BULI - Surface 1
WWCO - EAW 156 (23)
WWCO - EAW 30 (45)
WWCO - EAW 30 (53)
WWCO - EAW 30 (47)
WWCO - EAW 30 (48)
KBSS - Surface 8
ACSM - EDEN 124 (26)
ACSM - EDEN 124 (27)
ACSM - EDEN 125 (23)
ACSM - EDEN 127 (26)

6.2
11.8
1.2
1.1
11.8
22.5
12.8
36.7
40.2
43
3.3
54.2
38.8
28
53

K-f

Ab

Lab

Orth

Mic

0.4
25.5
11.5
7.2
4.5
0.7
3.3
3.5
3.1
6.8
0
0.3
0.6
0.7
0

0.9
12.1
25.5
27.5
6.5
2.9
3.1
3.2
5.8
5.2
2.5
1.6
2
2
1.6

1.6
0
11.2
2.6
5.3
0
0.5
0
0
0
0
0
0
0
0

1.1
1.1
6.9
3.1
12.1
2.5
4.5
4.3
5
4.9
1.6
3.3
3.2
2.3
4.5

Cl
TF
4
38.7
55.1
40.4
28.4
6.1
11.4
11
13.9
16.9
4.1
5.2
5.8
5
6.1

HML

ill

Ka

Di

Chl

Mix

TCl

0.7
2.5
0
0
11
3.8
17.9
0
3.4
1.5
0
0
2.6
5.3
0.1

1.1
11
0.2
7.6
25.4
16.8
0.2
2.3
0.9
0.4
0.3
24.4
27.7
26
13.9

2.9
0
2.8
0.5
3.8
2.2
0
1.5
1.4
0.6
2.1
1.8
2.2
4.1
1.2

0.7
0
2.2
6.3
1
0
1.3
0
0
0
0
0
0
0
0

5.9
9.1
10.2
9.2
4.9
44
48.5
11.5
31
29.8
4.3
6.7
16.9
19.2
7.1

11.3
22.6
15.4
23.6
46.1
66.8
67.9
15.3
36.7
32.3
6.7
32.9
49.4
54.6
22.3

Mu
8.1
4.3
6.3
29.8
2.8
0
1.3
0
1
0
1.1
0
0.1
0.4
0

Carb

He

Ho

Py

THML

Ca

Tcarb

Rock
type

0.5
0.6
0.3
0.6
2.2
0
0
0
0
0.4
0
0
0
0
0

0.2
0.3
0
0
0
1.8
2.2
4.2
5.6
4.4
2.9
4.7
3.8
3.2
5.5

0
0
3.5
2.7
0
0.5
0.4
0
0
0
0
0
0
0
0

0.3
0
0
0
0.5
0.5
1.1
0
0.4
0.1
0
0
0.3
0.4
0.1

1.9
0.4
0
0
6.2
0.1
0
0.7
1.5
0.7
2.7
0.6
1.2
1.8
1.4

0.5
0.4
1.4
0
0.6
0.4
0.4
0
0.1
0.2
0
0
0.2
0.2
0

3.4
1.7
5.2
3.3
9.5
3.3
4.1
4.9
7.6
5.8
5.6
5.3
5.5
5.6
7

65.5
17.9
13.1
0
1.4
1.1
1.3
0.3
0.4
1.7
0.4
1.3
0.3
6.1
1.3

0.5
0
0.4
1
0
0
1.1
0.3
0.1
0
78
0
0
0
0

0
2
2.1
0.8
0
0
0.1
20
0
0
0.3
0
0
0.1
2.4

1
1
1.2
0
0
0.2
0
11.5
0.1
0.3
0.5
1.1
0.1
0.2
7.9

67
20.9
16.8
1.8
1.4
1.3
2.5
32.1
0.6
2
79.2
2.4
0.4
6.4
11.6

Sid
Ba
Ba
Ba
Co
TF
TF
TF
TF
TF
Cac
TF
TF
TF
TF

Q = Quartz, F = Feldspar, Plag = Plagioclase, K-f = K-feldspar, Ab = Albite, Lab = Labradorite, Orth = Orthoclase, Mic = Microcline, TF = Total
feldspar, Cl = Clay minerals, ill = illite, Ka = Kaolinite, Di = Dickite, Chl = Chlorite, Mix = Mixed layer illite/smectite, TCl = Total clay minerals, Mu =
Muscovite, HML = Heavy minerals, He = Hematite, Ho, Hornblende, R = Rutile, Z = Zircon, T = Tourmaline, Py = Pyrite, THML = Total heavy
minerals, Carb = Carbonate cement, S = Siderite, Ca = Calcite, A = Ankerite, D = Dolomite, Tcarb = Total carbonate cement, Sid = Siderite, Ba =
Altered Basalt, Co = Coal, TF = Tuff, Cac = Carbonate cement sample.

787

Appendix 5-24: Thin section porosity (Po), density porosity (DNPO) and
equivalent porosity (EHP) in the Illawarra Coal Measures
Samples
LDSS - EAW 156 (14)
LDSS - EAW 156 (15)
LDSS - EAW 30 (30)
LDSS - EAW 30 (31)
LDSS - EAW 30 (32)
LDSS - EDEN 124 (13)
LDSS - EDEN 124 (14)
LDSS - EDEN 124 (15)
LDSS - EDEN 124 (16)
LDSS - EDEN 125 (15)
LDSS - EDEN 125 (16)
LDSS - EDEN 125 (17)
LDSS - EDEN 126 (16)
LDSS - EDEN 126 (17)
LDSS - EDEN 126 (18)
LDSS - EDEN 127 (15)
LDSS - EDEN 127 (16)
LDSS - EDEN 127 (17)
LDSS - EDEN 127 (18)
LRSS - Surface 1
LRSS - Surface 3
LRSS - Surface 4
LRSS - EAW 156 (18)
LRSS - EAW 156 (19)
LRSS - EAW 156 (20)
LRSS - EAW 30 (33)
LRSS - EAW 30 (34)
LRSS - EAW 30 (35)
LRSS - EDEN 124 (18)
LRSS - EDEN 124 (19)
LRSS - EDEN 125 (37)
LRSS - EDEN 125 (19)
LRSS - EDEN 126 (19)
LRSS - EDEN 126 (20)
LRSS - EDEN 126 (21)
LRSS - EDEN 127 (19)
LRSS - EDEN 127 (20)
UNM2 - EAW 30 (51)
UNM3 - EAW 156 (22)
UNM3 - EDEN 126 (22)
UNM3 - EDEN 127 (21)
WWCO - EDEN 124 (22)
KBSS - Surface 1
KBSS - Surface 3
KBSS - Surface 5
KBSS - Surface 6
KBSS - Surface 7
KBSS - EAW 156 (25)
KBSS - EAW 156 (26)
KBSS - EAW 156 (27)

P1

Po
P2

TPo

0
1.5
0
0.5
0
0
0
0.7
0.3
0
1
0
0
0.5
0
0
0
0.5
1.5
0
0
0
0
0.5
0.5
0
0
0
0
0
4
0.3
0
0
0
0
0
0
0
0
0
0
0
0
0
0.2
0
0
0.5
0

0
7
0
4.2
0
0
2
3
5
0
4.5
2
0
1.5
0
0
1.2
4
8
0
0
0
0
3
2.5
0.2
1
0
0
0
11
2.4
0
0
0
0
0
0
0
0
0
0
0
0
0
1
1
0
2.2
0.7

0
8.5
0
4.7
0
0
2
3.7
5.3
0
5.5
2
0
2
0
0
1.2
4.5
9.5
0
0
0
0
3.5
3
0.2
1
0
0
0
15
2.7
0
0
0
0
0
0
0
0
0
0
0
0
0
1.2
1
0
2.7
0.7

787

DNPO

EHP

3.4
1.4
0.1
2.8
3.6
0.8
4.4
2.3
1.9
3.1
6.7
9.4
4.2
3.8
2.1
7
7.4
6.7
11.3

2.6
13.1
2.6
8.4
2.6
2.6
5.1
7.2
9.2
2.6
9.4
5.1
2.6
5.1
2.6
2.6
4.1
8.2
14.4
2.6
2.6
2.6
2.6
6.9
6.3
2.8
3.8
2.6
2.6
2.6
21.2
5.9
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
4.1
3.8
2.6
5.9
3.4

6.6
0
4.2
5
2
0
1.5
4
2.3
6.3
4.9
5.1
7.2
7
8.5
0.6
1.4
4.4
8
25

3.7
8.8
0

Appendix 5-24 (Cont.): Thin section porosity (Po), density porosity (DNPO)
and equivalent porosity (EHP) in the Illawarra Coal Measures
Samples
KBSS - EDEN 124 (23)
KBSS - EDEN 124 (24)
KBSS - EDEN 124 (25)
KBSS - EDEN 125 (21)
KBSS - EDEN 125 (22)
KBSS - EDEN 126 (23)
KBSS - EDEN 126 (24)
KBSS - EDEN 126 (25)
KBSS - EDEN 127 (22)
KBSS - EDEN 127 (23)
KBSS - EDEN 127 (24)
ACFM - Surface 1
ACFM - Surface 3
ACFM - EDEN 124 (28)
ACFM - EDEN 126 (28)
DFSS - EDEN 124 (30)
DFSS - EDEN 125 (27)
DFSS - EDEN 125 (28)
DFSS - EDEN 126 (36)
DFSS - EDEN 126 (29)
DFSS - EDEN 127 (29)
DFSS - EDEN 127 (30)
BGCS - EDEN 127 (33)
WTFM - EDEN 124 (35)
WTFM - EDEN 126 (34)
WTFM - EDEN 127 (35)

P1

Po
P2

TPo

0
0.3
0.5
0
1
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

0
3.4
4.2
0
3.2
0
1
1.2
0
0.5
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

0
3.7
4.7
0
4.2
0
1
1.2
0
0.5
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0

787

DNPO

EHP

7.2
7.4
7.2
5.3
6.8
5.5
6.7
7.3
9.5
4.7
6.5

2.6
7.2
8.4
2.6
7.8
2.6
3.8
4.1
2.6
3.2
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6
2.6

22.2
0
7.4
2.8
3.1
2.2
7.7
11.8
2.4
11.2
25
9.4
22

Appendix 5-25: Thin section porosity (Po), density porosity (DNPO) and
equivalent porosity (EHP) in the Narrabeen Group.
Samples
NPFM - EAW 42 (2)
NPFM - EAW 42 (4)
NPFM - EAW 42 (6)
BACS - EAW 42 (14)
BGSS - EAW 42 (18)
BGSS - EAW 42 (19)
BGSS - EAW 42 (21)
BGSS - EAW 42 (22)
BGSS - EAW 42 (23)
BGSS - EAW 42 (24)
BGSS - EAW 42 (25)
BGSS - EAW 42 (27)
BGSS - EAW 42 (29)
BGSS - EAW 42 (30)
BGSS - EAW 42 (31)
BGSS - EAW 42 (32)
SPCS - EAW 42 (34)
SPCS - EAW 42 (35)
SBSS - EAW 42 (37)
SBSS - EAW 42 (38)
SBSS - EAW 42 (39)
SBSS - EAW 42 (42)
SBSS - EAW 42 (+42)
SBSS - EAW 42 (43)
SBSS - EAW 42 (44)
SBSS - EAW 42 (45)
SBSS - EAW 42 (+45)
SBSS - EAW 30 (1)
SBSS - EAW 30 (2)
SBSS - EAW 30 (4)
SBSS - EAW 30 (5)
SBSS - EAW 30 (6)
SBSS - EAW 30 (7)
SBSS - EDEN 125 (1)
SBSS - EDEN 125 (2)
SBSS - EDEN 125 (3)
SBSS - EDEN 125 (4)
WBCS - EAW 156 (1)
WBCS - EAW 156 (5)
WBCS - EAW 156 (6)
WBCS - EAW 42 (47)
WBCS - EAW 42 (48)
WBCS - EAW 42 (+48)
WBCS - EAW 42 (49)

P1
0.5
0
0
0
2.2
2.5
1.2
1
2.5
2
3.7
3
0
1.7
2.5
1
0
2.5
1
1.5
0
0.6
0
0
0
0.4
3
1
1.5
0.5
0
0
1.5
0.3
0.5
0.5
0.5
0.5
0
0
0
0
0
0

Po
P2
2
0
0
1.5
9.7
12.2
6
6.9
11.4
7.4
13.2
14.3
4.2
11.7
14
9.2
3
8
8.5
18
6.3
3.4
0
3.6
5.6
5.5
11
4
3
3.2
0
0
14
1.7
3.7
9.5
8.8
4
0.2
0
0.5
5.9
0
0

787

TPo
2.5
0
0
1.5
11.9
14.7
7.2
7.9
13.9
9.4
16.9
17.3
4.2
13.4
16.5
10.2
3
10.5
9.5
19.5
6.3
4
0
3.6
5.6
5.9
14
5
4.5
3.7
0
0
15.5
2
4.2
10
9.3
4.5
0.2
0
0.5
5.9
0
0

DNPO

EHP

0.7
3
0
0
0.7
14.6
4.9
4
13.2
4.9
0
13.9
0
0.3
15.8
5.9
0
0.9
1.5
4.3
6.6
0.6
0
0
1.5
0
1.4
3.9
2.3
10.8
0.8
0
0
6.9
9
4.2
2.5
2.4
5.1
4.6
0
1
3.3
15.8

5.7
2.6
2.6
4.4
17.4
20.9
11.5
12.4
19.9
14.3
23.6
24.1
7.8
19.2
23.1
15.3
6.3
15.6
14.4
26.8
10.4
7.5
2.6
7
9.5
9.9
20
8.8
8.2
7.2
2.6
2.6
21.8
5.1
7.8
15
14.1
8.2
2.8
2.6
3.2
9.9
2.6
2.6

Appendix 5-25 (Cont.): Thin section porosity (Po), density porosity (DNPO)
and equivalent porosity (EHP) in the Narrabeen Group.
Samples
WBCS - EAW 30 (9)
WBCS - EAW 30 (10)
WBCS - EAW 30 (11)
WBCS - EAW 30 (12)
WBCS - EAW 30 (14)
WBCS - EAW 30 (15)
WBCS - EAW 30 (16)
WBCS - EAW 30 (17)
WBCS - EAW 30 (18)
WBCS - EAW 30 (20)
WBCS - EAW 30 (21)
WBCS - EDEN 124 (2)
WBCS - EDEN 124 (3)
WBCS - EDEN 124 (4)
WBCS - EDEN 124 (5)
WBCS - EDEN 124 (6)
WBCS - EDEN 124 (7)
WBCS - EDEN 125 (6)
WBCS - EDEN 125 (7)
WBCS - EDEN 125 (8)
WBCS - EDEN 125 (10)
WBCS - EDEN 126 (38)
WBCS - EDEN 126 (3)
WBCS - EDEN 126 (4)
WBCS - EDEN 126 (6)
WBCS - EDEN 126 (7)
WBCS - EDEN 126 (8)
WBCS - EDEN 126 (10)
WBCS - EDEN 126 (12)
WBCS - EDEN 127 (2)
WBCS - EDEN 127 (3)
WBCS - EDEN 127 (4)
WBCS - EDEN 127 (5)
WBCS - EDEN 127 (7)
WBCS - EDEN 127 (8)
WBCS - EDEN 127 (10)
CCSS - Surface 1
CCSS - Surface 2
CCSS - Surface 3
CCSS - EAW 156 (12)
CCSS - EAW 156 (13)
CCSS - EAW 30 (22)
CCSS - EAW 30 (23)
CCSS - EAW 30 (24)

P1
0
0
1.5
0
2.2
0
0.8
3.7
0
0
0
0
0
0
1.5
1
1
0
2.2
0.8
1
0
0.8
0.8
1.8
0.5
0
0
0
0
0.5
0
1.7
0
0
0
0
0
0
0.8
0.5
2
0
1.3

Po
P2
0
0
9.2
0
10.3
0.3
8.4
0.5
0
0
0
0
0
0.2
4.5
8.2
3.8
2.7
7
10.2
10.5
0
11.4
12.7
4.2
5.5
0
0
0
0
2.5
0
3.8
3.5
0
0
1
0
1.7
15.2
12.2
9.2
0
7.7

787

TPo
0
0
10.7
0
12.5
0.3
9.2
4.2
0
0
0
0
0
0.2
6
9.2
4.8
2.7
9.2
11
11.5
0
12.2
13.5
6
6
0
0
0
0
3
0
5.5
3.5
0
0
1
0
1.7
16
12.7
11.2
0
9

DNPO

EHP

0
0
0
4
8.8
7.1
5.2
5.5
8.6
4
7
0
0.1
0.6
9.2
8
9.7
1.4
0.8
8.4
9.4
3.1
5.3
3.9
9.5
6.1
0.6
2.7
0.7
3.3
5.3
6.1
10
7.9
7.5
1.6

2.6
2.6
15.9
2.6
18.1
2.9
14
7.8
2.6
2.6
2.6
2.6
2.6
2.8
10
14
8.5
5.9
14
16.2
16.9
2.6
17.7
19.4
10
10
2.6
2.6
2.6
2.6
6.3
2.6
9.4
6.9
2.6
2.6
3.8
2.6
4.7
22.5
18.4
16.5
2.6
13.8

3
0.4
6.2
3.4
4.3

Appendix 5-25 (Cont.): Thin section porosity (Po), density porosity (DNPO)
and equivalent porosity (EHP) in the Narrabeen Group
Samples
CCSS - EAW 30 (25)
CCSS - EAW 30 (26)
CCSS - EAW 30 (54)
CCSS - EDEN 124 (9)
CCSS - EDEN 124 (10)
CCSS - EDEN 124 (11)
CCSS - EDEN 125 (11)
CCSS - EDEN 125 (12)
CCSS - EDEN 125 (14)
CCSS - EDEN 127 (11)
CCSS - EDEN 127 (12)
CCSS - EDEN 127 (13)

P1

Po
P2

TPo

2
0
0
0
0
0.5
0.5
0
0
0.7
0.7
0

8.5
3
0.2
0
0
3
7.7
0
0
5.5
5
1

10.5
3
0.2
0
0
3.5
8.2
0
0
6.2
5.7
1

787

DNPO

EHP

0
0
5.1
3.5
0
0
5.8
0
0.4
3.5
2.6
6.9

15.6
6.3
2.8
2.6
2.6
6.9
12.8
2.6
2.6
10.3
9.7
3.8

Appendix 5-26: Thin section porosity (Po), density porosity (DNPO) and
equivalent porosity (EHP) in the Hawkesbury Sandstone
Samples

Po
P1

P2

TPo

16.5
10.7
4.8
8.2
15
1.1
2.6
0
5.4

1.6
0.8
3.8
3
4.3
0.5
0.5
0
0.5

18.1
11.5
8.6
11.2
19.3
1.6
3.1
0
5.9

HBSS - EAW 18 a (3)

HBSS - EAW 18 a (4)

16

17

HBSS - EAW 18 a (5)

13

HBSS - EAW 18 a (7)

11.5

3.5

15

HBSS - Surface 6
HBSS - Surface 7
HBSS - Surface 8
HBSS - Surface 9
HBSS - Surface 10
HBSS - Surface 11
HBSS - Surface 12
HBSS - EAW 18 a (1)
HBSS - EAW 18 a (2)

HBSS - EAW 18 a (8)

HBSS - EAW 18 a (9)


HBSS - EAW 18 a (10)
HBSS - EDEN 115 (1)
HBSS - EDEN 115 (2)
HBSS - EDEN 115 (3)
HBSS - EDEN 115 (4)
HBSS - EDEN 115 (5)
HBSS - EDEN 115 (6)
HBSS - EDEN 115 (7)
HBSS - EDEN 115 (8)
HBSS - EDEN 115 (9)
HBSS - EDEN 115 (10)
HBSS - EDEN 115 (11)
HBSS - EDEN 115 (12)
HBSS - EDEN 115 (13)
HBSS - EDEN 115 (14)
HBSS - EDEN 115 (15)
HBSS - EDEN 115 (16)

2.2
0.2
14
0
16
8
0
9.2
0
15
2.7
11.2
0
14
3.7
14.2
2.5
0

1.3
0
3.7
0
2
1.5
0
2.3
0
3.1
1
3.5
0
2
0.5
2.3
1.7
0

3.5
0.2
17.7
0
18
9.5
0
11.5
0
18.1
3.7
14.7
0
16
4.2
16.5
4.2
0

781

DNPO

EHP

25.1
16.9
13.3
16.5
26.6
4.6
6.4
2.6

15.5
16.7
15.5
16.6
17
1.1
12.6

9.9
2.6
23.7
18.7
21.2
2.6
6.9

0
14.2
15.9
19
17.1
6.9
15.8
0
25
18.3
10.9
6.6
9.4
12.4
15.6
12.1
0

2.8
24.6
2.6
25
14.4
2.6
16.9
2.6
25.1
7.2
20.9
2.6
22.5
7.8
23.1
7.8
2.6

Appendix 5-27: Density porosity from well logs for all depths from all wells.
Fm

Max

Ave

Man

EAW 18a-HBSS
EAW 156-HBSS
EAW 156-NBFM
EAW 156-GRFM
EAW 156-BACS
EAW 156-BGSS
EAW 156-SPCS
EAW 156-SBSS
EAW 156-WBCS
EAW 156-CCSS
EAW 156-BUSM
EAW 156-LDSS
EAW 156-BASM
EAW 156-LRSS
EAW 156-CHSM
EAW 156-UNM2
EAW 156-HGSM
EAW 156-UNM3
EAW 156-WWCO
EAW 156-KBSS
EAW 30-HBSS
EAW 30-NBFM
EAW 30-GRFM
EAW 30-BACS
EAW 30-BGSS
EAW 30-SPCS
EAW 30-SBSS
EAW 30-WBCS
EAW 30-CCSS
EAW 30-BUSM
EAW 30-LDSS
EAW 30-BASM
EAW 30-LRSS
EAW 30-CHSM
EAW 30-UNM2
EAW 30-HGSM
EAW 30-UNM3
EAW 30-WWCO
EAW 30-KBSS

24.63
24.99
17.89
4.81
14.62
17.55
23.49
12.18
24.96
25.01

12.7
11.8
5
2.5
3.5
6.4
6.9
4.5
7.7
3.5

0.01
0
0.01
0.08
0
0
0
0
0
0.01

24.92

4.1

0.12

24.42

3.4

24.5

4.7

0.01

24.26

7.1

23.49
19.52
18.81
6.09
9.9
19.11
12.59
11.32
14.95
23.6

6.9
8.7
8.3
3.2
2.5
7.9
6
3.9
5.5
3.7

0.02
0
0.01
0.17
0
0
0
0
0
0

25

5.9

24.23

3.5

0.02

24.6

5.2

0.01

18.72

6.8

0.03

11.01

5.7

0.06

787

Appendix 5-27 (Cont.): Density porosity from well logs for all depths from all wells
Fm

Max

Ave

Man

EAW 42-HBSS
EAW 42-NBFM
EAW 42-GRFM
EAW 42-BACS
EAW 42-BGSS
EAW 42-SPCS
EAW 42-SBSS
EAW 42-WBCS
EAW 42-CCSS
EAW 42-BUSM
EAW 42-LDSS
EAW 42-BASM
EAW 42-LRSS
EAW 42-CHSM
EAW 42-UNM2
EAW 42-HGSM
EAW 42-UNM3
EAW 42-WWCO
EAW 42-KBSS
EDEN 115-HBSS
EDEN 124-HBSS
EDEN 124-NBFM
EDEN 124-GRFM
EDEN 124-BACS
EDEN 124-BGSS
EDEN 124-SPCS
EDEN 124-SBSS
EDEN 124-WBCS
EDEN 124-CCSS
EDEN 124-BUSM
EDEN 124-LDSS
EDEN 124-BASM
EDEN 124-LRSS
EDEN 124-CHSM
EDEN 124-UNM2
EDEN 124-HGSM
EDEN 124-UNM3
EDEN 124-WWCO
EDEN 124-KBSS
EDEN 124-ACSM
EDEN 124-ACFM
EDEN 124-DFSS
EDEN 124-BGCS
EDEN 124-TGSM
EDEN 124-WTFM

24.87
15.65
11.12
9.59
23.04
10.49
17.64
24.88
13.93

8.3
4.3
3.2
2.2
8
3.4
5.1
5.9
3.9

0
0
0.06
0.01
0
0.08
0
0
0

22.96

7.5

0.19

24.55

4.3

0.02

24.98

8.1

0.01

23.09
24.94
25
17.31
8.64
18.05
24.91
15.41
17.84
13.78
24.92

4.9
14.3
13.7
7.9
3.1
3.2
8.3
3.6
6.4
5.8
2.8

0.03
0.02
0
0.01
0.33
0
0
0
0
0
0

24.99

4.9

24.56

4.9

0.08

23.97

6.5

0.01

23.4

0.29

24.88
8.66
23.76

6.4
3.6
4.8

0.01
0
0

23.43

7.3

0.01

787

Appendix 5-27 (Cont.): Density porosity from well logs for all depths from all wells
Fm

Max

Ave

Man

EDEN 125-HBSS
EDEN 125-NBFM
EDEN 125-GRFM
EDEN 125-BACS
EDEN 125-BGSS
EDEN 125-SPCS
EDEN 125-SBSS
EDEN 125-WBCS
EDEN 125-CCSS
EDEN 125-BUSM
EDEN 125-LDSS
EDEN 125-BASM
EDEN 125-LRSS
EDEN 125-CHSM
EDEN 125-UNM2
EDEN 125-HGSM
EDEN 125-UNM3
EDEN 125-WWCO
EDEN 125-KBSS
EDEN 125-ACSM
EDEN 125-ACFM
EDEN 125-DFSS
EDEN 125-BGCS
EDEN 125-TGSM
EDEN 125-WTFM
EDEN 126-HBSS
EDEN 126-NBFM
EDEN 126-GRFM
EDEN 126-BACS
EDEN 126-BGSS
EDEN 126-SPCS
EDEN 126-SBSS
EDEN 126-WBCS
EDEN 126-CCSS
EDEN 126-BUSM
EDEN 126-LDSS
EDEN 126-BASM
EDEN 126-LRSS
EDEN 126-CHSM
EDEN 126-UNM2
EDEN 126-HGSM
EDEN 126-UNM3
EDEN 126-WWCO
EDEN 126-KBSS
EDEN 126-ACSM
EDEN 126-ACFM
EDEN 126-DFSS
EDEN 126-BGCS
EDEN 126-TGSM
EDEN 126-WTFM

24.99
24.64
5.37
21.35
24.61
13.62
16.62
17.18
22.17

13.7
9
1.5
4
8.6
3.9
6.9
5.4
4.2

0
0.05
0.19
0.02
0
0
0.01
0
0.01

24.12

5.7

0.08

24.11

5.4

0.07

23.05

0.32

24.81

6.6

0.16

24.97
14.21
24.23

6.9
4.7
4.5

0.03
0.03
0

23.65
24.59
19.04
11.21
14.33
20.97
23.9
18.85
15.96
23.71

8.5
12.8
7.9
3.8
3.9
8.8
5.7
9
4.5
2.9

0.17
0
0.01
0.02
0
0
0.01
0.02
0
0.01

24.92

0.05

24.93

6.5

0.11

23.07

6.8

0.02

23.9

7.7

0.62

18.97
12.82
23.75

6.8
4.8
4.7

0
0.06
0

24.1

9.3

0.16

778

Appendix 5-27 (Cont.): Density porosity from well logs for all depths from all wells
Fm

Max

Ave

Man

EDEN 127-HBSS
EDEN 127-NBFM
EDEN 127-GRFM
EDEN 127-BACS
EDEN 127-BGSS
EDEN 127-SPCS
EDEN 127-SBSS
EDEN 127-WBCS
EDEN 127-CCSS
EDEN 127-BUSM
EDEN 127-LDSS
EDEN 127-BASM
EDEN 127-LRSS
EDEN 127-CHSM
EDEN 127-UNM2
EDEN 127-HGSM
EDEN 127-UNM3
EDEN 127-WWCO
EDEN 127-KBSS
EDEN 127-ACSM
EDEN 127-ACFM
EDEN 127-DFSS
EDEN 127-BGCS
EDEN 127-TGSM
EDEN 127-WTFM

25
14.93
7.32
10.29
22.98
15.75
18.63
16.21
24.76

10.9
6.4
2.8
3.2
10
6.4
7.8
5.8
5.4

0
0.01
0.23
0
0
0.01
0.01
0
0.01

24.91

7.9

0.19

24.68

8.5

0.24

24.93

8.5

2.35

23.54

8.1

0.73

24.9
24.42
23.96

10.4
6.7
6

1.18
0
0.05

24.39

15.3

3.74

777

Appendix 5-28: Average of density porosity from well logs for all depths
from all wells
Fm

Max

Ave

Min

HBSS
NPFM
GRFM
BACS
BGSS
SPCS
SBSS
WBCS
CCSS
BUSM
LDSS
BASM
LRSS
CHSM
UNM2
HGSM
UNM3
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

24.3
18.3
7.8
14
21.9
16.5
16.2
18.3
22.6

11.9
7
2.9
3.2
8.3
5.1
6.2
5.8
3.8

0.003
0.01
0.2
0.004
0
0.01
0.01
0
0.01

24.5

0.1

24.5

5.2

0.1

24.7

0.01

23

7.1

0.5

21.9

6.7

0.3

23.4
15
23.9

7.6
5
5

0.3
0.02
0.01

23.9

10.1

777

Appendix 5-29: Values of compaction porosity loss (COPL) and


cementation porosity loss (CEPL) in the Illawarra Coal Measures.
Samples

COPL

CEPL

Samples

COPL

CEPL

LDSS - EAW 156 (14)

42.1

LDSS - EAW 156 (15)

34.8

2.9

UNM3 - EDEN 127 (21)

28.1

16.9

4.7

WWCO - EDEN 124 (22)

41.7

3.3

LDSS - EAW 30 (30)


LDSS - EAW 30 (31)

39.6

5.4

KBSS - Surface 1

41.6

3.4

29.9

11.8

KBSS - Surface 3

24

21

LDSS - EAW 30 (32)

42

KBSS - Surface 5

32.8

12.2

LDSS - EDEN 124 (13)

43.6

1.4

KBSS - Surface 6

40.5

3.7

LDSS - EDEN 124 (14)

36.6

7.2

KBSS - Surface 7

41.1

3.3

LDSS - EDEN 124 (15)

40.5

2.3

KBSS - EAW 156 (25)

42.9

2.1

LDSS - EDEN 124 (16)

37.9

3.8

KBSS - EAW 156 (26)

41.4

2.1

LDSS - EDEN 125 (15)

39.8

5.2

KBSS - EAW 156 (27)

40.3

4.2

LDSS - EDEN 125 (16)

20.4

20.2

KBSS - EDEN 124 (23)

15.3

29.7

LDSS - EDEN 125 (17)

40

3.8

KBSS - EDEN 124 (24)

32.8

9.7

LDSS - EDEN 126 (16)

44.2

0.8

KBSS - EDEN 124 (25)

35.4

6.6

LDSS - EDEN 126 (17)

38.8

KBSS - EDEN 125 (21)

44.2

0.8

LDSS - EDEN 126 (18)

43.1

1.9

KBSS - EDEN 125 (22)

2.7

38.3

LDSS - EDEN 127 (15)

42.6

2.4

KBSS - EDEN 126 (23)

24.6

20.4

LDSS - EDEN 127 (16)

42

2.3

KBSS - EDEN 126 (24)

38.9

5.5

LDSS - EDEN 127 (17)

41

1.4

KBSS - EDEN 126 (25)

42.4

1.9

LDSS - EDEN 127 (18)

37.5

1.6

KBSS - EDEN 127 (22)

33.3

11.7

LRSS - Surface 1

38.3

6.7

KBSS - EDEN 127 (23)

36.9

7.8

LRSS - Surface 3

44.2

0.8

KBSS - EDEN 127 (24)

41.2

3.8

LRSS - Surface 4

43.8

1.2

ACFM - Surface 1

35

10

LRSS - EAW 156 (18)

40.7

4.3

ACFM - Surface 3

40.8

4.2

LRSS - EAW 156 (19)

22.4

19.9

ACFM - EDEN 124 (28)

14.9

30.1

LRSS - EAW 156 (20)

30.6

12.3

ACFM - EDEN 126 (28)

43.6

1.4

LRSS - EAW 30 (33)

41.7

3.2

DFSS - EDEN 124 (30)

10.7

34.3

LRSS - EAW 30 (34)

35.5

8.8

DFSS - EDEN 125 (27)

45

LRSS - EAW 30 (35)

38.3

6.7

DFSS - EDEN 125 (28)

45

LRSS - EDEN 124 (18)

6.6

38.4

DFSS - EDEN 126 (36)

22.5

22.5

LRSS - EDEN 124 (19)

40.6

4.4

DFSS - EDEN 126 (29)

22.5

22.5

LRSS - EDEN 125 (37)

23.8

9.8

DFSS - EDEN 127 (29)

45

LRSS - EDEN 125 (19)

33.3

9.9

DFSS - EDEN 127 (30)

35.4

9.6

LRSS - EDEN 126 (19)

40.1

4.9

BGCS - EDEN 127 (33)

10

35

LRSS - EDEN 126 (20)

33.2

11.8

WTFM - EDEN 124 (35)

32

13

LRSS - EDEN 126 (21)

17.5

27.5

WTFM - EDEN 126 (34)

30.3

14.7

LRSS - EDEN 127 (19)

42.3

2.7

WTFM - EDEN 127 (35)

30

15

LRSS - EDEN 127 (20)

27.2

17.8

UNM2 - EAW 30 (51)

26.6

18.4

UNM3 - EAW 156 (22)

32.3

12.7

UNM3 - EDEN 126 (22)

32.3

12.7

777

Appendix 5-30: Values of compaction porosity loss (COPL) and


cementation porosity loss (CEPL) in the Narrabeen Group.
Samples

COPL

CEPL

Samples

COPL

CEPL

NPFM - EAW 42 (2)


NPFM - EAW 42 (4)
NPFM - EAW 42 (6)
BACS - EAW 42 (14)
BGSS - EAW 42 (18)
BGSS - EAW 42 (19)
BGSS - EAW 42 (21)
BGSS - EAW 42 (22)
BGSS - EAW 42 (23)
BGSS - EAW 42 (24)
BGSS - EAW 42 (25)
BGSS - EAW 42 (27)
BGSS - EAW 42 (29)
BGSS - EAW 42 (30)
BGSS - EAW 42 (31)
BGSS - EAW 42 (32)
SPCS - EAW 42 (34)
SPCS - EAW 42 (35)
SBSS - EAW 42 (37)
SBSS - EAW 42 (38)
SBSS - EAW 42 (39)
SBSS - EAW 42 (42)
SBSS - EAW 42 (+42)
SBSS - EAW 42 (43)
SBSS - EAW 42 (44)
SBSS - EAW 42 (45)
SBSS - EAW 42 (+45)
SBSS - EAW 30 (1)
SBSS - EAW 30 (2)
SBSS - EAW 30 (4)
SBSS - EAW 30 (5)
SBSS - EAW 30 (6)
SBSS - EAW 30 (7)
SBSS - EDEN 125 (1)
SBSS - EDEN 125 (2)
SBSS - EDEN 125 (3)
SBSS - EDEN 125 (4)
WBCS - EAW 156 (1)
WBCS - EAW 156 (5)
WBCS - EAW 156 (6)
WBCS - EAW 42 (47)
WBCS - EAW 42 (48)
WBCS - EAW 42 (+48)
WBCS - EAW 42 (49)
WBCS - EAW 30 (9)
WBCS - EAW 30 (10)
WBCS - EAW 30 (11)
WBCS - EAW 30 (12)
WBCS - EAW 30 (14)
WBCS - EAW 30 (15)
WBCS - EAW 30 (16)

40.1
39
38.1
28.8
35.2
29.8
34.1
31.5
28.7
25.9
27.7
25.9
37.4
31.3
30.6
31.9
26.2
20.9
27.1
21.5
32
31.1
37
32.1
35.9
29.2
30.8
26.5
18.8
30
30.7
31.3
28.9
37.3
38.2
34
34.9
40.8
37.5
40
37.2
33.3
35.6
34.1
38.1
32.9
23
38.9
28.5
39.2
32.3

3.4
6
6.9
15.1
2.1
4.9
6.2
8.1
6.4
12.2
5.1
6.3
5
4.5
2.9
6.1
16.6
15.8
11
8.2
8.7
11.2
8
10.5
5.5
11.6
4.5
14.9
22.6
12.4
14.3
13.7
5
6.5
4.2
4.4
4
1.5
7.4
5
7.5
7.8
9.4
10.9
6.9
12.1
13.8
6.1
7.6
5.6
6.5

WBCS - EAW 30 (17)


WBCS - EAW 30 (18)
WBCS - EAW 30 (20)
WBCS - EAW 30 (21)
WBCS - EDEN 124 (2)
WBCS - EDEN 124 (3)
WBCS - EDEN 124 (4)
WBCS - EDEN 124 (5)
WBCS - EDEN 124 (6)
WBCS - EDEN 124 (7)
WBCS - EDEN 125 (6)
WBCS - EDEN 125 (7)
WBCS - EDEN 125 (8)
WBCS - EDEN 125 (10)
WBCS - EDEN 126 (38)
WBCS - EDEN 126 (3)
WBCS - EDEN 126 (4)
WBCS - EDEN 126 (6)
WBCS - EDEN 126 (7)
WBCS - EDEN 126 (8)
WBCS - EDEN 126 (10)
WBCS - EDEN 126 (12)
WBCS - EDEN 127 (2)
WBCS - EDEN 127 (3)
WBCS - EDEN 127 (4)
WBCS - EDEN 127 (5)
WBCS - EDEN 127 (7)
WBCS - EDEN 127 (8)
WBCS - EDEN 127 (10)
CCSS - Surface 1
CCSS - Surface 2
CCSS - Surface 3
CCSS - EAW 156 (12)
CCSS - EAW 156 (13)
CCSS - EAW 30 (22)
CCSS - EAW 30 (23)
CCSS - EAW 30 (24)
CCSS - EAW 30 (25)
CCSS - EAW 30 (26)
CCSS - EAW 30 (54)
CCSS - EDEN 124 (9)
CCSS - EDEN 124 (10)
CCSS - EDEN 124 (11)
CCSS - EDEN 125 (11)
CCSS - EDEN 125 (12)
CCSS - EDEN 125 (14)
CCSS - EDEN 127 (11)
CCSS - EDEN 127 (12)
CCSS - EDEN 127 (13)

36.3
25.3
41.8
41.1
40.5
37.8
40
36.9
33.1
35.6
31.3
25.3
27.9
23.9
35.8
27.5
23.1
30.1
26.3
35.9
36.8
15.6
36.4
37.9
29.8
28.3
37.6
37.3
0
41.7
41.2
41.4
30.6
34.5
21.9
35
18.8
33.2
39.7
41.4
35.4
36.4
28.8
34.4
0
40.2
26.2
38.9
41.6

6
19.7
3.2
3.9
4.5
7.2
4.9
4.3
5.8
6.3
11.9
12.9
9.2
12.3
9.2
8.6
11.5
10.7
14.3
9.1
8.2
29.4
8.6
5.3
15.2
12.8
5.2
7.7
45
2.7
3.8
2.6
3.3
2.2
14.4
10
18.9
4.8
3.5
3.5
9.6
8.6
13.8
5.3
45
4.8
14.2
2.6
2.9

777

Appendix 5-31: Values of compaction porosity loss (COPL) and


cementation porosity loss (CEPL) in the Hawkesbury Sandstone.
Samples

COPL

CEPL

HBSS - Surface 6
HBSS - Surface 7
HBSS - Surface 8
HBSS - Surface 9
HBSS - Surface 10
HBSS - Surface 11
HBSS - Surface 12
HBSS - EAW 18 a (1)
HBSS - EAW 18 a (2)
HBSS - EAW 18 a (3)
HBSS - EAW 18 a (4)
HBSS - EAW 18 a (5)
HBSS - EAW 18 a (7)
HBSS - EAW 18 a (8)
HBSS - EAW 18 a (9)
HBSS - EAW 18 a (10)
HBSS - EDEN 115 (1)
HBSS - EDEN 115 (2)
HBSS - EDEN 115 (3)
HBSS - EDEN 115 (4)
HBSS - EDEN 115 (5)
HBSS - EDEN 115 (6)
HBSS - EDEN 115 (7)
HBSS - EDEN 115 (8)
HBSS - EDEN 115 (9)
HBSS - EDEN 115 (10)
HBSS - EDEN 115 (11)
HBSS - EDEN 115 (12)
HBSS - EDEN 115 (13)
HBSS - EDEN 115 (14)
HBSS - EDEN 115 (15)
HBSS - EDEN 115 (16)

27.1
32.9
37.4
33.7
25.4
39.8
32.8
26.4
35.1
35.8
27.4
29.7
28.5
37.9
37.1
39.6
26.8
40.2
25.4
34.4
40.9
24.7
32.2
24.9
36.6
23.1
31.5
23.5
27.3
25.5
34.4
35.6

4.7
4.4
2.2
3.8
5.2
4.3
10.1
18.6
6
9.2
5.2
6.2
5.8
7.1
5.7
5.3
5.3
4.8
6.2
4.4
4.1
11.7
12.8
6.5
6
10.6
13.5
9.3
14.6
7.2
7.9
9.4

777

Appendix Six

Appendix 6-1:
a) Monocrystalline quartz grain (Qm), quartz overgrowth (Qo) and lithic
grains (white arrows) in the Lawrence Sandstone (Sample N. 19, EDEN
125).
b) Monocrystalline quartz grain (Qm), sedimentary rock fragment (SRF) and
chert (Ch) in the Coalcliff Sandstone (Sample N. 22, EAW 30).
c) Monocrystalline quartz grains (Qm) and kaolinite (Ka) in the Hawkesbury
Sandstone (Sample N. 3, EAW 18 a).

111

Appendix 6-2:
a) Mineral inclusion (white arrows) in monocrystalline quartz grain (Qm).
Chert (Ch) grain is partially coated by siderite (S) cement. Rock fragment
is observed. No porosity can be seen in this sample from the Lawrence
Sandstone (Sample N. 37, EDEN 125).
b) Mineral inclusions (red arrows) in quartz grain. Siderite (white arrow)
cement occurs as grain coating on quartz grain. Some lithic grains are
present. No porosity can be seen in this sample from the Wombarra
Claystone (Sample N. 10, EAW 30).
c) Mineral inclusion (Min) and concavo -convex contacts (white arrow).
Dissolution (diss) on margins of monocrystalline quartz grain (Qm). Some
porosity (P) occurs in the Hawkesbury Sandstone (Sample N. 7, EAW 18
a).
111

Appendix 6-3:
a) Polycrystalline quartz grain (Qp) and sedimentary rock fragment (SRF)
are coated by ankerite (A) cement in the Kembla Sandstone (Sample N.
22, EDEN 127).
b) Polycrystalline quartz grain (Qp), chert (Ch), sedimentary and volcanic
rock fragments (SRF, VRF) are present. Siderite (white arrow) cement
fills pore and coats grain coating. Also, blue dye indicates porosity in the
Scarborough Sandstone (Sample N. 38, EAW 42).
c) Polycrystalline quartz grain (Qp) with some porosity (white arrows) in the
Hawkesbury Sandstone (Sample N. 6, EDEN 115).

111

Appendix 6-4:
a) Rounded and pitted of recycled sedimentary quartz (white arrow) is
observed. Also, carbonate cements (siderite S, and calcite Ca) coat margins
of grain. No porosity can be seen in this sample from the Kembla
Sandstone (Sample N. 23, EDEN 124).
b) Fibrous to feathery texture of chalcedony (white arrow), monocrystalline
quartz (Qm), chert (Ch) and porosity (P) are present with pore-filling
siderite (S) in the Bald Hill Claystone (Sample N. 14, EAW 42).
c) Ankerite (A) cement fills pore and coats monocrystalline quartz (Qm),
microcline (Mi) and chert (Ch) grains. Also, polycrystalline quartz grain
(Qp) and volcanic rock fragment (VRF) are observed. No porosity can be
seen in this sample from the Kembla Sandstone (Sample N. 27, EAW 156).
111

Appendix 6-5:
a) Plagioclase grain (Plag) with pseudoplastic deformation of mud intraclasts
within pseudomatrix (white arrow). Also, sedimentary rock fragment
(SRF) and chert (Ch) grains in the Loddon Sandstone (Sample N. 17,
EDEN 125).
b) Plagioclase grain (Plag), volcanic rock fragment (VRF), chert (Ch) and
quartz grain (Q) are coated partly by calcite cement in the Wombarra
Claystone (Sample N. 8, EDEN 125).
c) Kaolinite (Ka) is present in an altered plagioclase grain (red arrow) in the
Hawkesbury Sandstone (Sample N. 5, EAW 18 a).

121

Appendix 6-6:
a) Detrital grains such as quartz (Q), feldspar (white arrow) and rock fragments
(RF) are partly coated by finely crystalline calcite cement. No porosity is
observed in this sample from the DarkesForest Sandstone (Sample N. 28,
EDEN 125).
b) Quartz (Q), feldspar (F) and chert (Ch) are partly coated by micritic ankerite
(A) cement in the Wombarra Claystone (Sample N. 8, EDEN 127).
c) Volcanic rock fragments (VRF), chert (Ch) and quartz (Q) grains are partly
coated by ankerite (A) cement in the Kembla Sandstone (Sample N. 22, EDEN
125).

121

Appendix 6-7:
a) Volcanic rock fragment (VRF) and monocrystalline quartz (Qm) grains are
partly coated by siderite (S) cement. Also, blue dye indicates secondary
porosity in the Bulgo Sandstone (Sample N. 31, EAW 42).
b) Ankerite (A) cement fills pores and occurs as a grain-coating on a
sedimentary rock fragment (SRF). Other rock fragments (RF) also occur in
the Loddon Sandstone (Sample N. 14, EDEN 124).
c) Sedimentary rock fragment (SRF), quartz (Q) grains and porosity (white
arrow) exist in the Bulgo Sandstone (Sample N. 1, EAW 42).

122

Appendix 6-8:
a) Sedimentary rock fragment (SRF) and monocrystalline quartz (Qm) grains
are associated with secondary porosity (white arrow) in the Hawkesbury
Sandstone (Sample N. 1, EDEN 115).
b) Monocrystalline quartz (Qm), chert (Ch) and rock fragments (VRF, SRF)
grains are partially coated by ankerite cement which replaced detrital grain
(red arrow). No porosity occurs in this sample from the Kembla Sandstone
(Sample N. 22, EDEN 125).
c) Chert (Ch), sedimentary rock fragments (SRF), monocrystalline quartz
(Qm) and pores (white arrows) are observed in the Scarborough Sandstone
(Sample N. 32, EAW 42).

121

Appendix 6-9:
a) Chert (Ch), monocrystalline quartz (Qm) grains and porosity (P) in the
Hawkesbury Sandstone (Sample N. 6, EDEN 115).
b) Deformation of muscovite (Mu) between quartz grains, indicating
mechanical compaction. Some quartz and lithic grains (white arrows) are
observed in the Loddon Sandstone (Sample N. 13, EDEN 124).
c) Deformation of muscovite grains (red arrows) occur between quartz grains
indicating mechanical compaction in the Wombarra Claystone (Sample N.
47, EAW 42).

121

Appendix 6-10:
a) Muscovite grain is bent between monocrystalline quartz (Qm) grains.
Some porosity (white arrow) is observed in the Hawkesbury Sandstone
(Sample N. 15, EDEN 115).
b) Zircon is present (red arrow). Also, detrital grains are coated by siderite
cement in the Darkes Forest Sandstone (Sample N. 27, EDEN 125).
c) Zircon (red arrow) occurs as mineral inclusion in the monocrystalline
quartz (Qm) grain. Also, porosity (white arrow) in the Hawkesbury
Sandstone (Sample N. 8, EDEN 115).

121

Appendix 6-11:
a) Quartz cement (Qo) is present as overgrowth around quartz grain (Q) and
is partly coated by ankerite cement (A). Margins of very fine volcanic rock
fragment (VRF) and sedimentary rock fragment (SRF) are coated by
ankerite cement (A) in the Loddon Sandstone (Sample N. 16, EDEN 125).
b) Double quartz overgrowth (Qo), feldspar (F), sedimentary rock fragment
(SRF) and chert (white arrows) occur in the Darkes Forest Sandstone
(Sample N. 36, EDEN 126).
c) Poikilotopic siderite (S) is developed in large pore and coats margins of
monocrystalline quartz grain (Qm), sedimentary rock fragment (SRF) and
chert (Ch) grains. Altered feldspar (white arrow) is observed. No porosity
can be seen in this sample from the Loddon Sandstone (Sample N. 32,
EAW 30).
121

Appendix 6-12:
a) Poikilotopic siderite (S) is coating a volcanic rock fragment (VRF).
Polycrystalline quartz (Qp) in the Loddon Sandstone (Sample N. 16,
EDEN 124).
b) Siderite (S) is observed as replacement of monocrystalline quartz (Qm).
Sedimentary rock fragment (SRF) and some lithic grains (white arrows) in
the Lawrence Sandstone (Sample N. 4, Surface Samples).
c) Feldspar grain (white arrow) is replaced by siderite cement. Polycrystalline
quartz (Qp) and chert (Ch) grains in the Lawrence Sandstone (Sample N. 4,
Surface Samples).

121

Appendix 6-13:
a) Finely crystalline siderite (S) cement fills pores between grains and is
present as a grain-coating on detrital grains. Also, siderite (white arrow)
occurs as replacement of detrital grain. Chert (Ch) and volcanic rock
fragment (VRF) grains are observed in the Lawrence Sandstone (Sample
N. 34, EAW 30).
b) Siderite (S) cement fills small pores between grains and occurs as a graincoating on detrital grains in the Darkes Forest Sandstone (Sample N. 27,
EDEN 125).
c) Small dolomite crystals (D) fill pore space and coat the margins of
volcanic and sedimentary rock fragments (VRF, SRF) in the Loddon
Sandstone (Sample N. 16, EDEN 125).
121

Appendix 6-14:
a) Small crystals of ankerite filling the pore and coating the margins of
detrital grains of quartz (Q) and rock fragments (RF). No porosity can be
observed in this sample from the Lawrence Sandstone (Sample N. 18,
EDEN 124).
b) Detrital grains are partly coated by large crystals of ankerite (A) which also
fills pores. No porosity can be observed in this sample from the Loddon
Sandstone (Sample N. 16, EDEN 125).
c) Ankerite cement is developed in pores spaces and exists as a grain-coating
on detrital grains such as feldspar (white arrow), quartz (Q) and lithic
grains (red arrows). No porosity can be observed in this sample from the
Lawrence Sandstone (Sample N. 18, EDEN 124).
121

Appendix 6-15:
a) Detrital grains and quartz overgrowths (white arrow) are partly coated by large
crystal of ankerite (A) in the Kembla Sandstone (Sample N. 23, EDEN 124).
b) Ankerite cement replacing a rock fragment (white arrow) and margins of
volcanic rock fragment (VRF) and chert (Ch) grains in the Loddon Sandstone
(Sample N. 16, EDEN 125).
c) Ankerite cement replacing a quartz grain (white arrow). Margins of volcanic
rock fragment (VRF), sedimentary rock fragment (SRF) and chert (Ch) are
coated by ankerite cement in the Kembla Sandstone (Sample N. 23, EDEN
124).

111

Appendix 6-16:
a) Poikilotopic calcite cement (Ca) occludes a large pore and exists as a coating
on the margins of detrital grains such as volcanic rock fragment (VRF),
sedimentary rock fragment (SRF) and chert (Ch). No porosity is observed in
this sample from the Kembla Sandstone (Sample N. 5, Surface Samples).
b) Detrital grains such as quartz (Q), feldspar (white arrow), rock fragment and
chert (Ch) are partly coated by small crystals of calcite cement which fill
pores. No porosity is observed in this sample from the Darkes Forest
Sandstone (Sample N. 28, EDEN 125).
c) Detrital quartz (white arrow) is replaced by calcite cement. Chert grain (Ch) is
observed. No porosity is observed in this sample from the Loddon Sandstone
(Sample N. 14, EDEN 124).
111

Appendix 6-17:
a) Calcite cement (white arrows) has replaced a detrital chert (Ch) grain. A
sedimentary rock fragment (SRF) is present in the Kembla Sandstone (Sample
N. 5, Surface Samples).
b) Quartz overgrowth (Qo) is partly coated by late diagenetic calcite (Ca) cement
which also occurs as coatings on detrital grains such quartz (Q) and chert (Ch)
in the Kembla Sandstone (Sample N. 23, EDEN 124).
c) Dolomite (white arrow) is present as a replacement of a chert (Ch) grain. Also,
chert (Ch) grains are present in the Loddon Sandstone (Sample N. 16, EDEN
125).

112

Appendix 6-18:
a) Small pores are filled by dolomite cement which coats the margins of detrital
quartz (Q), chert (Ch) and volcanic rock fragment (VRF) grains. Sedimentary
rock fragments (SRF) also occur in the Kembla Sandstone (Sample N. 23,
EDEN 124).
b) Pore-filling authigenic pyrite (py), lithic grains (white arrow), chert (Ch) and
porosity occur in the Kembla Sandstone (Sample N. 25, EDEN 126).
c) Iron oxide (white arrow) fills a pore and coats margins of quartz (Q). Siderite
(S) occurs as coatings on chert grain (Ch) in the Lawrence Sandstone (Sample
N. 34, EAW 30).

111

Appendix 6-19:
a) Concavo-convex grain contacts (red arrow) are observed. Quartz (Q) and lithic
grains (red arrows) are shown in the Loddon Sandstone (Sample N. 30, EAW
30).
b) Long grain contacts (white arrow). Siderite (S) grain-coating on detrital quartz
(Q). Quartz overgrowth (Qo) on detrital quartz. Chert grain (Ch) in the Loddon
Sandstone (Sample N. 31, EAW 30).
c) Monocrystalline quartz (Qm) is dissolved to form secondary porosity (black
arrow). Secondary porosity (white arrow) is also formed by ankerite
dissolution. Margins of monocrystalline quartz (Qm), chert (Ch) and
sedimentary rock fragment (SRF) are coated by ankerite cement (A) in the
Kembla Sandstone (Sample N. 22, EDEN 125).
111

Appendix 6-20:
a) Secondary porosity (white arrows) result from dissolution of siderite (S)
cement. Sedimentary rock fragment (SRF) and chert (Ch) occur in the
Loddon Sandstone (Sample N. 18, EDEN 127).
b) Late diagenetic ankerite cement occurs following dissolution of early
ankerite cement. Lithic grains (white arrow) occur in the Kembla
Sandstone (Sample N. 25, EDEN 125).
c) Authigenic feldspar occurs as overgrowths (white arrow) on detrital
feldspar grains (F). Quartz (Q), chert (Ch) and sedimentary rock fragment
(SRF) grains are observed. Ankerite (A) cement fills a small pore and coats
a sedimentary rock fragment (SRF) in the Loddon Sandstone (Sample N.
16, EDEN 125).
111

Appendix 6-21:
a) Feldspar dissolution results secondary porosity (white arrow). Lithic grain (red
arrow) occurs in the Loddon Sandstone (Sample N. 16, EDEN 127).
b) Secondary porosity led to the dissolution of a rock fragment (white arrow).
Some lithic grains (red arrows) are observed in the Kembla Sandstone (Sample
N. 25, EDEN 124).
c) Quartz overgrowth (Qo) on detrital quartz grains fills pore space. Some lithic
grains (white arrows) in the Wombarra Shale (Sample N. 3, EDEN 126).

111

Appendix 6-22:
a) Double quartz overgrowth (red arrow), chert (Ch) and sedimentary rock
fragment (SRF) grains occur in the Wombarra Shale (Sample N. 7, EDEN
127).
b) Poikilotopic crystals of siderite (S) cement fill a large pore and coats some
detrital grains in the Scarborough Sandstone (Sample N. 7, EAW 30).
c) Coarsely crystalline siderite cement (S) fills pores and occurs as graincoatings around the margins of polycrystalline quartz (Qp) and volcanic
rock fragments (VRF) in the Scarborough Sandstone (Sample N. 42, EAW
42).

111

Appendix 6-23:
a) Margins of chert (Ch), monocrystalline quartz (Qm) and sedimentary rock
fragments (SRF) are coated by small rhombohedral crystals of siderite (S)
cement in the Wombarra Claystone (Sample N. 5, EDEN 127).
b) Chert (Ch) and sedimentary rock fragment (SRF) are partly coated by early
siderite (S) cement. Quartz (Q) and porosity (white arrow) occurs in the
Scarborough Sandstone (Sample N. 38, EAW 42).
c) Siderite (S) cement occurs during early diagenesis as a pore-filling cement and
as grain-coatings around the margins of detrital quartz (Q) and chert (Ch)
grains. Also, a volcanic rock fragment (VRF) occurs in the Scarborough
Sandstone (Sample N. 1, EAW 30).
111

Appendix 6-24:
a) Late diagenetic siderite (S) cement precipitated on quartz overgrowth
(Qo), monocrystalline quartz (Qm) and volcanic rock fragment (VRF).
Chert grain (Ch) is observed. No porosity is present in this sample from
the Scarborough Sandstone (Sample N. 4, EAW 30).
b) Plagioclase (Plag) is replaced by siderite (S) cement. Chert grain (Ch)
and sedimentary and volcanic rock fragments (SRF, VRF) are observed
in the Coalcliff Sandstone (Sample N. 3, Surface samples).
c) Siderite (S) cement exists as a replacement of rock fragments (white
arrow). Quartz (Q), lithic grain (red arrow) and porosity (blue dye) are
observed in the Bulgo Sandstone (Sample N. 19, EAW 42).

111

Appendix 6-25:
a) Coarse crystalline of ankerite (A) cement is observed as pore-fillings and
as grain-coatings on the margins of quartz (Qm, Qp) and rock fragments
(VRF, SRF) in the Stanwell Park Claystone (Sample N. 35, EAW 42).
b) Quartz (Q), feldspar (F) and chert (Ch) are partly coated by micritic
ankerite (A) cement in the Wombarra Claystone (Sample N. 8, EDEN
127).
c) Microcrystalline ankerite (A) cement fills a large pore and coats the
margins of quartz (Q) and rock fragments (RF) in the Wombarra
Claystone (Sample N. 10, EDEN 125).

111

Appendix 6-26:
a) Quartz overgrowth (Qo) and quartz (Q) grains are coated by
coarse crystalline ankerite (A) cement which also has replaced a
volcanic rock fragment (white arrow) in the Coalcliff Sandstone
(Sample N. 11, EDEN 125).
b) Volcanic rock fragment (VRF) is replaced by ankerite cement (red
arrow) which also partly coats the margins of monocrystalline
quartz (Qm) and chert (Ch) grains in the Coalcliff Sandstone
(Sample N. 22, EAW 30).
c) Quartz overgrowth (Qo), monocrystalline quartz (Qm), chert (Ch)
and rock fragment (VRF, SRF) grains are coated by
microcrystalline dolomite (D) cement which also fills pores in the
Wombarra Shale (Sample N. 6, EDEN 124).
111

Appendix 6-27:
a) Early coarsely crystalline calcite (Ca) cement was precipitated around
detrital quartz (Q) and replaced a detrital grain (white arrow). Margins of
chert (Ch), volcanic and sedimentary rock fragments (VRF, SRF) are coated
by calcite cement (Ca) in the Scarborough Sandstone (Sample N. 4, EAW
30).
b) Microcrystalline calcite cement coats margins of rock fragments (red
arrows) and quartz (white arrow) and occurs as replacement of detrital
grains in the Wombarra Claystone (Sample N. 4, EDEN 127).
c) Polycrystalline quartz (red arrow) is replaced by calcite cement (white
arrow) in the Coalcliff Sandstone (Sample N. 2, Surface sample).
112

Appendix 6-28:
a) Calcite cement (red arrow) appears as a replacement of feldspar (white
arrow). Monocrystalline quartz (Qm) and chert (Ch) grains in the Coalcliff
Sandstone (Sample N. 1, Surface sample).
b) Rock fragment is completely replaced by calcite cement in the Coalcliff
Sandstone (Sample N. 2, Surface sample).
c) Detrital grains (white arrow) are replaced by calcite (Ca) cement. Some lithic
grains (red arrow) are also partly coated by calcite cement in the Coalcliff
Sandstone (Sample N. 11, EDEN 127).

111

Appendix 6-29:
a) Chert (Ch) grain is partly replaced by dolomite (D) cement.
Polycrystalline quartz (Qp) grain and volcanic rock fragment (VRF)
are coated by dolomite (D) cement in the Wombarra Claystone
(Sample N. 18, EAW 30).
b) Margins of monocrystalline quartz (Qm) and volcanic rock fragment
(VRF) are coated by dolomite (D) cement which also occurs as a porefilling. Polycrystalline quartz (Qp) occurs in the Coalcliff Sandstone
(Sample N. 24, EAW 30).
c) Dolomite (D) cement occurs as a grain-coating on monocrystalline
quartz (Qm). Quartz overgrowths (Qo) and volcanic rock fragments
(VRF) are present in the Wombarra Claystone (Sample N. 7, EDEN
125).
111

Appendix 6-30:
a) Authigenic pyrite (white arrow). Detrital grains such as
monocrystalline quartz (Qm) and chert (Ch) are partially coated by
dolomite (D) cement which also has replaced a chert grain (Ch) in the
Wombarra Shale (Sample N. 18, EAW 30).
b) Iron oxide (white arrow) fills a pore and coats the margins of quartz
(Q) and sedimentary rock fragments (SRF). Chert grain is observed in
the Coalcliff Sandstone (Sample N. 22, EAW 30).
c) Deformation of muscovite (Mu) between quartz grains indicates
mechanical compaction. Some lithic grains (white arrows) occur in the
Bulgo Sandstone (Sample N. 22, EAW 42).

111

Appendix 6-31:
a) Sutured quartz (Qm) grain contact (white arrow). Siderite (S) cement
occurs as grain-coating on monocrystalline quartz (Qm) grains. No
porosity can be observed in this sample from the Wombarra Claystone
(Sample N. 11, EAW 30).
b) Silica dissolution (white arrows). Dolomite (D) cement coats the margins
of quartz (Q) grains in the Wombarra Shale (Sample N. 8, EDEN 125).
c) Feldspar albitization (white arrows), chert (Ch) grain and secondary
porosity (P2) occurs in the Bulgo Sandstone (Sample N. 19, EAW 42).

111

Appendix 6-32:
a) Altered feldspar (white arrow) plus carbonate cement that occurs as a
coating on detrital grains such chert (Ch) and volcanic rock fragment
(VRF) in the Coalcliff Sandstone (Sample N. 24, EAW 30).
b) Secondary porosity (white arrow) occurs within a partially dissolved lithic
grain. Margins of detrital grains such as quartz (Q) and chert (Ch) are
coated by dolomite cement in the Coalcliff Sandstone (Sample N. 24,
EAW 30).
c) Dissolution of carbonate (white arrow) is recorded, forming late
diagenetic secondary pore. Monocrystalline quartz (Qm), chert and
volcanic rock fragment (VRF) grains exist in the Coalcliff Sandstone
(Sample N. 25, EAW 30).
111

Appendix 6-33:
a) Dissolution of plagioclase (plag) happened inside the grain, generating
minor secondary porosity (white arrow) and is associated with
replacement by siderite (S) cement which occurs as grain-coatings on
margins of the sedimentary rock fragments (SRF) in the Wombarra
Claystone (Sample N. 5, EDEN 127).
b) Secondary porosity (white arrows) result from dissolution of siderite (S)
cement. Margins of sedimentary rock fragment (SRF) and chert (Ch)
grains are coated by siderite (S) cement in the Coalcliff Sandstone
(Sample N. 11, EDEN 125).
c) Secondary porosity (white arrows) result from dissolution of dolomite (D)
cement. Margins of quartz (Q), polycrystalline quartz (Qp), volcanic rock
fragment (VRF) and chert (Ch) grains are coated by dolomite (D) cement
in the Wombarra Shale (Sample N. 14, EAW 30).
111

Appendix 6-34:
a) Well developed quartz overgrowths (Qo) occur as a pore-filling cement along
the margins of detrital quartz grain and obliterates chemical compaction,
preserving small primary porosity (white arrow) in the Hawkesbury
Sandstone (Sample N. 1, EDEN 115).
b) Margin of monocrystalline guartz (Qm) grain is coated by fine-grained
rhombs of siderite (S) cement. Porosity (white arrow) is seen in the
Hawkesbury Sandstone (Sample N. 13, EDEN 115).
c) Siderite (S) cement occurs as fine and medium rhombs coating the margins of
monocrystalline quartz (Qm) grains. Silica dissolution (diss) and porosity
(white arrow) are observed (Sample N. 13, EDEN 115).

111

Appendix 6-35:
a) Medium-grained rhombs of siderite (S) exist as pore-filling cement between
quartz (Q) grains. Chert (red arrow) grain is observed in the Hawkesbury
Sandstone (Sample N. 16, EDEN 115).
b) Coarse crystalline siderite cement fills a large pore and coats the margin of
quartz grains in the Hawkesbury Sandstone (Sample N. 16, EDEN 115).
c) Siderite cement fills a small pore in the Hawkesbury Sandstone. Also,
porosity (white arrow) is observed (Sample N. 7, EAW 18 a).

111

Appendix 6-36:
a) Pseudoplastic deformation of mud intraclasts within pseudomatrix (black
arrow). Ankerite (A) occurs as a pore-filling cement (red arrow) between
monocrystalline quartz (Qm) grains. Secondary porosity (white arrow) is
formed by chert dissolution in the Hawkesbury Sandstone (Sample N. 4,
EAW 18 a).
b) Poikilotopic ankerite cement occurs as a grain-coating on detrital quartz (Q)
and quartz overgrowths (white arrow) in the Hawkesbury Sandstone (Sample
N. 1, EAW 18 a).
c) Iron oxide cement (red arrow) fills a pore between monocrystalline quartz
(Qm) grains in the Hawkesbury Sandstone (Sample N. 6, EDEN 115).
111

Appendix 6-37:
a) Grain contacts show chemical compaction and/or pressure solution in the
Hawkesbury Sandstone (Sample N. 2, EAW 18 a).
b) Long sutured grain contact (red arrow) in the Hawkesbury Sandstone (Sample
N. 2, EDEN 115).
c) Bent muscovite (Mu) grain indicates mechanical compaction. A sedimentary
rock fragment (SRF) occurs in the Hawkesbury Sandstone (Sample N. 3,
EAW 18 a).

112

Appendix 6-38:
a) Early stage clay (Cl) and late diagenetic quartz overgrowths (Qo) in the
Hawkesbury Sandstone. Also, porosity (white arrow) is observed in this
unit (Sample N. 6, EDEN 115).
b) Late diagenetic siderite (S) cement grows on a quartz overgrowth (Qo)
and monocrystalline quartz (Qm). Secondary porosity is observed in the
Hawkesbury Sandstone (Sample N. 13, EDEN 115).
c) Secondary porosity (white arrow) is formed by dissolution of silica.
Primary porosity (red arrow) also occurs in the Hawkesbury Sandstone
(Sample N. 4, EAW 18a).

111

Appendix 6-39:
a) Secondary porosity (white arrow) results from dissolution of siderite (S)
cement. Margins of quartz grains are coated by siderite (S) cement or
quartz overgrowths (Qo) in the Hawkesbury Sandstone (Sample N. 4,
EAW 18).
b) The dissolution of carbonate fragments between quartz (Qm) grains in
Hawkesbury Sandstone (Sample N. 7, EAW 18).
c) Detrital grains such as monocrystalline quartz (Qm) and chert (Ch) are
coated by siderite (S) cement. Also, a detrital grain (red arrow) is replaced
by siderite cement. Low porosity (white arrow) occurs due to incomplete
pore-filling siderite cement. Also, sedimentary rock fragment (SRF) is
observed in the Lawrence Sandstone (Sample N. 19, EAW 156).
111

Appendix 6-40:
a) Detrital grains such as monocrystalline quartz (Qm), polycrystalline
quartz (Qp) and chert (Ch) are coated by siderite (S) cement. Intergranular
secondary porosity (white arrow) is formed by dissolution of pore-filling
siderite cement in the Lawrence Sandstone (Sample N. 19, EAW 156).
b) Monocrystalline quartz (Qm) and chert (Ch) grains are present. Volcanic
rock fragments are partly coated by calcite (Ca) cement and are partly
dissolved, forming intragranular secondary porosity (white arrow) in the
Loddon Sandstone (Sample N. 16, EDEN 127).
c) Grain-coating ankerite cement on detrital grains (white arrow) and
polycrystalline quartz (Qp), volcanic and sedimentary rock fragments
(VRF, SRF). Intragranular secondary porosity (blue arrow) occurred as
result of dissolution of a volcanic rock fragment in the Kembla Sandstone
(Sample N. 24, EDEN 124).
111

Appendix 6-41:
a) Detrital grains such as monocrystalline quartz (Qm) and sedimentary rock
fragments are observed. Intragranular secondary porosity (white arrow) is
formed by dissolution of volcanic rock fragment in the Kembla Sandstone
(Sample N. 25, EDEN 124).
b) Polycrystalline quartz grains (Qp) are replaced by siderite cement (red
arrow). Sedimentary rock fragments (SRF) and chert (Ch) are present.
Mouldic pore is seen as result of dissolved plagioclase in the Loddon
Sandstone (Sample N. 17, EDEN 125).
c) Feldspar is totally dissolved, forming oversized porosity (white arrow).
Volcanic and sedimentary rock fragments (VRF, SRF) occur in the
Loddon Sandstone (Sample N. 14, EDEN 124).
111

Appendix 6-42:
a) Primary pore (white arrow) is observed between monocrystalline quartz
grains (Qm) which are coated by siderite (S) cement in the Bulgo
Sandstone (Sample N. 24, EAW 42).
b) Primary pores occur between grains in the Bulgo Sandstone (Sample N.
24, EAW 42).
c) Volcanic and sedimentary rocks fragments are partly coated by ankerite
cement which is dissolved, forming intergranular secondary porosity
(white arrow) in the Coalcliff Sandstone (Sample N. 24, EAW 30).

111

Appendix 6-43:
a) Intergranular secondary porosity (white arrow) occurred due to dissolution
of pore-filling ankerite cement. Monocrystalline quartz (Qm), volcanic,
sedimentary rocks fragments (VRF, SRF) and chert (Ch) are partly coated
by ankerite cement in the Wombarra Claystone (Sample N. 7, EDEN
125).
b) Monocrystalline quartz (Qm) is partly coated by ankerite cement, which
dissolved forming intergranular secondary porosity (white arrows). Also,
volcanic and sedimentary rocks fragments (VRF, SRF) are present in the
Wombarra Claystone (Sample N. 14, EAW 30).
c) Siderite (S) cement occurs as a grain-coating on volcanic rock fragments
(VRF). Dissolution of a volcanic rock fragment lead to the formation of
intragranular secondary porosity (white arrow) in the Stanwell Park
Claystone (Sample N. 34, EAW 42).
111

Appendix 6-44:
a) Volcanic rock fragments (VRF) are partly coated by ankerite (A) cement.
Intragranular secondary porosity (white arrows) is observed due to
dissolution of a volcanic rock fragment in the Stanwell Park Claystone
(Sample N. 35, EAW 42).
b) Siderite (S) cement occurs as grain-coatings on the margins of
monocrystalline quartz (Qm), chert (Ch) and sedimentary rock fragments
(SRF). Partial dissolution of a feldspar grain (white arrow) results mouldic
porosity in the Coalcliff Sandstone (Sample N. 13, EDEN 127).
c) Monocrystalline quartz (Qm), chert (Ch) and volcanic rock fragment
(VRF) grains are clear. Also, intragranular secondary porosity and
oversized porosity (white arrow) are present as result of dissolution of
chert and feldspar grains, respectively, in the Coalcliff Sandstone (Sample
N. 12, EAW 156).
111

Appendix 6-45:
a) Feldspar grain is largely dissolved, forming oversized porosity (white
arrow). Margins of detrital grains such as monocrystalline quartz (Qm),
volcanic and sedimentary rock fragments (VRF, SRF) are coated by
ankerite (A) cement. Also a chert (Ch) grain is observed in the Wombarra
Claystone (Sample N. 7, EDEN 127).
b) Primary porosity (white arrow) occurs between monocrystalline quartz
(Qm) grains in the Hawkesbury Sandstone (Sample N. 7, EAW 18a).
c) Primary porosity occurs between quartz grains in the Hawkesbury
Sandstone (Sample N. 6, EDEN 115).

111

Appendix 6-46:
a) Intergranular pressure dissolution of quartz grains. Siderite (S) cement fills
pores between quartz grains, A chert grain (Ch) is observed, Primary and
secondary porosity (P1 and P2) occur in the Bulgo Sandstone (Sample N. 21,
EAW 42).
b) Pseudomatrix (white arrow) and secondary porosity (P2) are present as well
as siderite (S) cement, which occurs as grain-coatings on monocrystalline
quartz (Qm) in the Bulgo Sandstone (Sample N. 23, EAW 42).
c) Secondary porosity (white arrow) occurs as result of dissolution of carbonate
cement in the Newport Formation (Sample N. 2, EAW 42).

111

Appendix 6-47:
a) Kaolinite (Ka) is generated from alteration and dissolution of feldspar (F),
lithic grains (white arrows) and quartz (Q) are observed. Siderite (S) cement
is present as a grain-coating on detrital grain, Calcite (Ca) cement occurs as
grain-coating on altered feldspar in the Loddon Sandstone (Sample N. 32,
EAW 30).
b) Kaolinite (Ka) is generated from dissolution of feldspar (F), chert grains
(white arrows) and quartz (Q) are observed in the Coalcliff Sandstone
(Sample N. 11, EDEN 125).
c) Microporosity (white arrow) results from kaolinitization of feldspar. Also,
volcanic rock fragments (VRF) are observed in the Kembla Sandstone
(Sample N. 25, EDEN 124).

112

Appendix 6-48:
a) Partial dissolution of feldspar grain (white arrow) results in mouldic porosity.
Lithic grains (red arrows) are present in the Loddon Sandstone (Sample N.
16, EDEN 124).
b) Clays, probably illite or chlorite (white arrow), are formed as a result of
dissolution of feldspar. Also, alteration of feldspar (F), monocrystalline quartz
(Qm) grains and chert (Ch) are observed in the Loddon Sandstone (Sample N.
32, EAW 30).
c) Monocrystalline quartz (Qm) grains, siderite (S) cement and pseudomatrix
(white arrow) in the Hawkesbury Sandstone (Sample N. 9, EAW 18a).

111

Appendix 6-49:
a) The interpenetration of grains such as lithic grains (red arrows) and quartz
grains is observed. No porosity can be seen in this sample from the Kembla
Sandstone (Sample N. 24, EDEN 127).
b) Secondary porosity generated by the dissolution of polycrystalline quartz may
be filled by carbonate cement (white arrow). Also, volcanic rock fragments
(VRF) and chert (Ch) grains occur in the Loddon Sandstone (Sample N. 17,
EDEN 126).
c) Deformation of lithic grain (white arrow), chert (Ch) grains, ankerite (A)
cement and alteration of feldspar (F) are present. Also, no porosity can be
seen in this sample from the Scarborough Sandstone (Sample N. 42, EAW
42).

111

Appendix 6-50:
a) The interpenetration of grains such as lithic grains and quartz grains in the
Coalcliff Sandstone (Sample N. 10, EDEN 124).
b) Secondary porosity (white arrow) occurs as a result of incomplete pore-filling
by ankerite (A) cement. Also, monocrystalline quartz (Qm) grains,
sedimentary rock fragment (SRF) and chert (Ch) grains in the Scarborough
Sandstone (Sample N. 38, EAW 42).
c) Secondary porosity (white arrow) occurs as result of carbonate grain
dissolution. Also, secondary porosity (P2), quartz (Q) and chert (Ch) grains
occur in the Scarborough Sandstone (Sample N. 38, EAW 42).

111

Appendix 6-51:
a) Secondary porosity (white arrow) is formed by dissolution of lithic grains.
Also, Monocrystalline quartz (Qm), chert (Ch) and primary porosity (P1) are
present in the Bulgo Sandstone (Sample N. 19, EAW 42).
b) Secondary pores, generated by dissolution of feldspars (F) may be filled by
calcite (Ca) cement. Also, quartz (Q), chert (Ch) and volcanic rock fragments
(VRF) occur in the Coalcliff Sandstone (Sample N. 22, EAW 30).
c) Quartz overgrowth (Qo), sedimentary rock fragments (SRF), chert (Ch), and
primary and secondary porosity (P1, P2) occur in the Bulgo Sandstone
(Sample N. 23, EAW 42).

111

Appendix Seven

Appendix 7-1:
a) Scattered platelets of late diagenetic chlorite clays (Chl) post-date quartz
overgrowth (Qo) and preserves primary pore (red arrow) in the Loddon
Sandstone (Sample N. 16, EDEN 127).
b) Grain coating clays on detrital grains with the observation of primary pore
(red arrow) in the Lawrence Sandstone (Sample N. 35, EAW 30).
c) Chlorite clays (Chl) fill a large pore in the Kembla Sandstone. A primary
pore (red arrow) is also observed in the Kembla Sandstone (Sample N. 24,
EDEN 124).
d) Space pore is filled by mixed-layer illite/smectite in the Loddon Sandstone.
Also, primary pores (red arrows) are present (Sample N. 31, EAW 30).
e) Very tightly packed quartz in the Loddon Sandstone (Sample N. 13, EDEN
124).
f) Albitized feldspar grain (red arrow) with secondary porosity (P2) in the
Kembla Sandstone (Sample N. 23, EDEN 127).

761

Appendix 7-2:
a) Quartz overgrowth (Qo) fills pore and is partly coated by late diagenetic illite
clays (white arrow), thus porosities (red arrows) are preserved in the Loddon
Sandstone (Sample N. 31, EAW 30).
b) Euhedral crystals with the rhomohedron shape of ankerite (A) is present as a
pore-filling cement and is also coated by clays. Minor primary porosity (red
arrows) is observed in the Kembla Sandstone (Sample N. 23, EDEN 127).
c) Well developed ankerite cement (A) is partly coated by late mixed-layer
illite/smectite (Mix) in the Kembla Sandstone. Also, primary pore (red arrow)
occurs in the Kembla Sandstone (Sample N. 24, EDEN 126).
d) Kaolinite (Ka) occurs as booklets and vermicular aggregates and is associated
with dickite in the Kembla Sandstone (Sample N. 22, EDEN 125).
e) Late diagenetic mixed-layer illite/smectite (Mix) coats quartz overgrowth (Qo)
and preserves porosity (red arrow) in the Loddon Sandstone (Sample N. 15,
EDEN 124).
f) Quartz grain (Q) partly is covered by mixed-layer illite/smectite (Mix). Large
pore (red arrow) is observed in the Kembla Sandstone (Sample N. 23, EDEN
127).
761

Appendix 7-3:
a) Early stage quartz overgrowth (Qo) is coated by later diagenetic chlorite (chl)
which preserves porosity (red arrow) in the Loddon Sandstone (Sample N. 16,
EDEN 125).
b) Chlorite (Chl) occurs as a grain-coating on kaolinite (Ka). Small pores (red
arrows) are present in the Kembla Sandstone (Sample N. 22, EDEN 125).
c) Authigenic kaolinite (Ka) is intergrown with authigenic illite (ill). Also,
primary pores (red arrows) are recorded in the Kembla Sandstone (Sample N.
22, EDEN 125).
d) Mixed layer illite/smectite (Mix) is observed as a grain coating on albitized
feldspar (F). Secondary pores (red arrows) are present in the Kembla
Sandstone (Sample N. 24, EDEN 124).
e) Late diagenetic mixed-layer illite/smectite (Mix) is precipitated on siderite
cement (S) and quartz overgrowth (Qo). Porosities (red arrows) are recorded in
the Loddon Sandstone (Sample N. 18, EDEN 127).
f) Illite pore-filling (ill) cement in the Loddon Sandstone. Also, primary pores
(red arrows) are observed (Sample N. 19, EDEN 125).
761

Appendix 7-4:
a) Detrital grain (white arrow) is coated by ultra-thin layers and thin mat-like
authigenic illite. Also, primary pores (red arrows) occur in the Loddon
Sandstone (Sample N. 16, EDEN 127).
b) Ultra-thin layers and thin mat-like grain coating illite bordering pores in the
Lawrence Sandstone. Primary pore (red arrow) is observed (Sample N. 34,
EAW 30).
c) Quartz overgrowth (Qo) is enclosed by chlorite (Chl), thus porosities (red
arrows) are present in the Kembla Sandstone (Sample N. 24, EDEN 124).
d) Chlorite occurs as a replacement of detrital grains in the Loddon Sandstone
with the observation of a secondary pore (red arrow) (Sample N. 15, EDEN
124).
e) Chlorite occurs as replacement of detrital grains in the Loddon Sandstone
(Sample N. 16, EDEN 124).
f) Chlorite occurs as replacement of detrital grains in the Loddon Sandstone
(Sample N. 37, EDEN 125).

711

Appendix 7-5:
a) Authigenic kaolinite (Ka) has filled a pore and is partly coated by fibrous illite
(ill), also primary porosity (P1) occurs in the Newport Formation (Sample N.
2, EAW 42).
b) Pore-filling authigenic kaolinite with primary pore (red arrow) in the Bulgo
Sandstone (Sample N. 21, EAW 42).
c) Siderite cement (S) fills a pore and is coated by late diagenetic mixed layer
illite/smectite (Mix) in the Bulgo Sandstone (Sample N. 21, EAW 42).
d) Detrital feldspar (F) in the Bulgo Sandstone and is coated partly by mixed
layer illite/smectite (Mix) (Sample N. 27, EAW 42).
e) Quartz overgrowth (Qo) is partly coated by late chlorite (chl), thus large
porosity (Po) is observed in the Bulgo Sandstone (Sample N. 21, EAW 42).
f) Late diagenetic mixed layer illite/smectite (Mix) is present as coatings around
quartz overgrowth (Qo) and authigenic kaolinite (Ka). Also, primary pore (red
arrow) occurs in the Wombarra Claystone (Sample N. 6, EDEN 124).

717

Appendix 7-6:
a) Authigenic kaolinite (Ka) occurs as grain-coatings around a large pore with
secondary porosity (P2) in the Bulgo Sandstone (Sample N. 18, EAW 42).
b) Siderite cement (S) fills a pore and is partly coated by late diagenetic mixedlayer illite/smectite. Also, primary pores (red arrows) are observed in the
Scarborough Sandstone (Sample N. 3, EDEN 125).
c) Quartz overgrowth (Qo) is partly coated by late mixed-layer illite/smectite,
thus porosity (red arrow) is preserved in the Scarborough Sandstone (Sample
N. 1, EAW 30).
d) Mixed-layer illite/smectite (Mix) occurs as a pore-filling cement in the
Scarborough Sandstone. Also, secondary pores (red arrows) are present
(Sample N. 45, EAW 42).
e) Ultra-thin layers and thin mat-like crystals of illite (ill) exists as a graincoating on kaolinite (Ka) which fills a large pore. Small pores (red arrow)
occurs in the Newport Formation (Sample N. 2, EAW 42).
f) Fibrous authigenic illite (ill) occurs as pore-filling in the in the Scarborough
Sandstone (Sample N. 38, EAW 42).
711

Appendix 7-7:
a) Late stage of ultra-thin layers and thin mat-like crystals of illite (ill) occurs as a
grain-coating on quartz overgrowths (Qo); primary porosity (P1) is present in
the Bulgo Sandstone (Sample N. 27, EAW 42).
b) Large pore is filled by chlorite (Chl) in the Coalcliff Sandstone (Sample N. 12,
EAW 156).
c) Authigenic chlorite (Chl) is coating on detrital grains and is associated with
mixed-layer illite/smectite (Mix). Primary pore (red arrow) is observed in the
Coalcliff Sandstone (Sample N. 11, EDEN 124).
d) Authigenic chlorite (Chl) is associated with fibrous illite (ill). Secondary pores
(red arrows) occur in the Newport Formation (Sample N. 2, EAW 42).
e) Detrital feldspar (F) is associated with authigenic kaolinite (Ka). Also,
secondary pores (red arrows) are present in the Scarborough Sandstone
(Sample N. 2, EDEN 125).
f) Quartz overgrowth (Qo) occurs in the Wombarra Claystone (Sample N. 10,
EDEN 125).

711

Appendix 7-8:
a) Well developed kaolinite (Ka) exists as booklets and vermicular aggregates
associated with illite (ill). Also, microporosity is seen between kaolinite plates
in the Hawkesbury Sandstone (Sample N.1, EDEN 115).
b) Small pore between authigenic kaolinite (Ka), detrital feldspar (F) and primary
porosity (P1) in the Hawkesbury Sandstone (Sample N.4, EAW 18 a).
c) Fibrous illite (ill) occurs as a grain coating around kaolinite (Ka) and quartz
grains (Q). Kaolinite (Ka) fills pore between grains. Minor secondary porosity
(P2). Quartz overgrowth (Qo) is enclosed by authigenic kaolinite (Ka) in the
Hawkesbury Sandstone (Sample N.1, EDEN 115).
d) Kaolinite (Ka) is observed as a pore-filling cement and is partially coated by
ultra-thin layers and thin mat-like crystals of illite (ill) which also coats quartz
overgrowths (Qo). Minor primary porosity (Po) in the Hawkesbury Sandstone
(Sample N.1, EDEN 115).
e) Late stage quartz overgrowth (Qo) is coated by later diagenetic kaolinite (Ka)
in the Hawkesbury Sandstone (Sample N.7, EAW 18 a).
f) Margins of large pore are coated by kaolinite (Ka) in the Hawkesbury
Sandstone (Sample N.7, EAW 18 a).
711

Appendix 7-9:
a) Well developed kaolinite (Ka) exists as booklets and vermicular aggregates
and is coated with ultra-thin layers and thin mat-like crystals of illite (ill).
Primary pores (red arrows) occur in the Hawkesbury Sandstone (Sample N.4,
EAW 18a).
b) Kaolinite (Ka) is observed as pore-filling cement and is associated with ultrathin layers and thin mat-like crystals of illite (ill) in the Hawkesbury Sandstone
(Sample N.10, EDEN 115).
c) Fibrous illite (ill) fills pore space and in the Hawkesbury Sandstone. Also,
detrital feldspar (F) and primary pores (red arrows) are observed in the
Hawkesbury Sandstone (Sample N. 4, EAW 18).
d) Kaolinite (Ka) exists as booklets and vermicular aggregates and is associated
with illite (ill). Primary pores (red arrows) occur in the Hawkesbury Sandstone
(Sample N.6, EDEN 115).
e) Scattered crystals of chlorite (Ch) occur as rims on kaolinite (Ka) in the
Hawkesbury Sandstone (Sample N. 4, EDEN 115).
f) The conversion of kaolinite into dickite (white arrow) in the Hawkesbury
Sandstone (Sample N. 4, EDEN 115).
711

Appendix 7-10:
a) Late diagenetic kaolinite (Ka) overlies quartz overgrowth (Qo). Porosity (Po)
is preserved in the Hawkesbury Sandstone (Sample N.7, EAW 18 a).
b) Authigenic kaolinite (Ka) is precipitated on quartz overgrowth (Qo). Porosity
(Po) is preserved in the Hawkesbury Sandstone (Sample N.7, EAW 18 a).
c) Mixed-layer illite/smectite (Mix) coats onto authigenic kaolinite (Ka) with the
observation of a secondary pore (red arrow) in the Hawkesbury Sandstone
(Sample N.4, EAW 18 a).
d) Late diagenetic mixed-layer illite/smectite is associated with late diagenetic
illite (ill) which occurs as grain-coating on quartz overgrowth (Qo). Porosity
(red arrow) is preserved in the Hawkesbury Sandstone (Sample N.1, EDEN
115).
e) Illite (ill) occurs as a grain-coating on the margin of a quartz grain (Q) and
quartz overgrowth (Qo) in the Hawkesbury Sandstone (Sample N.4, EDEN
115).
f) Pore-filling chlorite (Chl) in the Hawkesbury Sandstone (Sample N.14, EDEN
115).

716

Appendix 7-11:
a) Late diagenetic chlorite (Chl) occurs as rims around quartz overgrowth (Qo) in
the Hawkesbury Sandstone. Also, primary pores (red arrows) are observed
(Sample N.14, EDEN 115).
b) Authigenic chlorite (Chl) in the Hawkesbury Sandstone (Sample N.3, EAW
18a).
c) Chlorite (Chl) is associated with kaolinite (Ka) in the Hawkesbury Sandstone
(Sample N.4, EDEN 115).
d) Kaolinite (Ka) occurs as a later diagenetic coating on quartz overgrowths (Qo)
and is associated with dickite. Also, primary pores (red arrows) are present in
the Bulgo Sandstone (Sample N.30, EAW 42).
e) Partially altered detrital feldspar (white arrows) in the Hawkesbury Sandstone
(Sample N.3, EAW 18).
f) Detrital muscovite (Mu) is shown by scanning electron microscope in the
Hawkesbury Sandstone (Sample N.3, EAW 18).

711

Appendix 7-12:
a) Microcrystalline quartz rims (white arrow) occur as a grain-coating on quartz
overgrowth, preserving porosity (Po). Kaolinite (Ka) and quartz overgrowth
are observed in the Hawkesbury Sandstone (Sample N. 6, EDEN 115).
b) Microcrystalline quartz rims (white arrow) occur as a grain-coating on quartz
overgrowths, preserving porosity (Po) and are associated with authigenic illite
(ill) in the Hawkesbury Sandstone (Sample N. 6, EDEN 115).
c) Microcrystalline quartz rims (white arrows) occur as a grain-coating on quartz
overgrowths, preserving porosity (Po) and are associated with kaolinite (Ka) in
the Hawkesbury Sandstone (Sample N. 6, EDEN 115).

711

Appendix Eight

450

HBSS

400
NPFM

Height above base of Wilton Fm (m)

350

BACS

300
BGSS

250

Tq
Qm

200

Qp

SPCS
SBSS

150
WBCS

100

50

0
0

20

40

60

80

Component percentage

Appendix 8-1: The variation of average values of total


quartz (Tq), monocrystalline and polycrystalline quartz (Qm,
Qp) with stratigraphic height based on point count data
(Appendix 5-14).

971

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

HBSS

500
450

NPFM

Height above base of Wilton Fm (m)

400

BACS

350
BGSS

300
Tq

250

Qm
200

Qp

SPCS
SBSS

WBCS

150

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

100
50
0
0

20

40

60

80

100

Components

Appendix 8-2: The variation in all subsurface samples of total quartz


(Tq), monocrystalline and polycrystalline quartz (Qm, Qp) with
stratigraphic height based on point count data (Appendices 5-7, 5-8
and 5-9).

981

450

HBSS

400
NPFM

350
Height above base of Wilton Fm (m)

BACS

300
BGSS

250
TF
K-f

200

SPCS

Plag

SBSS

150
WBCS

100

50

0
0.0

0.5

1.0

1.5

2.0

2.5

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

Component percentage

Appendix 8-3: The variation of average values of total feldspar


(TF), plagioclase (Plag) and K-feldspar (K-f) with stratigraphic
height based on point count data (Appendix 5-14).

989

HBSS

500
450

NPFM

Height above base of Wilton Fm (m)

400

BACS

350
BGSS

300
TF

250
200

K-f

SPCS

Plag

SBSS

WBCS

150

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

100
50
0
0

0.5

1.5

2.5

3.5

Component percentage

Appendix 8-4: The variation in all subsurface samples of total


feldspar (TF), plagioclase (Plag) and K-feldspar (K-f) with
stratigraphic height based on point count data (Appendices 5-7, 5-8
and 5-9).

981

450
HBSS

400
NPFM

350
Height above base of Wilton Fm (m)

BACS

300
BGSS

250
RF
200

Ch

SPCS
SBSS

150
WBCS
CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

100

50

0
0

10

20
30
40
Component percentage

50

60

Appendix 8-5: The variation of average values of rock


fragments (RF) and chert (Ch) with stratigraphic height based
on point count data (Appendix 5-14).

981

500

HBSS

450
NPFM

400

BACS

350

BGSS

300
RF

250

Ch

SPCS
SBSS

200

WBCS

150

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

100

50

0
0

10

20

30

40

50

60

70

Appendix 8-6: The variation in all subsurface samples of rock


fragments (RF) and chert (Ch) with stratigraphic height based on
point count data (Appendices 5-7, 5-8 and 5-9).

981

450
HBSS

400
NPFM

350
Height above base of Wilton Fm (m)

BACS

300
BGSS

250
HML
Mu

200

Chl

SPCS
SBSS

150

WBCS

100

50

0
0.0

0.5

1.0

1.5

2.0

2.5

Component percentage

Appendix 8-7: The variation of average values of heavy


minerals (HML), muscovite (Mu) and chlorite (Chl) with
stratigraphic height based on point count data (Appendix 514).

981

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

HBSS

500
450

NPFM

400
Height above base of Wilton Fm (m)

BACS

350
300

BGSS

HML

250
200

Mu

SPCS

Chl

SBSS

WBCS

150

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

100
50
0
0

10

12

Component percentage

Appendix 8-8: The variation in all subsurface samples of heavy


minerals (HML), muscovite (Mu) and chlorite (Chl) with
stratigraphic height based on point count data (Appendices 5-7, 5-8
and 5-9).

981

450

HBSS

400
NPFM

Height above base of Wilton Fm (m)

350

BACS

300
BGSS

250
Qo
Carb

200

SPCS

TC

SBSS

150
WBCS
CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

100

50

0
0

10

20
30
Component percentage

40

50

Appendix 8-9: The variation of average values of quartz


overgrowth (Qo), carbonate cement (Carb) and total cement
(TC) with stratigraphic height based on point count data
(Appendix 5-14).

987

HBSS

500
450

NPFM

Height above base of Wilton Fm (m)

400

BACS

350
BGSS

300
Qo

250

Carb
200

TC

SPCS
SBSS

WBCS

150

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

100
50
0
0

10

20

30

40

50

60

70

Component percentage

Appendix 8-10: The variation in all subsurface samples of quartz


overgrowth (Qo), carbonate cement (Carb) and total cement (TC)
with stratigraphic height based on point count data (Appendices 57, 5-8 and 5-9).

988

450
HBSS

400
NPFM

350
Height above base of Wilton Fm (m)

BACS

300
BGSS

250
Po
200

Ma

SPCS
SBSS

150
WBCS

100

50

0
0

10

20

30

40

50

Component percentage

Appendix 8-11: The variation of average values of total


porosity (Po) and matrix (Ma) with stratigraphic height based
on point count data (Appendix 5-14).

981

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

HBSS

500
450

NPFM

Height above base of Wilton Fm (m)

400

BACS

350
300

BGSS

250

Po
Ma

200

SPCS
SBSS

WBCS

150

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

100
50
0
0

20

40

60

80

100

Component percentage

Appendix 8-12: The variation in all subsurface samples of total


porosity (Po) and matrix (Ma) with stratigraphic height based on
point count data (Appendices 5-7, 5-8 and 5-9).

911

450

HBSS

400
NPFM

350

Height above base of Wilton Fm (m)

BACS

300
BGSS

250
GSZ

200

SPCS
SBSS

150
WBCS

100

50

0
0

100

200

300

400

500

600

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

Grain size (microns)

Appendix 8-13: The variation of average values of grain size


(GSZ) with stratigraphic height based on point count data
(Appendix 5-14).

919

HBSS

500
450

NPFM

Height above base of Wilton Fm (m)

400

BACS

350
BGSS

300
250
GSZ

SPCS
SBSS

200

WBCS

150

CCSS
LDSS
LRSS
UNM2
UNM3
WWCO
KBSS
ACFM
DFSS
BGCS

100
50
0
0

500

1000

1500

2000

Grain size (microns)

Appendix 8-14: The variation in all subsurface samples of grain


size (GSZ) with stratigraphic height based on point count data
(Appendices 5-7, 5-8 and 5-9).

911

450
HBSS

400
NPFM
GRFM
BACS

Height above base of Wilton Fm (m)

350

300
BGSS

250
Q
Cl

200

SPCS
SBSS

150

WBCS

100

50

0
0

20

40
Component percentage

60

80

Appendix 8-15: The variation of average values of quartz (Q)


and clay minerals (Cl) with stratigraphic height (XRD results;
Appendix 5-22).

911

CCSS
LDSS
BASM
LRSS
UNM2
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

HBSS

500
450

NPFM
GRFM
BACS

Height above base of Wilton Fm (m)

400
350

BGSS

300
250

Q
Cl

200

SPCS
SBSS

150

WBCS

100
50
0
0

20

40

60

80

100

Component percentage

Appendix 8-16: The variation in all subsurface samples of quartz (Q)


and clay minerals (Cl) with stratigraphic height (XRD results;
Appendices 5-15, 5-16 and 5-17).

911

CCSS
LDSS
BASM
LRSS
UNM2
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

450

HBSS

400
NPFM
GRFM
BACS

Height above base of Wilton Fm (m)

350

300
BGSS

250
TF
K-f

200

Plag

SPCS
SBSS

150
WBCS

100

50

0
0

10
Component percentage

15

20

CCSS
LDSS
BASM
LRSS
UNM2
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

Appendix 8-17: The variation of average values of total feldspar


(TF), K-feldspar (K-f) and plagioclase (Plag) with stratigraphic
height (XRD results; Appendix 5-22).

911

HBSS

500
450

NPFM
GRFM
BACS

Height above base of Wilton Fm (m)

400
350

BGSS

300
TF

250

K-f
200

Plag

150

SPCS
SBSS

WBCS

100
50
0
0

10

15

20

25

30

Component percentage

Appendix 8-18: The variation in all subsurface samples of total


feldspar (TF), K-feldspar (K-f) and plagioclase (Plag) with
stratigraphic height (XRD results; Appendices 5-15, 5-16 and 517).

911

CCSS
LDSS
BASM
LRSS
UNM2
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

450
HBSS

400
NPFM
GRFM
BACS

Height above base of Wilton Fm (m)

350

300
BGSS

250
HML
Mu
200

Carb

SPCS
SBSS

150
WBCS

100

50

0
0

10
15
Component percentage

20

25

CCSS
LDSS
BASM
LRSS
UNM2
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

Appendix 8-19: The variation of average values of heavy


minerals (HML), muscovite (Mu) and carbonate minerals (Carb)
with thickness (XRD results; Appendix 5-22).

917

HBSS

500
450

NPFM
GRFM
BACS

Height above base of Wilton Fm (m)

400
350

BGSS

300
HML

250
200

Mu

SPCS

Carb

SBSS

WBCS

150
100
50
0
0

10

20

30

40

50

60

70

Component percentage

Appendix 8-20: The variation in all subsurface samples of heavy


minerals (HML), muscovite (Mu) and carbonate minerals (Carb)
with thickness (XRD results; Appendices 5-15, 5-16 and 5-17).

918

CCSS
LDSS
BASM
LRSS
UNM2
WWCO
KBSS
ACSM
ACFM
DFSS
BGCS
TGSM
WTFM

You might also like