02 Whole
02 Whole
02 Whole
by
Meshack L. N. Kagya
B.Sc. (Dar es Salaam), M.Sc. (Newcastle'upon'Tyne)
Doctor of Philosophy
March 1997
Abstract vll
Declaration ix
Acknowledgements X
CHAPTER 1 INTRODUCTION I
1.1 Rationale 1
I
2.4.4 Hy drocarbon generation
2.5 Stable carbon isotope signature of rock extracts and crude oils
2.5.1 Occurrence and abundance of carbon
2.5.2 Living organisms and the carbon cycle
2.5.3 Photosynthesis and carbon isotope fractionation
2.5.3.1 Carbon isotopic fractionation in biosynthetic
processes
2.5.3.2 Carbon isotope record in sdiments
2.5.4 Carbon isotopic composition of hydrocarbon fractions
of crude oils
2.5.5 Normal alkane distributions and carbon isotopic signatures
2.6 Saturated hydrocarbon biomarkers
2.6.1 Normal alkanes
2.6.I.1 O dd- c arb on-numb ered n- alkane s of hi ghe r
molecular wei ght (C
rr-C rr)
2.6. I .2 O dd - c arb on- numb e red n - alkane s of me dium
molecular weight (C,rand C17)
1l
2.1 .2.4 Methylnaphthalene homolo gues 95
2.7 .3 Alteration of crude oils in reservoirs 96
ru
4.2.3.2 Detro-inertinite sub group 115
4.2.3.3 Gelo-inertinite subgroup 115
IV
6.2 Carbon isotopic composition of saturated and aromatic hydrocarbons
in source rocks and oils 182
6.2.1 Saturates in source rock extracts 185
6.2.2 Saturates in crude oils 186
6.2.3 Aromatics in source rock extracts 189
6.2.4 Atomatics in crude oils 189
6.2.5 Relationship between isotopic composition of saturated
and aromatic hydrocarbons 192
6.2.6 Isotopic signatures of oils in stacked reservoirs 195
6.3 Relationship of r¿-alkane distribution to source biota and saturate
õ13C signature 198
v
7.3.2.1 Terpanes 244
7.3.2.2 Steranes 245
7.3.3.I n-Alkanes 246
7.3.3.2 Pristane and phytane 248
7.3.4 OiI source affinity based on cyclic biomarkers 248
7.3.4.1 Terpanes 248
7.3.4.2 Steranes 248
7.4 Biomarker maturity parameters 250
7.4.1 Source rock maturity based on acyclic biomarker parameters 250
7.4.2 Source rock maturity based on cyclic biomarker parameters 250
7.4.2.I Terpanes 250
7.4.2.2 Steranes 25t
7.4.3 Oll maturity based on acyclic biomarkers 253
7.4.4 OiI maturity based on cyclic biomarkers 253
BIBLIOGRAPHY 292
VI
Abstract
Source rock evaluation based on Rock-Eval pyrolysis and organic petrography was
undertaken on silty shales, carbonaceous shales and coal samples from the Early Jurassic
fluviolacustrine Poolowanna Formation in oil fields located along the western edge of the
Eromanga Basin, South Australia. Potential source rock lithofacies were delineated using
gamma ray and sonic wireline logs. Their total organic carbon (TOC) values range from
I.5Vo in silty shales to70Vo in coal facies. Maceral analysis data show that vitrinite (notably
fluorescent desmocollinite) is the dominant maceral group (vitrinitùliptinite>inertinite). The
liptinites are mainly resinite and sporinite in the Poolowanna Trough .
In the Patchawarra
Trough, inertinite is the dominant maceral group (inertiniteäiptinite>vitrinite) and the
liptinite group comprises mainly Botryococcøs-like telalginite, indicative of lacustrine
depositional environments. Generally the whole study area seems to have been subject to
fluvial and paludal processes leading to a variety of sub-oxic to oxic terrestrial depositional
environments, as illustrated by a wide range of vitrinite to inertinite ratios (V/I = 0. 1 3- 1 3 .0).
Low V/I ratios suggest either intense oxidation of autochthonous humic organic matter prior
to burial, or a strong input of reworked allochthonous inertodetrinite. Both Rock-Eval and
organic petrographic data indicate the presence of oil-prone Type IIIIII (or, more rarely,
resinite-enriched Type II) kerogen. The maturity of source rock lithofacies range from early
mature (Ro = 0.5-O.6Vo) in the Patchawarra Trough to mature (Ro = O.9Vo) in the
Poolowanna Trough.
In order to ascertain whether the potential source rocks of the Poolowanna Formation have
actually contributed hydrocarbons to adjacent basin-edge reservoirs, f,rfteen representative
rock samples and eighteen oils from five oil fields (Poolowanna, Tantanna, Sturt, Sturt East
and Taloola) were selected for comparative isotopic and biomarker analysis. Three of these
fields produce from stacked reservoirs which range in age from Cambrian (Mooracoochie
Volcanics), through Permian (Patchawarra Fm), to Jurassic (Poolowanna Fm, Hutton Sst,
Birkhead Fm and Namur Sst). All the oils appear to be early expulsion products (R" = 0.5-
Aprimary higher plant input to both oils and source rocks is shown by the relative
O.\Vo).
abundances of acyclic isoprenoids and n-alkanes (Prln-Cs versus PWn-Cß) and the
presence of conifer resin-derived diterpenoid hydrocarbons. Intense microbial activity in the
depositional environments of the source rocks is indicated by high hopane/sterane ratios and
the presence of 2s"-, 2þ-, and 3B-methyl-17cr(H), 21B(H)-hopanes in the Czs-Czt
pseudohomologous series.
vll
Permian/Jurassic. The distinction of Jurassic from Permian and pre-Permian hydrocarbons is
substantiated by cross plots of calculated reflectance (Rc) versus aromatic isotopic signature
(õ13Caro) and methylphenanthrene index (MPI) versus trimetþlnaphthalene ratio (TNR-2).
In these plots the Poolowanna source rocks coincide with most of the oils in Jurassic
reservoirs. However, an unusually high abundance of 3O-norhopane (C2elC3s hopane -1)
was observed in the intra-Poolowanna shales and coals at Sturt, Sturt East and Tantanna, a
feature seen in none of the oils. The Poolowanna-l (Poolowanna) crude has a distinctively
light isotopic signature, analogous to that of one of the waxy Jurassic source facies in Sturt
Field. Sterane distributions are without exception ethylcholestane-dominant in both oils and
source rock extracts.
Vru
Declaration
This work contains no material which has been accepted for the award of any other degree or
diploma in any university or other tertiary institute and, to the best of my knowledge and
belief, contains no material previously published or written by another person, except where
due reference has been made in the text.
I give consent to this copy of my thesis, when deposited in the University Library, being
available for loan and photocopying.
Meshack L.
2lMarch1997
lx
Acknowledgements
I thank the South Australian Department of Mines and Energy (MESA) for their permission
to access the well completion reports and to obtain rock samples from their core library at
Glenside. The technical advice and assistance provided by Dr. David Gravestock, Ms. Elinor
Alexander and other MESA personnel are very much appreciated. I extend my thanks to Mr.
Ardy Mitchell of the National Centre for Petroleum and Geophysics and Mr. Rob Mempes
of the Department of Geology and Geophysics for their assistance in acquiring the wireline
log data. I do appreciate the hospitality and technical support by the staff at the MESA Core
Library. The oil samples were kindly provided by Santos Ltd and I am very indebted to
those personnel, particularly Mr. Andrew Pietch, who tried hard to make those samples
available to me.
My special thanks go to Geotechnical Services Pty Ltd. of Victoria Park, 'Western Australia
for carrying out the Rock-Eval pyrolysis and total organic carbon (TOC) analyses; Amdel
Ltd, Thebarton, for allowing me to use their facilities to undertake reflectance measurements;
and the Australian Wine Research Institute, Urrbrae, for access to their gas chromatography-
mass spectrometry (GC-MS) facility.
Most of the analytical work was carried out in house at the Department of Geology and
Geophysics, University of Adelaide. Accordingly, I am very much indebted to Mr. 'Wayne
Mussared for preparing the polished blocks for petrographic analysis; Dr. Keith Turnbull for
his cordial assistance with the carbon isotopic analysis; Ms Sherry Proferes for her frequent
guidance in computer drafting; Mr. John Stanley for his promptness in providing the X-ray
diffraction data and Mr. Bernd Michaelsen for running the Organic Geochemistry Laboratory
and his considerable help with the GC-MS analyses,
X
the Enfield Lions Club who have kindly supported me and my family in different ways
during this hard period of my studies. My sincere thanks to Ms. Pauline Burger for her
considerable help with editing.
Finally, my sincere thanks and appreciation go to my wife Anna and our children Asiimwe
and Alinda for being supportive and tolerant; and also to my family in Tanzania especially
my parents Mrs Amelia and the late Mr. Mathias Ndyekobora, for their constant guidance
and encouragement.
XI
Chapter 1
Introduction
1.1, Rationale
Recent additions to global oil and gas reserves have been discovered by the application of
increasingly sophisticated petroleum exploration techniques. Petroleum geoscientists in their
efforts to locate commercial accumulations of hydrocarbons, now ask specific questions
concerning source rocks and their related crude oils. Major issues are the occurrence and
distribution of source rocks in the basin, their potential to generate hydrocarbons, likely
migration pathwaysof the expelled hydrocarbons and finally entrapment mechanisms.
Moreover, knowledge of the geochemical composition of a petroleum sample provides
insights on its source rock characteristics (e.g. lithology, depositional environment, thermal
history, type of organic matter and age) and its post expulsion history, including the relative
migration distance, arrd in situ alteration by water washing or biodegradation and thermal
maturation.
In the Eromanga Basin of central Australia (Fig. 1.1), exploration for hydrocarbons
commenced in the early 1900's but the first discoveries of gas and oil were not made until
1976 and 1977 respectively. The origin of the subsequently discovered hydrocarbons in this
Mesozoic basin has been the subject of much discussion. Did they originate from within the
Eromanga Basin or were they derived from underlying Palaeozoic basins? For instance, the
oils which flowed from Early Jurassic reservoirs at Cuttapinie-l in the Patchawarra Trough
and Poolowanna-l in the Poolowanna Trough are reported to have shown features
attributable to either Triassic (Barr and Youngs, 1981) or Early Jurassic source rocks
(Thomas, 1982).
In order to solve this enigma the techniques of petroleum geochemistry were applied to the
study of source rocks within the Early Jurassic Poolowanna Formation, in Poolowanna
Trough where the first well (Poolowanna-1) to recover Jurassic oil was drilled; and in the
south western Patchawarra Trough where there are several oilfields with multiple stacked
reservoirs (Fig. 1.2). The oil bearing reservoirs in this part of the Patchawarra Trough
include the Cambrian Mooracoochie Volcanics of the 'Warburton Basin; the Permian
Patchawarra Formation of the Cooper Basin; and the Jurassic Poolowanna Formation,
Hutton Sandstone, Birkhead Formation and Namur Sandstone of the Eromanga Basin.
1
136' r40' 44
20 I
I
/ l^r
\ nl itr GALILEE
RGINA \
(-rt ta
BASIN
BASIN \.
l--- -,
\
\
/
24 /
\ SIMPSON -l / \/
I
\
PEDIRKA
DES
\ N
\
t
z'! BAS
\ / COOPER
N I
I \
I I BASIN
a
(
/ I
Moo
I WARBURTON \
I
'l B
BASI
)
W77Ã Poolowanna Trough 0
L+¡ 200
km
Figure 1. 1. The Eromanga Basin and its underlying basins. Cross section A-B is
illustrated in Figure 1.6.
Depth (m)
BSL 84 2 6 7 weils
- 1600
- 1640
\\
\-r
- 1800 Birkhead
Hutton Sandstone
- 1850
{
{
{ na
- 1900 { { (
{
a
{{{{
{
{
a
{
<(({
{ { { 1 I
{ { { t{
{ x
{l { { {<
{ { {t {
{
{ {
{
.l' {
{
{
{
{ {
{
{{ {
{
{ 4
- 1950 { { {
{ I {
{ { {
{ (
{ a
{ {
{
{
( {ta
(a
{
{
t
{ {
{ a
I { 1 a { a a a a { a {
{ { { { a 4 a{ { { {
1 { { I
{ { { {
{ 1
{ { {
1
{
{ {
{
{
{
{
{ I
{ {
a
{ { { {
{ {
{{
{{
{
I
{ { {
{ {
{ { {
{
(
{
{ { I
{
{
{ { t
{
a
I( { {
{ f
{ {
{{ t {
{ t
I { a{
- 2000
. oil < DST interval
Sandstone Coa
{{{{
{{{{
<((t
Volcanics
Figure 1.2a. Cross section of Sturt Field showing multiple stacked reservoirs (modified
from T\rrner (1991)).
o o
Nealyon 1 Tantanna Taloola Lake Hope 1
Murta
Birkhead
Hutton
Poolowanna
Nappamerri
Patch awalra Epsilon
Patchawarra
Merrimelia
2km
Figure 1.2b. Tantanna-Taloola cross section showing the occurrence of stacked Jurassic
reservoirs (Heath et a1.,1989)
Chapter I
1.2 Objectives and scope of this work
The aims of this study are to establish the presence of source rocks in the Poolowanna
Formation, to evaluate their hydrocarbon generative potential and to determine whether any
of the generated hydrocarbons have been expelled as crude oil. Source rock evaluation was
performed by employing the techniques of organic petrography; Rock-Eval pyrolysis;
column chromatography of rock extracts; and gas chromatography, stable carbon isotope
analysis and gas chromatography-mass spectrometry of saturated and aromatic
hydrocarbons. Using these methods, the organic matter content, kerogen type, source rock
richness, maturation levels, hydrocarbon yield and biomarker composition were determined.
Source rock depositional environments were also reconstructed. Comparison of the source
rock characteristics of the Poolowanna Formation in the Poolowanna and Patchawarra
Troughs is of particular interest.
The depth and thickness of drillhole sections of the Poolowanna Formation are shown in
Table 1.1 where the intervals which are likely to be good source rocks because of their
lithological composition (shale, siltstone and coal) are identified as potential source units.
Intervals sampled for source-rock analysis, together with their lithology and mineralogy (for
selected samples), are shown in Figure L4a-d and Table L2.The oil samples recovered
from drill stem tests (DST) of Jurassic and underlying reservoirs in fields from the study
afea are shown in Table 1.3, along with their API gravities and pour points.
4
a)
\. \ NT I
OLD
\ þil.l
t
!
I \\
I
T t
\ I
\
t\ l'
E
\ I \ SA
t
\\
J
t{ I
o I I
b \ POOLOWANNA I t
\ I I
J \ I
zÉ. Oodnadatta r \ , COOPER
¡
ul J{ / BASIN I
l-U) GH --/ I
i
u¡
7 STURT/ STURT EAST Moombd
tl
NA TALOOLA
a
I l\)
a/
I
a
I
!
NT
I z
CN
QLD i€
WA I
0 100 km
N r-l i
I
1 380 1400
b)
o 3
Or o6 a7
26.21'.
4
t'!,'.t' o(De.z a"
oo
Oe.r 3 28"1
o2 199"36' o 1
Figure 1.3 Study area showing a) oilfield locations and b) well locations in PELs 5 and 6
Chapter I
Table 1.1 Drillhole intersections of Poolowanna Formation in Poolowanna and Patchawarra
Troughs
6
Table 1.2 Lithological and mineralogical composition of Poolowanna Formation samples selected for source rock analysis
ray
No. (m) API Quartz Kaolinite Illite Gypsum
D = Dominant
SD = Sub dominant
M = Minor
T = Trace
Figure 1.4a. Poolowanna Formation sample locations in the Poolowanna Field
3 Poolowanna 2
Poolowanna Poolowanna 1
E
E
Þ z
z
3
c
c
co
2500
>
o
?
)
Þ
It
Þ
Ê
I
!r
Þ
:
¡
-- GR DT
261
GR DT
GR DT
Figure 1.4b. Poolowanna Formation sarnple localions in the Tantanna Field
I
GAP| GAPI GAPI
I
m m
786
- m m
792
796 1
q
I 1800
!
i
)
I 1800
\
! (
ì z
z
j
800 1
¡ d c=
1 ( õ
't cfL
E
J E
GR DT
i
E ¡
E
o 1822
181 I (
I
GR DT
GR DT 1830
-I
1
GR DT
G GR
GR DT
0 200 140 40 0
tt
200 40._10 g_J200 É0
1
10 g_J200 1110 4,0 Q_300 1Ég 4,0
GAPI pVft GAPI Usfft GAPI pvft GAPI tÆ/ft GAPI ps/ft
lr L
m m m m
1 1
1
I
J
ì
Il
GR
1
I
DT o
+ I
1
1860 B
E
1 860
B J
o
(
(
1 t
J o r 1
GR DT
1
E GR
å
DT 1
o
GR DT GR DT
+ DST interval O Oil sample Source rock sample
Figure 1.4d. Poolowanna Formation sample locations in tlre Sturt East Field.
Sturt East 4 Sturt East 2 Sturt East 1 Sturt East 3
0 200 140 40 0 200 0
t-,
200 140 0t-J 200 1
I I L
lj
2
2
c
e B
c 1840
o 1860
1840
+
o
854
Â
GR
GR DT 1884
868
Dr + uBourr"
GR DST interval O Oil sample B rock samptBr
Chapter )
Table 1.3. Oils from selected drillholes in the Poolowanna and Patchawarra Troughs
t3
FORMATIONS AGE
EROMANGA WINTON FM
BASIN L
(t,
MACKUNDAFM
f
OODNADATÏA FM ALLARU MUDSTONE o
I,JJ
o
BULLDOG
SHALE
E
I,JJ
WALLUMBILI-A FM
SUB
Í.
TRANSITION BEDS
(CADNA- OWE FM)
WYANDRA SST MBR.) o
MURTA FM
ALGEBUCKINA SST
NAMUR SST
+ L
ssT
RKHEAD FM
H SST
M IU'
+ U)
:::::f::::l::::l NA FM E É.
l
fE
I,TJ
CN
ã=-
LrJ
-u)
z- -¿
w t\^
M
L o
oò
U)
{ E
o-
6 f rooucxee rn
o- L
l
o
DABALINGIE BEDS É.
(5 COOPER ^¡
z BASIN
6 o_
MT. TOONDINA
FM. dt
J z
(5
z IJJ
(5 E
Ø o
z c0 FM (t É
cÍ. IJJ
o-
Y
o TIRRAWABRASST
fE
MEFRIMELIA FM
BOORTHANNA FM
ARROWIE CAR.
ICER BASI
AMADEUS RTON
BASIN BASIN .':':r:iiiiiiliiii M OO taCOOChi e
o O¡l * Gas
Figure 1.5 Generalised stratigraphy and significant hydrocarbon occrurences in the
western Eromanga Basin region (after Kantsler et a1.,1986). Stiple indicates reseryoir
units from which oils were analysed in this study.
Chapter I
1,.4.1 Stratigraphy and basin development
The generalised stratigraphy of the western Eromanga Basin and its intrabasins is illustrated
in Figure 1.5. During the Cambro-Ordovician period, sedimentation took place in the
Warburton, Arrowie, Amadeus and Georgina Basins (Gravestock, 1995). The clastics,
redbeds, carbonates and volcanics which occur beneath the westem Cooper Basin are of
Early to Late Cambrian age, whereas the dark grey pyritic siltstone and sandstone attributed
to the development of the Larapintine Sea, represent Ordovician sedimentation (Gatehouse,
1986). Low grade metamorphic rocks (argillites and arenites, probably of Silurian age),
intruded by late Silurian and Devonian granite, underlie the Cooper Basin in southern
Queensland (Green et aL.,1989).
Deposition in the Cooper and Pedirka Basins is reported to have been triggered by the decay
of an Early Permian ice sheet which released an enormous volume of sediment. A non-
marine to shallow matine, fluvio-lacustrine environment appears to have prevailed during
this period giving rise to a typical Gondwana basin succession comprising basal diamictites
overlain by non-marine and marine sediments (Gilby and Foster, 1988). The late
Carboniferous to Early Permian record of the Pedirka Basin consists of a basal diamictite
(Crown Point Formation) and the meandering fluvial deposits of the Purni Formation (Fig.
1.s).
The Cooper Basin is characterised by a major depocentre of Permo-Triassic age. It has a total
sediment thickness of up to 2000 m, comprising cycles of fluvial sandstone, fluvio-deltaic
coal measures, lacustrine shale and, in places, glacial sediments. This succession has been
assigned to two groups. The first of these is the Gidgealpa Group of Late Carboniferous to
Late Permian Age. At its base is the glacigenic Merrimelia Formation Q-ate Carboniferous).
This unit is unconformably overlain by an Early Permian sequence consisting of the braided
fluvio-deltaic and lacustrine Patchawarra Formation; the lacustrine Murtereee Shale; the
fluvio-deltaic Epsilon Formation; the lacustrine Roseneath Shale; and fluvio-deltaic
Daralingie Formation. Another major unconformity separates those Early Permian sediments
from the Late Permian fluvio-deltaic Toolachee Formation. The second is the Late Permian to
Early Triassic Nappameni Group which consists of the nonmarine Arrabury and Tinchoo
Formations. The Arrabury Formation contains the lacustrine, floodplain and meandering
stream sediments of the Callamuna Member; and the braided fluvial channel and floodplain
sediments of the Paning and the Wimma Sandstone Members. At the top the fluvio-lacustrine
Tinchoo Formation conformably overlies the Arrabury Formation.
The Simpson Desert Basin is represented by two Triassic units, viz. the Walkandi and Peera
Peera Formations. The Walkandi Formation is characterised by interbedded shale, siltstone
and minor sandstone, deposited on floodplains and in shallow ephemeral lakes. The Peera
Peera Formation conformably overlies the V/alkandi Formation and consists of carbonaceous
l4
Chapter I
in a fluvial environment. A slight but
shale, coal and thin sandstone interbeds, deposited
widespread compressional deformation, regional uplift and erosion at the end of the Late
Triassic appears to have terminated deposition in the region.
The Eromanga Basin contains about 3000 m of Jurassic to Cretaceous sediments which were
deposited in lacustrine, fluvial
peat swamp and shallow marine environments. The
lowermost unit is the Poolowanna Formation of South Australia (Moore, 1986) which is the
time equivalent of the 'Basal Jurassic' unit of southern Queensland (Geological Survey of
Queensland, 1985), the'Windorah Formation of the eastern Eromanga region and the upper
Precipice Sandstone and Evergreen Formation of the Surat Basin (Passmore and Burger,
1e86).
Other Jurassic units of the Eromanga Basin include the Hutton Sandstone, Birkhead
Formation, Adori Sandstone, Westbourne Formation, Namur Sandstone and Murta
Formation. Collectivel], these units are laterally equivalent to the Algebuckina Sandstone of
the Poolowanna Trough (Nugent, 1969). Of these, the Birkhead Formation is characterised
by interbeds of siltstone, coal and sandstone deposited in a flood-plain and lacustrine
environment; the Westbourne Formation consists of fluvial and overbank siltstone and
sandstone; and the Murta Formation comprises thinly interbedded lacustrine dark grey
siltstone, shale, and sandstone.
15
Chnpter I
Field, its thickness ranges from 14 m in Tantanna-l to 31m in Tantanna-2. These intervals
generally consist of siltstone and sandstone with minor coal interbeds. The depositional
environment seems to have varied from meandering or anastomosing fluvial to associated
flood plain lacustrine, overbank and low energy swamp settings.
Fig. 1.5 Stratigraphy
In the Sturt and Sturt East Fields, the Poolowanna Formation unconformably overlies the
Patchawarra Formation and ranges in thickness from 5m at Sturt-1 to 48m at Sturt East-4. It
contains siltstone, coal and sandstone interbeds. In some intervals the siltstones a.re notably
carbonaceous. Fluvio-lacustrine environments appear to have been prevalent, leading to
interdistributary siltstone interbedded with minor distributory channel sands.
Poolowanna Trough
The Poolowanna Trough is a large synclinal area which appears to be influenced by the
structural settings of both the Pedirka and Simpson Desert Basins. Its broad structural low
coincides with that of the underlying basins and is suggestive of a common depocentre. In
the western part of the Trough, a structural setting similar to that of Cooper Basin has been
suggested (Moore, 1986). Here, structural development in the Early Permian is marked by
small horsts and grabens, suggesting a tensional tectonic regime. Erosion and a large time-
break mark the separation of Permian, Pedirka Basin strata from the overlying Mesozoic
strata. In the central and eastern part of the Trough, the Permian sediments are absent and
there is very little sign of Permian or Mesozoic structural development. The main phase of
structuring appears to have occurred in the mid-Tertiary. It was associated with epeirogenic
movements which are verified by the presence of an elongate, anticlinal dome with strong
fault control on the western margin. The main faults extend upwards from the basement into
the Cretaceous Cadna-owie Formation, and possibly as far as the Toolebuc Formation
(Moore and Pitt, 1984).
Patchnwarra Trough
Deformation in this part of the Cooper Basin region has resulted in a group of anticlinal
structures (Tantanna, Taloola, Sturt and Sturt East). All these structures are associated with
pre-Permian palaeohighs and show evidence of continued growth from the Early Permian to
the Tertiary, including co-axial closures and isopach thinning (Anonymous, 1988; Beech,
198e).
t6
EROMANGA BASIN \-
A B
MSL Ku
M
J
-2000 J
\\\\\\
PEDIRKA \\\\\\
Ìttttt
-4000 BASIN
(volcanics)
tc-Pl
-6000 SIMPSON DESERT
BASIN COOPER BASIN
trrl c-T
Tr Triassic Ku Upper Cretaceous
P Permian Kl Lower Cretaceous
C Cambrian J Jurassic
Basement
Cooper Basin source rocks contain an average of 3.9Vo total organic carbon (TOC) and have
an average hydrocarbon yield of 6.9 kg/t (Jenkins, 1989). In the intraformational fluvial and
deltaic shales of the Toolachee and Patchawarra Formations TOC values range from 3 to 6Vo
(Kantsler et a1.,1983; Smyth, 1983; Cook and Struckmeyer, 1986). Abundant inertinite-rich
coals occur in the Patchawarra Formation of the Patchawarra Trough (Hunt, 1988) where
Type trI kerogen was identified in both coal and dispersed organic matter (DOM).
Better quality Type IVIII kerogen with hydrogen index values up to 320 mg HC/g TOC was
reported to be present in the Toolachee Formation of the Naccowlah Block, southwestern
Queensland (Vincent et al., 19S5). These high HI values may be caused by bacterial
degradation of terrigenous organic matter which is thought to improve the quality and oil
generation potential of the source rock (Kantsler et a1.,1983).In PELs 5 and 6 and ATP
259P, the Toolachee Formation is reported to have an average TOC content of '7.2Vo and a
hydrocarbon yield of 15.7 kg HC/t, whereas equivalent figures for the Patchawarra
Formation are somewhat lower (3.OVo TOC; 4.8 kg HCit) (Jenkins, 1989).
18
Chapter I
Source rocks of Jurassic to Early Cretaceous age from the Eromanga Basin, have an average
TOC content of l.6%o, and an average hydrocarbon yield of 4.3 kg/t (Jenkins, 1989). The
Poolowanna Formation in the Poolowanna Trough, contains up to I5Vo TOC and Type IVIII
kerogen (Cook, 1982). The equivalent 'Basal Jurassic' unit in the northern area of the
Cooper Basin has TOC values ranging from 0.5 to 2.OVo, whereas in the Naccowlah Block
TOC values range from 1 to lIVo (Vincent et al., 1985). Hydrogen index values ranging
from 150 to 450 mg HC/g TOC appear to be quite common confirming the presence of Type
IIIIII kerogen.
The Birkhead Formation in the northern Cooper Basin region has TOC values ranging from
2 to 4Vo, with a mean hydrocarbon generation potential of 4l kg/t in the non-coal facies
(Scholefied, 1989). Coals in this area were found to be vitrinite-rich (>507o) and inertinite
poor (<llVo). Their liptinite fraction has a high content of sporinite and cutinite and lesser
amount of resinite, suberinite and alginite. In the Naccowlah Block this unit is a good source
rock with of this Type II/I[ kerogen TOC values up to 4.57o (Yincent et ø/., 1985). The
hydrogen-rich liptinite component HI value in the range 150-450 mg HCig TOC, is thought
to have been enhanced by aerobic bacterial decay of vitrinite (Powell, 1984). Throughout
PELs 5 and 6 and ATP 259P, the Birkhead Formation is reported to have an overall average
TOC value of 2.5Vo and a mean hydrocarbon yield of 10.8 kg/t, with a preponderance of
vitrinite and liptinite contents varying from 10 to'|OVo of the DOM. The liptinite is mainly
cutinite and sporinite, with minor resinite and alginite (Jenkins, 1939).
The Murta Member has somewhat lower TOC values ranging up to 3.lVo and hydrogen
indices in the range 100 to 350 mg HClg ToC (Vincent et al., 1985; Jenkins, 19g9;
Michaelsen and McKirdy, 1989). Its kerogen is of Type IIIIII composition. Potential
hydrocarbon yields are typically in the range 2-8 kg/t. The DOM contains a high proportion
of inertinite and liptinite. Sporinite and liptodetrinite are the most abundant liptinites although
telalginite contents ap to 26Vo have been recorded in some samples from the Murteree Horst
(Michaelsen and McKirdy, 1989; Powell et a1.,1989).
t9
Chapter I
The saturated hydrocarbon fractions of source rock extracts from the Eromanga Basin
(Murta, and Birkhead Formations) and Cooper Basin (Toolachee and Patchawarra
Formations) exhibit variable wax (n-C2e*) contents and high pristane to phytane ratios
(prlpb3). High wax contents have been reported in the Westbourne and Birkhead
Formations, whereas the Murta Formation saturates are non-waxy (Powell et aL, 1989;
Gilby and Mortimore, 1989). Cooper Basin siltstones and shales have non-waxy n-alkane
distributions, whereas the associated coals and carbonaceous shales tend to have waxy
profiles (Jenkins, 1989).
Pristane to phytane ratios ranging from 3 to t have been recorded in the Murta Formation
whereas the Birkhead Formation had ratios in the range 4 to 5 (Vincent et a1.,1985). The
'Basal Jurassic' has pristane to phytane ratios up to 5. Steranes and triterpanes were
identified in almost all the aforementioned source rocks from the Cooper and Eromanga
Basins (Jenkins, 1989). An average hopane to sterane ratio of 2.5 was reported from both
basins, with occasional local variations indicating variable microbial influence on source rock
input (McKirdy, 1984).
20
COOH
R --R COOH
R=H, CH3 R=CH2OH, COOH R=CH2OH, COOH
I trm ry
AgathicAcid and Communols and Sandaracopima¡adienol Abietic Acid
Methylagathate CommunicAcids and SandaracopimaricAcid
X XI C 19 czo
CTg
1,2,5-Trimethyl- 1,7-Dimethyl- V VI Vtr
naphthalene phenanthrene 15,19-Bisnor- 19-Norlabdane Labdane
labdane
zMATURATION
Figure 1.7 Relationships between saturated and aromatic biomarkers and their conifer
resin precursors (after, Alexander et al., L988).
Chnpter I
1.5.3 Source rock maturity
Source rock maturities have been reported as vitrinite reflectance measurements (Kantsler
andCook, 1979; Cook 1982; Kantsler eta1.,1982; Passmore and Boreham, 1982; Kantsler
et aI., 1983) and in molecular parameters such as the metþlphenanthrene index (MPÐ
(Alexander et a1.,1988; Michaelsen and McKirdy, 1989; Boreham et
al., 1988; PowelI et
al., 1989; Tupper and Burckhardt, 1990); the dimetþlnaphthalene ratio (DNR-l) and
trimethylnaphthalene ratios (TNR-1 and TNR-2) (Alexander et a\.,1985; Radke et aI., 1986
Alexander et a1.,1988); and various triterpane and sterane isomer ratios (Mackenzie,1984;
Michaelsen and McKirdy, 1989; Powell et a1.,1989).
Vitrinite reflectance data show that source rocks in the Cooper and Eromanga Basins have
overlapping maturities ranging from 0.5 to l.ÙVo Ro, and indicative of the onset to main
stage of oil generation (Gilby and Mortimore, 1989; Powell et a\.,1989). However, at some
localities in the Cooper Basin source rocks have Ro values up 4.OVo which are ovemature
for oil generation and correspond to late-stage gas generation (Kantsler and Cook, 1979;
Kantsler et a1.,1983 and Hrurnt et al., 1989). However, immature source rocks occur on the
flanks of both basins (Scholefield, 1989).
In the Poolowanna Trough, the Poolowanna Formation is at the early mature stage (Cook,
1982; Moore, 1986) with vitrinite reflectance varying from 0.7 to 0.9Vo Ro. In the central
and southern Eromanga Basin, the maturity of this unit varied from 0.55 to 0.70% Ro
(Kantsler et aL, 1982; Passmore and Boreham, 1982; Scholefield, 1989). Along the
Naccowlah-Jackson trend and in the northern part of the basin, Vincent et aI. (1985) reported
maturities of 0.60-0.70Vo R.lo, whereas in the central Nappamerri Trough area a maturity
greater than O.8Vo Ro was infened for the Poolowanna Formation. In the northern
Naccowlah Block the Birkhead Formation is early mature to mature (0.65-O.8Vo Ro) while
the Murta Formation has an early oil generation matunty (0.45-O.6Vo Ro) shown from
isoreflectance mapsat the base of each formation (Vincent et al., 1985). In the
Nappacoongee-Murteree and Moomba Blocks a maturity range of 0.57-0.62Vo Ro was
reported for the Murta Formations (Michaelsen and McKirdy, 1989; Powell et al., 1989).
Alexander et al. (1988) observed that for samples of Jurassic age such calculated vitrinite
reflectance values are likely to be least reliable in the low maturity zone (Rc <O.7Vo) because
of anomalouslyhighconcentrations of the araucariacean marker l-methylphenanthrene. The
22
Chapter Ì
of this effect is for low maturity samples to have an artificially low Rc value. To
net result
overcome this problem, Alexander and coworkers developed several alternative maturity
parameters based upon the abundance of other isomeric aromatic hydrocarbons including
dimetþlnaphthalenes and trimetþlnaphthalenes. The first of these ratios (DNR-1) is defined
as [2,6-DMN + 2,7-DMN]/1,5-DMN where DMN is dimethylnaphthalene (Alexander et aL,
1985). The second ratio (TNR-l) is defined as 12,3,6-TMN/[1,4,6-TMN + 1,3,5-TMN]
where TMN is trimetþlnaphthalene (Alexander et al., 1985). A cross-plot of DNR-I and
TNR-I for samples of different maturities shows an approximately linear relationship.
Jurassic sediments from the Eromanga Basin have low maturity (TNR-I <0.5; DNR-I <5)
to moderate maturity (TNR-I -0.8; DNR-1 -6). Permian samples from the Cooper Basin
have moderate to high maturity values.
The Murta Formation has been shown to be an effective early mature source rock as
indicated by various saturated biomarker parameters (Michaelsen and McKirdy, 1989;
Powell et aI., 1989). Its C32 hopane isomerisation ratio (225/225+22R) rcnges between 52
and 6lVo, its moretane to hopane ratio varies from 0. 10 to 0.13, and its sterane isomerisation
ratio (20S/20S+20R) based on etþlcholestane (CZÐ 5s",14u,174 epimers varies from 30
to 4OVo.
23
Chapter I
The carbon dioxide (COz) content of the CooperlEromanga gases is highly variable. Values
ranging from 10 to lSVo have been recorded. CO2 content is a function of source rock
maturity (Hunt, 1979; McKirdy, 1982). Moreover, COzmay play a key role in the primary
migration of liquid hydrocarbons from source rocks containing hydrogen-poor terrestrial
organic matter (McKirdy and Chivas, 1992). In the Patchawarra Trough, CO2-rich gas
(CrlCrC4 <0.85; 6l3Ccoz = -12 to -Il%o appears to be associated with expulsion of
paraffinic oil and condensate from the Permian coals and/or carbonaceous shales at
maturation levels of 0.9-1.Ivo Rrc.In the Nappameni Trough, where (CrlC1-C4> 0.85 and
ðr3cco2 = -J to0%o,carbon dioxide is accompanied by aromatic condensate generated from
very mature (I.2-l.5Vo Rc) Type trI-IV kerogen. Powell et aI. (I99I) suggested that lower
molecular weight hydrocarbons could have migrated with co-generated gas or carbon dioxide
in a supercritical state during the main or late phase of oil generation to produce paraffinic
condensate and light oils such as those of Cooper Basin type.
24
Gidgealpa-47 Gidgealpa-33
DST 2 DST 4
czz
t( 7
Cg Cn
C3 1
I
?,,
C7
Gidgealpa-46 Sturt-5
DST 3 DST 1
cr¡
czz Pr czz
Tantanna-1
DST 1
Cn
Figure 1.8 Va¡iation of whole-oil alkane profiles of selected oils from the Poolowanna
reservoirs in the Eromanga Basin.
Chapter I
Oils from Permian reservoirs (Tirrawarra, Patchawarra, Toolachee) are usually characterised
by weakly bimodal Ce-C¡o* n-alkane profiles with maxima at Cs-1s and C1e-23. Such a profîle
is typical of fully mature non-marine crude oils derived from Type III kerogen (D. M.
McKirdy, pers. comm., 1993). Exceptions to this rule for the Cooper Basin are the broad
non-uniform profiles of the Gidgealpa-49 (Tinawarra), Strzelecki-16 (Toolachee) and
James- 1 (Triassic) crudes.
Inspection of the oil GC traces assembled by Kagya (1993) showed that the weak n-alkane
bimodality of the Permian oils could be traced upward into younger reservoirs in some fields
(e.g. Gidgealpa). This could be due to vertical migration from Permian source rocks into
younger Jurassic reservoirs. The above observations imply that different source rock facies
in the Eromanga Basin were responsible for the oils found in its reservoirs. However,
vertical migration of oil generated from Cooper Basin source rocks into younger formations
cannot be ruled out. The Patchawarra and Tirrawarra oils can be singled out as having a
specific source rock because of their unique n-alkane bimodality. High wax oils with
n-alkane maxima in the C26.. range may be either relatively immature or water washed, or
both.
Biomarkers
Steranes and triterpanes have been identified in oils from the Eromanga and Cooper Basins.
Using their relative abundance and distribution, Vincent et al. (1985) were unable to reliably
distinguish between oils of Permian, Jurassic and Cretaceous source affinity (see also Fig.
1.9). The saturated biomarkers 25,28,30-trisnorhopane and l9-norisopimarane identified in
some Eromanga-reservoired oils were suggested by Jenkins (1939) to be diagnostic of their
Jurassic source. The l9-norisopimarane marker was found in greatest abundance in crude oil
pools from the northern region of ATP-259P at Cuddapan-l, Marengo-l (Aquitaine Block)
and Toby-l (Wareena Block). A suite of aromatic biomarkers similar to that found in
Jurassic source rock extracts (Alexander et aL.,1988; see also Fig. 1.7) was also identified in
these oils, although a significant number of them did not conform to the DNR-I versus
TNR-1 relationship as shown by the sediment samples.
26
TRITERPANES STERANES
F F
É, É,
E
=
o
z
o
F
F I
f,
T =
!
c
1 t
s 3
X
lr
=
Ê, =
Ê,
UJ IU
o. o.
Figure 1.9 Uniform sterane (m/2217) and triterpane (mlz 191) signatures of Cretaceous,
Jurassic and Permian oils in the Naccowlah Block, CooperÆromanga Basin (Vincent et al.,
1 985).
Chapter I
Oilmaturity
A maturity range of 0.5-1.09Vo F'lc, with a mean value of 0.74 ToRc was reported for a large
suite of Eromanga-resevoired oils (Tupper and Burckhardt, 1990). The Murta-reservoired
oils are of low maturity, analogous to the immature condensates described by Connan and
Cassou (1980). Their Rc values vary from 0.54 to 0.6l%o, with the exception of those of the
Merrimelia (Murta) oils which reflect a higher maturity (0.73-O.84Vo Rc). The corresponding
C2e sterane isomerisation ratios (20S/20R = 0.53-0.82) and C36 moretane to hopane ratios
(0.16-0.23) are likewise low (Michaelsen and McKirdy, 1989). The Jurassic oils are
typicaly more mature than the Cretaceous oils. Michaelsen and McKirdy (1989) reported Rc
values of 0.64-1.I4Vo; C2s sterane 20S/20R ratios of 0.82-1.16; and moretane to hopane
ratios in the range 0.09 to 0.17.
The maturity of oils reservoired in the Cooper Basin varies from 0.65 to 1.09 VoRc, with a
mean value of 0.89Vo Rc, (Tupper and Burckhardt, 1990). Michaelsen and McKirdy (1989)
reported C2e sterane (20S/20R) ratios ranging from 0.98 to 1.14 and moretane to hopane
ratios of -0.02 for these oils.
Based on their gasoline range (Cs-Cz) composition, the carbon isotopic signatures of their
(Cn) saturated and aromatic hydrocarbon fractions, and their pristane to phytane ratios,
Vincent et al. (1985) and McKirdy (1985) categorised three separate oil families of different
source affinity. The first category includes the Cretaceous oils which have higher pristane to
phytane ratios than those of the Jurassic and Permian oils. The gasoline-range hydrocarbon
data indicate a mixed to microbial source for their organic matter; and the isotopic data
confirm that bacterial lipids were the predominant source of their hydrocarbons.
The second family is comprised of Jurassic oils which also had a bacterial contribution to
their source material, except that in this case expulsion from the source rock occurred at
higher maturity levels. As a result, generation of waxy hydrocarbons from leaf cuticles and
spore coatings was achieved. Subsequently, the wax contents of many Jurassic oils were
enhanced by removal of light ends through water washing in the reservoirs.
28
Chapter I
The third family consists of Permian oils and condensates, which may be distinguished from
Jurassic oils on the basis of their gasoline range composition. They are clearly of higher
plant origin, whereas the Jurassic oils are of mixed higher plant and microbial source.
Tupper and Burckhardt (1990) argued that, if the oil's maturity exceeds that of the local
reservoir or source rock, a Cooper Basin source is implied. This argument, however, can
also be true if the oil is derived from highly mature basal Jurassic source rocks in the
Eromanga Basin. Vertical migration from Cooper source rocks into Eromanga reservoirs can
only be confirmed if there exists a unique biomarker for Permian source rocks. Indications
of long distance migration (either lateral or vertical) in the chemistry of an oil could just as
easily be explained by migration within the Eromanga Basin sequence.
The source affinity approach discussed by Vincent et al. (1985) could be more confidentþ
used to determine the origin of these crude oils if all the parameters (including those based on
gasoline-range hydrocarbons) were applied to their respective source rocks. Cretaceous oils
were shown to originatein situblt the low maturity of these source rocks casts doubt on
their effective hydrocarbon generating potential.
As McKirdy (1982) and Kanstler et al. (1933) observed, and as later confirmed by
Alexander et al. (1988), some of the oil in Jurassic and Cretaceous reseryoirs was derived
from within the Eromanga Basin and some may have had a Permian source. The Cooper
Basin oils were of course derived largely from Permian source beds although some might
have been derived from older sedimentary units. Oil expulsion from the Cooper Basin began
in the Triassic with a signif,rcant phase of generation in the Jurassic to Late Cretaceous time
interval (Kanstler et a\.,1983; Tupper and Burckhardt, 1990), whereas oil expulsion from
potential Eromanga source rocks occurred during the early Late Cretaceous to Tertiary
(Tupper and Burckhardt, 1990).
29
Chapter 2
Chapter 2
2.1, Background
Based on the observed associations of petroleum with organic-rich sedimentary rocks, and
the presence of optically active compounds and biological markers in petroleum, it is obvious
of petroleum (oil and gas) is of organic origin. This concept has been
that the vast bulk
proved by various methods such as oil-source rock correlation (Philippi, 1965) and
experimental conversion of buried organic matter into petroleum (Hoering and Abelson,
1963). Marine and terrestrial lipids have been suggested to be the major source of petroleum
(Silverman, 1965a). The great chemical similarity between lipid molecules and many
petroleum hydrocarbons appears to corroborate this suggestion. The non-lipid part of
bacteria and algae may also contribute to the formation of oil (Lijmbach, 1975).
Elaboration of this concept has been achieved by studies which embrace organic petrology
and organic geochemistry. These studies have elucidated the, microscopic, visual and
chemical composition of organic matter in unconsolidated sediments and rocks, and the
changes induced in it by microbial activity and thermal alteration. The organic geochemistry
of hydrocarbons in reservoirs is also investigated and compared to that of organic matter in
source rocks.
30
PALAEOPHYTIC MESOPHYTIC CENO-
ERA
PHYTIC
T
T
TI
TI
a--
¡_
¡TII
ANGIOSPERMAE
IT IIIT
CAYTONIALES
II rTIII
CONIFERALES o
-¡ BENNETITALES 3
IT II! I II--
.
a
!
rII
: -
I
CYCADALES
GINKGOALS
f
o
(n
It
o
=
r-
I -.- I I I -I
I-
-- - CORDAITALES 3
II - ¡E--
I
-- ïÈ I II I PTERIDOSPERMALES
EQUISETALES
Ð
o
-FT
I¡ - t-
L--
-!I
I
¡
II ¡ I
II I
- CALAMARIACEAE
SPHENOPHYLLALES T
I -I I II SIGILARIACEAE
o
I ( - I I tI LEPIDODENDDRACEAE o
=.
o.
IT r SELAGINELLACEAE
LYCOPODIACEAE
E
=
0)
II T> - PTEROPSIDA
<T (D o o m t I,
EI Rd m PERIOD
TT Ð o
Þ
z
F cn=
fl CN
Ø
(t,
Ø
t! ft
z
z o ft
I
z z z õ
PALAEOZOIC MESOZOTC
CENO
ERA
zotc
Figure Z.la T}re stratigraphic ranges of major members of the plant kingdom showing the
difference between plant- and animal- based eras (Diessel, 1992).
Chapter 2
The Precambrian thucholite deposits of the Witwatersrand (Snyman, 1965) and the
Ordovician kukersites (formed from Gloeocapsomorpha prisca) in Estonia (Foster et al.,
1989, 1990) are examples of early boghead coals or oil shales which formed from the
remains of algae or cyanobacteria.
Oil shales continued to form after the advent of higher plants. Examples include the Eocene
deposits of the Green River Formation in the western USA and the Tertiary deposits at
Rundle and Julia Creek in Queensland, Australia (Diessel,1992). Torbanites contain mainly
the fossil green algae PiIa and Reinschia which occur in many Permian boghead coals of
Australia, South Africa and South America. Both PiIa and Reinschia appear to be closely
related to the extant genus Botryococcøs (Stach,1982) and are found in both fresh and
brackish water. The species Botryococcus braunil is regarded as the source of coorongite, a
Recent bituminous substance found near Salt Creek in South Australia. Tasmanites sp. is
another fossil chlorophyte found in some marine oil shales (tasmanite). It is related to the
extant genus Pachysphaera, and is known from Permian deposits in Tasmania and
Cretaceous deposits in Alaska (Tissot and'Welte, l9l8).
Algal bodies show little differentiation, do not possess a vascular circulatory system and are
therefore dependent on osmosis and the presence of water to sustain their life cycle. Apart
from being diverse throughout the geological record, their presence indicates that the host
rocks were deposited in sub-aquatic environments.
The first land plants, the Psilophyta (also called Psilopsida), appeared in the Early Devonian.
They were descendants of the algae but are regarded as a separate group and assigned to a
class of the pteridophytes (spore plants). They were the first plants to achieve the transition
from water to land, having a simple vascular system and displaying cell differentiation. The
early transitional forms lived half submerged in water and bore sporangia (spore capsules) at
the tþs of their leafless stems. Later fossil forms bear evidence of a more terrestrial habit and
were well equipped for life in swampy environments. During the Carboniferous Period they
gave rise to banded humic coals in the Northern Hemisphere. Ground water fluctuation
seems to have affected their growth and diversity and consequently fewer economic coal
seams were produced.
During the Permian Period, the effects of ground water fluctuation on peat formation a¡e less
evident. This time marked the appearance of arborescent vegetation which was dominated by
seed plants. The arborescent gymnospenns were among the first plants to grow on relatively
dry ground and therefore they are characterised by longer roots. The Permian flora had
greater tolerance towards environinental change enabling them and their vegetational
successors to continue accumulating peat where the Carboniferous flora would have ceased
to do so. In addition to vitrinite-rich bright coals, Permian coal measures contain many dull
32
Chapter 2
coals with high proportions of inertinite, which is commonly formed under relatively dry
conditions.
The spore content of Carboniferous humic coals is higher than that of equivalent younger
deposits which, from the Permian onward, were increasingly based on seed plants.
However, it has been observed that reproduction of pteridophytes requires a moist
environment and fresh water to fertilise the spores (Collinson and Scott, 1987; Phillips,
1979). Because of adverse conditions during the Carboniferous, fertilisation had a low
success rate. The pteriodophyte flora overcame this problem by producing spores in large
quantities in order to assure that reproduction was kept at high level. In some lacustrine
environments, spores were so concentrated that they formed a special type of coal known as
'cannel' coal.
The Permian Period marks the end of the Palaeophytic and the beginning of Mesophytic Era
(Fig.2.1a), in which the gymnospenns (plants with naked seed) constituted the leading plant
group. Like the psilophytes, which were the forerunners of most of the pteridophytes, the
pteridosperm (seed ferns) constitute the link between spore-bearing ferns and seed plants.
They reached their maximum development on the southern continents (Gondwana) during
the late Carboniferous and Permian. They are responsible for the formation of rich coal
deposits in Australia, South Africa, South America, India and Antarctica. One of their
leading genera gives its name to the whole plant association, which is often referred as the
Glossopteris flora (Gould, 1975; Gould and Shibaoka, 1980; Retallack, 1980). Other
gymnosperms which were common in the Permian are the Cordaitales, the Coniferales and,
toward the end of the period, the Cycadales. Post-Carboniferous coals therefore frequently
have low spore contents but are rich in derivatives of plant resins and waxes (Diessel, 1992).
Figure 2.lb shows the stratigraphic distribution of conifer fossils in the Cooper and
Eromanga Basins of Australia. Near the close of the Permian, the pteridosperms underwent
an abrupt change with the introduction of the Dicroidium flora and an associated sporeþollen
assemblage known as the Falcisporites microflora. The Dicroidium flora as a whole included
only a low proportion of conifers, notably Voltziopsis and Rissikiø (Gould, 1975; Gould
and Shibaoka, 1980; Retallack, 1980; Townrow, 1969).
33
TOTAL BIOMARKER
LITHO. SELECTED DIST. IN
AGE STRATIGRAPHY CONIFER
CONIFERS SOURCE
DIST.
ROCK
LrJ
L
- 10C
U'
f Idt¡nkEtiEÉ+tD
o
IJJ
Bulldog Shalê Wellumbllla Fm
o ìcÍ.
I,JJ trJ
s@E
Cadna-Owlê Fm
É
O
Murta Fm
!a
-éu)
IJJ
- 15C L Namur Ssl
3
Blrkhsad Fm
I
U)
trJ
J
U) ô
ô Hunon Sêt
É.
lI =
Poolowenne Fm
E
ì(E tl)
-200 B€an bush Beds
t!
PE
no
:r¿
Morney Beds
o(Ú
o IJJ
L
Cuddapan Fm
ÈE
(úo
3
CD
Ø 5F
iñe
E ô
=
IIJ
Nappamenl Grcup
aØl Ëg
-250
z
IJJ
L frotachteFm
æ Él SF
3
ffi ,El o
ú ì fLl =
L
lrJ
fT
Petchawarrå Fm
ol Oo
fL
t! Tlrrawafia Sst ol (ú
-280
9 Ê
Figure 2.lb The stratigraphic distribution of selected plant genera in sediments of the
Cooper and Eromanga Basin (modified afterAlexander et a1.,1992).
Chapter 2
The Triassic flora were dominated by the pteridosperm pollen Falcisporites. Microfossil
observations have recorded the occurrence of a very low proportion of pinacean-like,
podocarpaceanlike and inaperturate pollen, which are thought to represent plants with
araucariacean affinities.No megafloral remains of the Araucariaceae have been recorded in
Australian Triassic sediments (Townrow, 1969). However, Alexander et aI. (1988) noted
some tentative evidence of sparse araucariacean forms in the Triassic of the Cooper Basin.
Of the Triassic conifers, only the podocarp Russikia is related to a modern family
(Townrow, 1969).
The beginning of Jurassic is marked by another floral change whereby, for the first time,
conifers attained dominance over the pteridosperms and a low but significant proportion of
araucariacean forms. The earliest Jurassic microfloral assemblages, such as those from the
Poolowanna Formation, are strongly dominated by Classopolis (Carollfutn) pollen, believed
to be derived from the Cheirolepidiaceae, an extinct conifer group. Cretaceous members of
this family are allied to extant Cupressaceae (Miller, l9l7), which are widely represented in
Australian flora.
The Cretaceous plant record is marked by the dominance of angiosperms which first
appeared in Triassic Period. In spite of the preponderance of angiosperms, the gymnosperms
remained important contributors to peat deposits. Their abundance in the resulting coals is
related to their resinous wood which resists rapid aerobic decay (Diessel, 1992). The modern
podocarpacean genera, Dacrydium and Phyllocladus, first appeared in the Cretaceous and
together withPodocarpus gave rise to the distinctive the diterpane biomarkers, identified by
Alexander et al. (1987) in oils and source rocks of the Gippsland Basin.
35
Chapter 2
liptinite group comprises macerals which are derived from cuticular or other resistant vegetal
matter rich in resinous and waxy substances.
The International Committee for Coal Petrology (1975) distinguishes between macerals and
submacerals by reference to either their particular vegetal origin or their respective state of
preservation. In addition, a unified brown/black coal maceral classification was adopted by
Standards Association of Australia (1986), wherein three subgroups of tissue-derived
macerals are recognised according to their state of humification prior to physico-chemical
gelification (Table 2.1). The following discussion of macerals which make up the vitrinite ,
inertinite and liptinite groups draws heavily on the summary accounts of Stach et al. (1982)
and Diessel (1992).
36
Chapter 2
V/oody and cortical cell tissues are thought to be the main progenitors of this subgroup.
Under severe and prolonged conditions of degradation, the cell texture of lignin-rich wood
will tend towards total breakdown. The rate of degradation of these tissues preservation will
depend also on their botanical origin. For instance, wood tissues produced by gymnosperms
are more resistant to mechanical, microbial and chemical degradation because of their high
content of resins and tannins; whilst angiosperm wood is frequently badly degraded.
Macerals identifiable in brown coals are as shown in Figure 2.2a. Textinil¿ is distinguished
by its cell structural features which differ little from that of original wood. In some cases,
the unaltered anisotropic cellulose and lignin with primary fluorescence are retained (Russell,
1984). It is equivalent to telinite of bituminous coal rank. Plant genera and species are
readily identified from textinite tissue, whereas in telinite identification is possible only with
difficulty.
37
INTENSITY OF UMIFICATION ERALISATION
¡o
.9
Soft Tissues o
(tt
(mainly cellulose) à
E
Ë
Woody Tissues d
CL
(mainly cellulose
.9.
and lignin) o
E
/,/,/,/,/,/, /////,/,
Preservation of
ATTRINITE DENSINITE
HUMODETRINITE
Macerals
in higher rank TELINITE TELOCOLLINITE DESMOCOLLINITE
after gelification
and polymeri- DETROVITRINITE "$er!o6_corrrr'r
TELOVITRINITE GELOVITRINITE
sation of hunic
colloids
- GELINITE -*CORPOHUMINITE
./\
Telogelinite Phlobaphinits Pseudo-Phlobaphinite
Porigelinite Levigelinite
./ Detrogelinite
Eugelinite
Figure 2.2a Schematic outline of the formation of the vitrinite group fiom tissued
vegetable matter subjected to varying intensities of humification (modified after Diessel,
1992)
Chapter 2
Post-depositional epigenetic gelification of cell tissue begins in the catotelm during the peat
stage and, after polymerisation of the humus colloid, culminate in formation of telovitrinite,
mainly in the form of telinite (Fig. 2.2a). The diagnostic feature of telovitrinite is that it is
free from contamination by other macerals. All telovitrinite displays a reflectance 5Vo to l}Vo
higher than that of its associated detrovitrinite which is subjected to a higher degree of cell
destruction before final burial (Robert 1979).
Attrinite consists of loosely packed cell fragments and other plant debris, including liptinite
and inertinite. The latter are discontinuous phases in humodetrinite. Epigenetic gelif,rcation of
attrinite is common.
H u m o c o llínit e- g eI ov itrinít e
Humocollinite is a coalified humic colloid without any inclusions of remnant cell tissue.
Although rare, it occurs sporadically in the form of gelinite and corpohuminite in brown
coals. They occur as precipitates in cavities, cleats, and fissures of peat and are generated
from a humic hydrosol (humic acid) during humification of decomposing vegetal matter.
39
Chapter 2
Gelocollinil¿ is the black coal equivalent of gelinite. It occurs as infillings of cell lumens and
gaps between cell walls in semifusinite. This colloidal humic matter commonly contains
inclusions of cell fragments and other organic debris. The spheroidal and elliptical bodies of
flocculated humic colloids contained in cell lumens of brown coal are known as pseudo-
phlobaphinite.These are secondary infîllings which resemble the phlobaphinite, a substance
consisting of tannin and non-humic excretions like that in the cork cells of bark tissue. On
disintegration of the cell tissue these corpohuminite bodies become isolated, forming the
corpocollinite of black coal.
Inertinites are characterised by high reflectance in incident light microscopy because of being
rich in aromatic carbon. They are relatively brittle and hard and thus tend to develop a high
polishing relief. The proportion of inertinite in humic coals varies over a wide range, but is
frequently between 20 and3ÙVo.Figures which deviate substantially either way from this
average, occur in coals which have been formed under particular sets of environmental
conditions. The inertinite group is divided into three subgroups according to the degree of
cell tissue preservation.
Telo-inertinite
This subgroup comprises structured macerals in which gelification is either absent or minor.
The most coÍìmon members are pyrofusinite (or simply fusinite), semifusinite and
sclerotinite. Pyrofusinir¿ is fossil charcoal resulting from incomplete combustion of wood
tissue. Of all the inertinites, it shows the highest degree of preservation of cell tissue, as well
as the highest reflectance and polishing relief. It has a distinct yellow tinge in ordinary
incident light, whereas it is opaque in transmitted light and does not fluoresce under [fV
illumination.
40
INTENSITY OF HUMIFICATION MINERALISATION
.9
Soft tissues -o
o
(mainly o
(ú
Woody Tissues
à
.(ú
( cellulose Ë
(ú
o-
and lignin)
Figure 2.2b Schematic outline of the formation of macerals from tissued vegetable
matter subjected to varying intensities of humification and drying before burial below the
ground water table (Diessel, 1992)
Chapter 2
The cell walls of pyrofusinite are rather thin, as suggested by Barghoorn (1949), only their
resistant lignified portions survive combustion. In most cases pyrofusinite occurs as broken
curved fragments, known as bogen (bow) structure, which are caused by either overburden
or tectonic pressure. Its higtrly aromatic structure does not undergo further change during
physical-chemical coalifi cation.
Detro-inertinite
This subgroup consist of two macerals, inertodetrinite and micrinite. They both originate
from fusinitised detrital plant fragments and are distinguished on the basis of their size.
Inertodetrinite is a fragmented inertinite with the longest diameter ranging between 2 and 30
pm. Micrinite consists of the smallest inertinite grains and is commonly referred to as
granular opaque maffer by(Teichmüller, 1974a). Much micrinite forms at the sub-
bituminous rank level from lipid rich material by a disproportional process which results in
the formation of petroleum-like materials such as exsudatinite and bitumen, leaving behind
micrinite as a residue (Teichmüller,l974b).
42
Chapter 2
The Australian Standard 2856 (Standard Association of Australia, 1986) regards all
inertinites smaller than 2pm as micrinite. The proportion of inertodetrinite in coal varies quite
considerably, but it is generally small. Carboniferous coals contain more micrinite than
younger coals (average content is 3-6Vo but may be as high as l9%o). The micrinite contents
of Permian and the post-Carboniferous coals rarely exceed3Vo.
Gelo-inertinite
Macerals of this subgroup are derived from plant material which was first biodegraded into
humus colloids and subsequently dehydrated and oxidised (Stach and Alpern,1966). The
only defined representative maceralis macrinite wl'ttch is subdivided into corpomacrinite, and
l.ammncrinire (Diessel, 1992). In both transmitted and fluorescent light modes, macrinite
covers not only a wide range of reflectance in relation to the associated vitrinite, but like
semifusinite, there is an inverse relationship between reflectance and the intensities of
translucency and fluorescence.
Liptinites are coÍrmonly subdivided into primary and secondary liptinite (Table 2.1).
Primary liptinites are part of coalified plants, whereas secondary liptinites are a group of
products derived from thermal condensation and dissociation reactions. Both categories
occur in only relatively small proportions and rarely exceed ZOVI inmost of humic coals, but
they can be very concentrated in sapropelic coals (i.e. coals formed by subaquatic
sedimentation of floating vegetation (algae) and allochthonous organic matter).
43
Chapter 2
Primary liptínítes
Macerals of this subgroup are as listed in Table 2.1 and their description is as follows:
Sporinite is the protective skin (exine) of spores and pollen grains. It is made up of
sporopollenin which according to Shaw (1970) consists of oxidative polymers of carotenoid
esters. Taylor and Liu (1987) suggested that the resistance to decay of sporinite is due to the
high degree of molecular cross-linking in sporopollenin. Sporinite occurs in high
concentrations in cannel coals and it is the most common primary liptinite maceral in humic
coals and other carbonaceous sedimentary rocks. Because of compaction caused by
overburden pressure, sporinite appears as small flattened lenses in sections normal to
bedding.
Size and shape are the main features by which sporinite is normally distinguished. Large
spores ranging in (flattened) diameter between several hundred micrometers and several
millimetres are referred to as macro- or megaspores (hence macro- or megasporinite).
Macrospores are the mostly female spores of heterosporous plants such as lycopods
(Kosanke, 1969). Their exines are omamented by various protrusions and small semi-
detached spheroidal appendages which represent abortive spores. In polished sections of low
rank coals their dark surface has reddish internal reflections and commonly also granular
inclusions. They are relatively rare in Carboniferous coals and with the reduced contribution
of pteridophytes to younger sediments it is most likely that macrosporinite will be even less
conìmon in post-Carboniferous coals. Smaller spores with diameters measuring less than a
few tens of micrometers are referred to as microsporinite. These spores are derived from
homosporous plants. Male microspores of heterosporous plant constitute the main liptinite
maceral in some Carboniferous humic coals (up to l5Vo) and in some cannel coals (up to
807o: Diessel,1992). The sporinite content of Permian coals in Australia averages only 3Vo.
As a consequence of the advent of seed plants (gymnosperms and angiosperms), younger
coals and carbonaceous sedimentary rocks contain an increasing proportions of pollen
grains. The term for both microspores and pollenis miospores (Guennel,1952).
Studies of its chemistry have shown that sporinite is not a major source for normal
hydrocarbons. Liptinite-rich samples with abundant sporinite and liptodetrinite gave low
yields of normal hydrocarbons, and those obtained were of lower molecular weight i.e. non-
44
Chapter 2
waxy (Powell et al., l99l). However, Kruge et al. (1991) noted that the pyrolysate of
sporinite had a greater predominance of z-alkanes over branched and cyclic alkanes as
compared to that of the parent coal; and it also showed a lesser predominance of
polyaromatics over n-alkyl benzenes. Mild oxidation products of sporinite were also found
to be significantly different, in that sporinite was less polycondensed than the parent coal.
Allan and Larter (1983) reported the dominance of aliphatic, alkylated phenolic and alþlated
aromatic components in the pyrolysates of both vitrinite and sporinite macerals, although
sporinite of any given rank was found to have greater amounts of aliphatic material.
Cutinite under the microscope appears as straight or wavy lines either translucent in
transmitted or dark in incident light, usually with palisade ridges on one side. It is formed
from cuticles, the waxy cover on the epidermis of leaves and young shoots. Goodarzi (1984)
and Bartram et al. (1987) suggested that some chitinous cuticles may have been derived from
the epidermis of arthropods. Cuticles consist of cutin, a glycerine ester of fatty acids (Stach
et a1.,1982) from which hydroxy and epoxy fatty acids can be derived by depolymerisation
(Kolattukudy, 1976).
Cutinite resists biodegradation better than its associated mesophyl tissue which decompose
rapidly, and as a result it sometimes appears to be densely packed. However, cutinite is not
as resistant to biological and chemical attack as sporinite (Taylor and Liu, 1987). Cuticles
concentrate mainly in shallow ponds, soon after detachment from the parent plant, without
much transportation (DiesseI, 1992). Cutinite is of relatively low abundance in most coals
and rarely exceeds 2 or 37o.
The thickness of cutinite seems to be influenced by the growth and depositional environment
of the parent plant. This is because the thickness of the cuticle which protects the inner tissue
of plant from drying, is related to the availability of water. It has been observed that plants of
wet environments have thin cuticles, whereas those growing in comparatively dry areas are
characterised by thick cuticles (Strasburger, 1983: as cited by Diessel, r99z).
In Australian coals and carbonaceous shale of Carboniferous to Tertiary age, the yield of
high molecular weight n-alkanes upon pyrolysis correlates strongly with the content of
cutinite plus liptodetrinite (Powell et al.,l99l).
45
Chapter 2
Resinite is derived from the resins of vascular plants. It displays a wider range of optical
properties than other liptinite macerals because of its varied origin and post-depositional
history. It consists of high molecular weight (mainly) aliphatic compounds including resin
acid, resin esters and terpenes (Selvig, 1945). Cunningham et aI. (1983) noted that fossil
resinite (amber) had been formed by photolytic polymerisation, upon exudation from the
host tree.
In coal, resinite occurs either in situ in resin ducts and the cells of xylem, cortex, mesophyl
and seeds (White, 1914; Selvig, 1945) or in dispersed form as lumps, nodules and rodlets
(Kosanke and Harrison, 1957; Lyons et aL, 1984). Some concentrate as residuum after its
host tissue has decayed. Substantial portion of resinous maffer contained in vascular plants
become absorbed by humus colloids during advanced humification, thus causing low
reflectance readings in detrovitrinite.
Based on the pyrolysate results, Mukhopadhyay and Gormly (1984) reported that resinite
begins to generate hydrocarbons at a very early stage of maturation and may not have much
hydrocarbon potential left beyond a vitrinite reflectance of O.\Vo. They also reported that
different types of resin produce different classes of compound. Some resins contain both
terpenoid compounds and n-alkanes whilst others (brown-fluorescent resinites) contain
mainly n-alkanes. Most resins generate sesquiterpenoid compounds which on thermal
maturation give rise to aromatic hydrocarbons.
Other types of resinite are derived from resin based primarily on polymers of labdatriene
diterpenoid carboxylic acids, especially various isomers of communic acid, ozic acid and
zanzlbaic acid. Modern resins based on these compounds are produced, often in significant
amounts by various plant taxa, including the Araucariaceae, Leguminosae and Cupressaceae
(Anderson et al., 1992). Others are derived from resins based on polycadinene structures
(van Aarsen et al., I99l). Analogous modern resins, sold commercially as 'dammaf , aÍe
produced in large quantities by trees belonging to the taxon Dipterocarpaceae.
Suberinite is the preserved suberin impregnated cell walls of the cork tissue. It is only
known from Tertiary and few Mesozoic coals. Suberinite commonly occur concentrated into
layers from 0.02 to 0.5mm thick. Under the microscope it displays an outline of its bearing
cell tissues (i.e. rectangular, brick-like, or irregularly polygonal four- to six-sided shapes).
46
Chapter 2
In cases such as biodegradation where the corpocollinite cell f,rllings disappear, it appears in
schlierren form like resinite. In reflected light it is dark to medium grey in colour; green to
red internal reflections, particularly, in the low rank (i.e. brown coal stage) and is
unrecognizable beyond a vitrinite reflectance of about 0.8Vo Rm. It shows variable
fluorescence intensties, ranging from greenish yellow to brown, with increasing rank. At
higher rank (from 0.67o Rm) the fluorescence intensity becomes very weak. Suberinite is
similar to cutinite in composition but the suberin is less strongly polymerized than the cutin
and thus is more easily attacked. Suberin is considered to be a polymer containing aromatics
and polyesters (Kolattukudy, 1980).
Alginite is attributable to small unicellular algae. Fossil alginites have been encountered in
various geological settings, from the Ordovician kukersite of Estonia which is formed from
the remains of a marine planktonic algal or cyanobacterial species Gloeocapsomorpha prisca
(Downie, 1967;Foster et al., 1989, 1990), to lacustrine deposits of the colonial green alga
Botryococcus braunü (Robert, 1988). It is classified into two major types, telalginite and
Iamalginite, according to their microscopic morphology. Telalginite comprises intact
structured bodies having spheroidal to flat lenticular shapes. In contrast, lamalginite occurs
as thin anastomosing lamellae formed by algal mats inter-layered with mineral grains (Hutton
Botryococcus braunü has in its outer cell walls a resistant aliphatic biopolymer that is a
possible source of n-alkanes (Berkaloff et al., 1983). On pyrolysis this biopolymer yields
n-alkanes up to C31 and therefore appears to be a major building block for Botryococcus-
derived Type I kerogen. Furthermore, McKirdy et al. (1986) noted that different clonal
races of Botryococcus algae synthesize unsaturated hydrocarbons (C23-C3 3, n-alkenes and
C3a botryococcenes) which are potential sources of long-chain n-alkanes and botryococcane
4l
Chapter 2
Secondary liptinites
These are liptinites formed during coalification as by-products of chemical dissociation of
primary liptinites. They lack any distinctive morphology and some of them are characterised
by intense fluorescence under blue/UV light. At the time of their formation they occupy
mostly cell lumens and small fractures. This maceral subgroup includes fluorinite, bituminite
and exsudatinite.
48
Chapter 2
proceeds, fluids and gases are released by diffusion from these non-aromatic moeities. As
the proportion of non-aromatic compounds decreases, there is an increase of aromaticity (fu=
aromatic C/total C) and consequently also an increase in reflectance. Aromaticity increases
from approximately 0.56 in brown coal to 0.78 in medium-volatile bituminous coal and 0.95
in anthracite (Iyengar and Lahir, 1959).
Conventionally, vitrinite (particularly telocollinite) is used for the determination of coal rank
by micro-reflectance; hence the term 'vitrinite reflectance'. Factors favouring the use of
vitrinite for reflectance measurements include its ubiquity, homogeneity and ease with which
a well-polished surface can be obtained. As condensation of the aromatic complexes is
associated with an increasingly parallel orientation of their planar layers, the rise in
reflectance is accompanied by the development of optical anisotropy. Therefore, reflectance
'measurements can be carried out in either polarised light or non-polarised light. In non-
polarised light, R¡¡ is the mean maximum reflectance (measured in oil immersion) and Rs is
the mean random reflectance. The relationship between maximum and random reflectance
can be expressed as % R¡¡ = 1.106 Ro - 0.024 (Neavel et aL, 1981). Diessel and McHugh
(1986) suggested an alternative relationship for Australian coals: Vo F.rr¡= l.O7 Ro - 0.01.
In stagnant reducing environments (e.g. Black Sea, Lake Tanganyika) the boundary between
the oxic and anoxic zones is higher up in the water column (Demaison and Moore, 1980). It
has been observed that usually only small amounts of organic matter are finally deposited.
49
Chapter 2
However, in
conditions where photosynthetic activþ is abnormally high because of
enhanced nutrient supply considerable amounts of organic matter can accumulate. This
occurs in areas of coastal up-welling. In such cases the oxic zone of the water-column is
drastically attenuated as a result of heterotrophic bacteria consuming much of the dissolved
oxygen. Consequently, anoxic conditions are established in the bottom waters where
degradation of the sinking organic matter by anaerobic bacteria is slow and incomplete
(Foree and McCarty, 1970; Otsuki and Hanya,1972).In this way the original organic matter
is reworked and converted into a biomass consisting of bacterial bodies mixed with the
microbially resistant parts of the planktonic bodies and terrestrial organic matter supplied
from elsewhere. Organic-rich sediments (3-2OVo TOC) can be deposited under these
conditions.
The first step in identifying a hydrocarbon source rock is to determine its content of total
organic carbon (TOC) and extractable organic matter (EOM) or bitumen. The second step is
to determine the type of kerogen and the composition of the solvent-extractable
hydrocarbons. Finally, the evolutionary stage or maturation level of the kerogen is
determined.
50
Chapter 2
polycondensed nuclei with alkyl side chains linked by heteroatoms such as S and O (Tissot
et aI., I974a); or of
sedimentary rocks that is soluble in neither
as an organic constituent
aqueous alkaline solvents nor the common organic solvents (Tissot and 'Welte, 1934). It is
commonly classified visually on the basis of its maceral group composition or chemically on
the basis of its carbon (C), hydrogen (H) and oxygen (O) elemental composition. It is
isolated from sediments by hydrofluoric and hydrochloric acid attack to remove the mineral
matrix, or by density separation using heavy liquids. The amount and type of hydrocarbon
products are determined by its biological precursors and the degree of preservation (Durand,
1980).
'When
the elemental ratios (FVC and OiC) are cross plotted on a van Krevelen diagram (Fig.
2.3a) they provide an immediate reference to the three main types of kerogen (viz. kerogen
Type I, II and III: Fig. 2.3b). An approximate equivalence of various terms used in kerogen
description are as summarise d in T able 2.2.
Provenance
Aquatic
Liptinite Amorphous Type I
Amorphous
Herbaceous
(fibrous) Type II
ru
Woody Vitrinite
(plant structure) Humic
Sub-aerial
(tenestrial) (angular
to sub-angular
fragments) Inertinite Residual
Due to the mixed nature of sedimentary organic matter, the Type I, II, and III pathways on
the HI vs. OI plot are only generalised. Thermal maturation and early diagenetic oxidation of
the organic matter also influence the position of sample on the HI vs. OI plot (Durand and
Monin, 1980; Demaison and Bourgeois, 1984). For instance, coals of low maturity (Ro
<0.67o) generate large amounts of carbon dioxide during pyrolysis; whereas at higher levels
of maturity, proportionally more pyrolytic oxygen is released as carbon monoxide, which is
not analysed by the Rock Eval pyrolyser and thus may give erroneously low OI values.
51
Chapter 2
Parameter Measurement
S1 Quantrty of hydrocarbons and related materials desorbed from rock
sample during heating at 250'C for 5 min in helium - approximately
equivalent to material extracted in solvent procedures
S2 Quantlty of hydrocarbons and related material produced during
pyrolysis of kerogen and in volatile bitumens over the temperature
range 250-550"C at250'C min-l in helium - approximately equivalent
to thermal degradation of kerogen
s3 Quantity of carbon dioxide produced during pyrolysis of sediment over
the range 250-390"C
T** Temperature of maximum kerogen decomposition rate (maximum peak
52 evolution) - relatable to the maturity of the sample
51+52 Genetic potential
S1/( 51 + 52) Production index (PI)
s2Æoc Hydrogen index (HI)
s3Æoc Oxy gen index (OI)
52
2.0
c
1.5 W
H/C 1.0
0.5
0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
o/c
Figure 2.3a T}lie atomic IVC versus O/C diagram of van Krevelen (1961) illustrating the
convergence of maceral composition during the physicochemical stage of coalification.
(A= ¿Ílgal matter; C= cellulose; E= spore eúnes and other liptinite; F= inertinite; L=
lignin; V= vitrinite;V/= wood)
2.
I
.50
U II
o
o
.00
0.50
The reflectance intervals shown above for the oil and gas zones are only approximate. There
has been much discussion of this topic (see e.g. Dow, 1977; Hercux et al., 1979; Alpern,
1980; Robert, 1980; Teichmüller, 1982).
T,o* is the temperature in "C at which the maximum kerogen pyrolysate (S2) containing
hydrocarbons and other compounds is produced. It also increases progressively with
increasing depth thus qualifying as a maturation parameter. Unlike the production index, it is
not affected by migration factors and a good correlation between these two parameters is
quite common. It also correlate well with other maturation parameters such as vitrinite
reflectance for Type III kerogen and humic coals. It is particularly valuable in the case of
marine or lacustrine kerogen (Type I and II) where vitrinite is often scarce or absent.
Approximate T,o* boundaries for oil generation are < 435"C (immature), 435 to 460'C
(mature) and >460'C (postmature). However, the type of organic matter seems to influence
T,n* values during diagenesis and early catagenesis when values are generally higher in
kerogen Types I and II (Tissot et al., 1978). However some exceptions have been noted,
where Type II kerogen do not always show higher T,n* values than Type III kerogen of the
same rank. Tn,o* values are almost equivalent for the different kerogen types in the peak zone
ofoil generation and later in the gas zone (Peters, 1986).
54
Hydrocarbonsgenerated
-+
0
Biogenic CH4
.o o
U'
o J
c al
o
o) E
.g 1
o -E
3
o
E
C
th
'6 '=
o
c o
o
o) E
C' .Y
G' 3
o aú
o)
E
o-
o o
ô
=
4
.att
at,
o tt,
co (¡l
o) 5
ôà
trt)
ct
o
Biomarkers oit
N Gas
Recycled organic matter (OM) and altered samples appear to have elevated T-* values. If the
total 52 from the recycled OM greatly exceeds that from any accompanying primary OM, the
T.* value can be much higher (40'C or more) than expected based on the actual maturity of
the sediments. Normally, the T,o* difference between recycled OM and mature or postmature
OM is generally less than 5"C. Oxidised, highly mature, or Type fV kerogens usually show
high T** values or lack an Sz peak (e.g. in siltstones and sandstones accessible to
oxygenated ground water). This is because oxidation tends to remove hydrogen and add
oxygen to the kerogen, thus resulting in depletion of 51 and 52 whilst enriching 53 values.
This 53 enrichment results in erroneous OI signatures which give rise to ambiguous kerogen
maturation pathways on the HI vs. OI cross plot. Therefore, in order to avoid such
problems, a HI vs. T*u* cross plot is used to differentiate maturation pathways rather than
the conventional HI vs. OI plot (Espitalié et al., 1984).
Changes in kerogen type or geothermal gradient within the sequence may have an effect on
the distribution and amount of bitumen, thus making it difficult to delineate regular trends of
maturation and hydrocarbon generation. Short-range migration and minute accumulations
within the source rock may also alter the amount of bitumen.
Further assessment of the maturity of the organic matter in source rocks may be carried out
by using the detailed composition, rather than the abundance, of their hydrocarbons. These
hydrocarbon compositional parameters (e.g biomarker isomer ratios) and their relation to
maturation effects are discussed in later sections of this chapter.
56
Marine Standing
bicarbonate Kinetic biomass
1V effect
-lóô 21To
Kinetic
Equilibri isotope
Equilibri
ne Atmospheric
carbonate carbon d¡oxid
ca. 4x 0 I
Sedimentary
carbonate organic
75To -26
(ca. g
carbon
1001o
Figure 2,5 Approximate relative sizes of carbon reseryoirs together with average ô l3C
values within the earth's surface, crust and mantle (Killops and Killops,1993)
Chapter 2
As the depth of burial of the source rock increases further, the 'metagenesis' stage is
reached. Here cracking becomes so severe that the kerogen and any retained oils is broken
down into light hydrocarbons and eventually into methane. At this stage more gas is
generated and any unsaturated hydrocarbons formed become saturated by hydrogen derived
from disproportionation reactions which lead to polyaromatic compounds and ultimately
graphite.
2.5 Stable carbon isotope signature of rock extracts and crude oils
The ability of carbon to form chains and rings enables it to produce an immense variety of
compounds, including hydrocarbons and their derivatives. It is therefore necessary to
understand its mode of occurrence, and how it is first incorporated in living organisms, then
into sedimentary organic matter and eventually into crude oils. One way of doing this is to
examine its stable isotopic composition.
sunlisht/chlorophvll
6Co2 + 6Hro , C6Hl2c)6 + 6C.2
58
Light
P
Chlo NADP a
ê-
Ribulose
Triose ate
p P (cs)
ATP
ADP
Light stage Dark stage
AEROBIC by PHOTOSYNTHESIS
ENVIRONS mêtazoa, by
cyanobacteria,
algae and
higher plants
INORGANIC CARBON
FROM
ATMOSPH ERE,
OCEAN & ROCK
ANOXYGENIC
ANAEROBIC PHOTOSYNTHESIS
and by
ANAEROBIC
FERMENTATION sulphur bacteria,
ENVIRONS
by microbes cyanobacteria
MATTE
Sediments
E.9
õg)
co
oo_
UMIN
'Eo
(EL
lo
lc
É-c l=.
o)=
=å t-
10,
KEROGEN
@
c
OILAND =.
0)
GAS
GRAPHITE
Figure 2.6b The carbon cycle and the interactions between its major participants
(modified after Summons, 1993)
Chapter 2
The photosynthetic process has two stages, one of which operates in the light and the other
in the dark (Fig. 2.6a). In the first stage, light energy emitted by the sun is adsorbed by a
green pigment, chlorophyll. This results in the formation of carbohydrate units by transfer of
hydrogen atoms from water to carbon dioxide molecules and the liberation of oxygen from
water molecules. Carbohydrates are used in the biosynthesis of other organic compounds
and to provide an energy store for the performance of normal cell functions.
During the dark stage, various carbohydrates ale synthesised through the assimilatory
pathway known as the'Calvin cycle'. Under this pathway carbon dioxide is first combined
with a C5 compound ribulose diphosphate (RDP), which then splits into two molecules of
the C3 compound phosphoglyceric acid (PGA). The PGA is used to synthesise further RDp
while some of it is reduced by NADPH (nicotinamide adenine dinucleotide phosphate-H),
utilising the energy supplied by the ATP/ADP system, to generate triose phosphate which
consequently is converted to the glucose phosphate which is the source of various
carbohydrates. All autotrophic carbon fixation, photosynthetic and chemosynthetic involve
this cycle.
Most plants and cyanobacteria (blue-green algae) fix carbon using the Calvin or C3 pathway.
Those which use the Calvin cycle alone are termed C3- plants to designate the involvement of
PGA which contains three carbon atoms. Two additional biochemical pathways -the Ca or
Hatch-Slack pathway and the CAM (crassulacean acid metabolism) which fix CO2 at night
are used by smaller groups of plants. Under these processes the carbon dioxide fixed during
the night is released again within the plant tissue during the day for incorporation via the
Calvin cycle. This mechanism reduces photorespiration by closing stomatal pores during the
day while maintaining essential supplies of CO2. Two plant groups are named after their
additional mechanisms, Ca-plants (tropical grasses, desert plants and salt marsh plants) and
CAM-plants (succulents). Some photosynthetic anaerobic bacteria use H2S as a source of
hydrogen instead of water, and so liberate sulphur instead of oxygen. Other species, such as
the pulple non-sulphur bacteria (Rhodospirillum), use simple organic compounds as a
source of hydrogen.
60
Chøpter 2
To summarise, carbon in the form of carbon dioxide, is continuosly cycled through plants,
animals and sediments by the processes of respiration, photosynthesis, growth, decay and
preservation in sediments. The organic carbon trapped in sedimentary rocks is involved in
the generation of oil and gas, and some of it is liberated by weathering of outcropping rocks.
Eventually, this carbon is returned, via the biosphere and the hydrosphere, back to the
atmosphere (Fig. 2.6b).
13
2 in sample
ô ,C -1 x 1000
in standard
C
The units are per mn (%o) and the standard is usually PDB i.e. a belemnite from the
t2C
Cretaceous Peedee Formation, USA (Craig, 1953). For PDB lt3c = 88.99 and its ôl3C
value = O%o by definition. Therefore negative values for a sample indicate depletion in the
heavier isotope compared with PDB.
The approximate relative sizes of carbon reservoirs (surface, crust and mantle) and their
exchange pathways within each of the three main compartments is as illustrated in Figure
2.5. The isotopic signature for the earth as awhole has been assumed to be that inherited
from the parent solar nebula as represented by the isotopic signature of the mantle (ôt'C - -
5%o: Ganels et al., 1915, Deines, 1980; Ganels and Lerman, 1984). All autotrophic
processes favour incorporation of the lighter isotopes of carbon into live organic tissue.
Consequently the isotopic signature of sedimentary organic matter is ô13C - -26Voo.
The isotopic lightness of biogenic substances results from isotope fractionation processes in
the main assimilatory pathways and, to a lesser extent, in ensuing metabolic reactions. For
example, the enzyme ribulose l,5-bisphosphate carboxylase (RuBP carboxylase) is
responsible for the primary fractionation of carbon isotopes in green plants, algae and most
autotrophic bacteria (Fogel et al.,
1988; O'Leary, 1988). In most cases, the isotopic
t'C -27%o) and atmospheric or dissolved
fractionation between photosynthetic carbon (õ -
t'C
CO2 (ô - -7%o¡ (Stuiver, lg18) is on avera ge 2O%o, a value l0%o greater than that for the
enzymatic fractionation alone (Fogel and Cifuentes, 1993).
Several models of isotope fractionation have been developed (Farquhar, 1980; Farquhar et
aL, 1982; Guy et al., 1986). These models assume that the amount of carbon isotope
fractionation expressed in the tissues of plants is dependent on the ratio of the carbon dioxide
61
Chapter 2
concentration inside the plant to that in the external environment. Therefore, when diffusion
of CO2 into the plant is a limiting process, the isotope fractionation of the plant decreases.
Typically, the õr3C values of the most plants (-20 to -30 %r) show that both diffusion and
carboxylation are limiting processes.
Plants that utilise the Ca pathway for photosynthesis use an altemate eîzyme, phosphoenol
pyruvate carboxylase (PEP carboxylase), in the first steps of carbon dioxide fixation. The
isotope effect associated with this eÍ-zyme (-2.2%o) is much smaller than that for RuBP
carboxylase (O'Leary, 1988), such that Ca-plants have more positive isotopic compositions
in the range -8 to -18%o (Smith and Epstein,l97l; O'Leary, 1981; Deleens et al., 1983).
Figure 2.7a tllustrates how the CO2 fixed by PEP carboxylase is ca:ried as part of a Ca acid
into the internal bundle sheath cells of vascular plants. Because the environment in the
bundle sheath cell is an almost closed system, most of the CO2 entering the cell is fixed back
into organic matter. Thus little isotope fractionation by RuBP-carboxylase is expressed in the
whole plant tissue (Farquhar, 1983; Fogel and Cifuentes, 1993).
Phosphoenol- ) Pyruvate
pyruvate
Figure 2.7a. Simplified diagram of CO2 fixation in Ca photosynthesis (Fogel and Cifuentes,
t9e3)
In aquatic environments the diffusion of CO2 into the plant cell is often the limiting step
(O'Leary, 1981; Raven et a1.,1987). This is due to the fact that CO2diffuses more slowly in
water than in air. Consequently, most aquatic plants have some membrane-bound
mechanism that actively transports dissolved inorganic carbon (DIC) into the
photosynthesising cells (Lucas, 1983,). Studies have shown that when DIC (mostly COz
and HCO3-) concentration in aquatic environments is low the plants will turn on an HCO3- or
CO2 pump which will accumulate a higher concentration of DIC in the cell for
photosynthesis (Fig. z.lb). The exact mechanism of transport differs from organism to
organism, but the final effect is the same. Despite living in CO2limiting environments,
aquatic plants are able to carry on high rates ofphotosynthesis.
62
Chapter2
HCq HCOr-
porter
HCq-
\
Carbonic
CO" anhydrase
moeity
Figure 2.7b. Model for dissolved inorganic carbon transport (Fogel and Cifuentes, 1993)
Sharkey and Berry (1985) reported on the effects of a DlC-concentrating mechanism on the
isotope composition of algae. When cells are growing in a high concentration of CO2 $Vo)
the carbon isotope fractionations between cells and DIC are large and similar to those seen in
terrestrial higher plants. In a low concentration of CO2 (0.03Vo), however, the concentrating
mechanism is activated whereby active transport of DIC into the cell results in large intemal
pool of substrate available for photosynthesis. Much of the accumulated CO2 does not leave
the cell before it is fixedby RuBP carboxylase. The model describing this fractionation is a
modification of that developed for Ca photosynthesis. Autotrophic organisms can be
grouped according to their degree of isotopic fractionation, largely determined by the primary
carboxylation step, as summarised by the Table 2.4beIow.
Table 2.4: Isotopic composition of major autotrophs and methanogens (Schidlowski, 1988)
Organism õ'3c %o
Higher Plants
c3 -23 to -34
c4 -6 to -23
CAM -11 to -33
Algae -8 to -35
Cyanobacteria -3 to -27
Photosynthetic bacteria
Green -9 to -21
Purple -29 to -36
Red -19 to -28
Methanosenic bacteria -6 to -41
63
Chapter 2
Different plant constituents such as carbohydrates, lignin, proteins and lipids have different
carbon isotope compositions within the same plant (Deines, 1980; Benner et al., l98l).
13C
Lipids are reported to be depleted in relative to carbohydrates (Park and Epstein, 196l).
Depletion of total plant lipid relative to total leaf tissue of terrestrial C3 plants averages 5 *
2%o (Deines, 1980). n-Alkanes are likely to provide an adequate guide to the ôt3C
composition of homologous families of straight-chain lipids as they are closely related
biosynthetically.
It is known that different species of plants synthesise differing proportions of the various r¿-
alkane homologues (Harwood and Russell, 1984) and that such differences might be
preserved in the sedimentary record (Cranwell et aL,1987; Poynter et aI., 1989). However,
no diagnostic features of n-alkane distributions were observed for a given metabolic
grouping of plants. Thus Collister et al. l199Ð suggested that n-alkane distributions from
terrestrial plants cannot be used to distinguish between Cz, Cq and CAM plants. However,
because the ôl3C value of Cz, Cq and CAM plants are known to be characteristic for each
group (Smith and Epstein, 1971; Bender, 1971; Deines, 1980), the carbon isotopic
compositions of individual n-alkanes and of total surface lipids from terrestrial plants are
potentially useful in distinguishing between these metabolic groups.
Relatively constant isotopic signatures throughout the sedimentary rock record may imply
64
Chapter 2
that a steady state in the balance between organic and carbonate carbon was reached at an
early stage in the evolution of life (Schidlowski, 1988). This is verified, for instance, by the
occurrence of stromatolites in the earliest sediments. These ancient stromatolites also signify
that a high level of productivity was achieved by the cyanobacterial mats which built them.
Such mats today can produce 8-12 g Corg l-24 lfilops and Killops, lgg3).It is also
possible that photosynthesis has improved little in effectiveness and quantitative importance
since the first microbial communities became established. The highly negative ô13C values in
Precambrian stromatolitic kerogens is thought to be related to the lack of an active transport
system for DIC at that time (Sharkey and Berry, 1985; Raven and Sprent, 19S9).
Stahl (1978) produced isotopic type curves based on the ôt3C valo"s of the saturated,
aromatic, NSO and asphaltene fractions of crude oils. He demonstrated that the l3C content
of a crude oil fraction increased with increasing polarity. Due to the structural and chemical
similarities between asphaltene and kerogen, Yen (1912) suggested that the asphaltene
fraction of an oil should have a
ttclt'C ratio very similar to that of kerogen
from which it
was derived. The general trend in the isotopic composition of both source rock EOM and oil
fractions is found to be: ôt3Ck"rog"n > ôl3Casphalternes ) ôl3cNso >
õl3Csarurates (Galimov, l9'73;Stahl, 1978).
The difference between the isotopic composition of the aromatic and saturated hydrocarbon
fraction ranges up to 3.5%o with the aromatic fraction being isotopically heavier. In very rare
cases the aromatic fraction can be slightly lighter than the saturate fraction (Sofer, 1984).
Fuex (1977) suggested that the causes for variations in the difference between these two
fractions might be related to maturity effects, whereas Stahl (1977, 1979) suggested that the
difference is related to the source of the oil. According to Fuex (1971), the C15a saturate and
aromatic hydrocarbon fractions of different oils or rock extracts can be analysed and
compared isotopically. A positive correlation is usually established when the equivalent
fractions of different oils differ by less than l%o. Bacterial degradation, maturation and
possibly migration effect on the oils, and minor inhomogeneities in the source material, are
65
Chapter 2
Source; Sofer (1984) demonstrated that the isotopic difference between the saturated and
aromatic hydrocarbon fractions (ôl3caro - ðl3csat, or Aô1) exhibits a linear relationship
which can differentiate oils of terrigenous source (Aù = 0. 1 2õr3Csat + 5.45) and those of
marine source (^ôr = 0.10ô13Csat + 3.75). The source of the oil (i.e. marine or
non-marine), the absolute isotopic value of the oil and, to a lesser extent, its maturity were
suggested to be major factors which influence its Aù value.
Maturation affects the whole oil (Fuex, lgll), such that with increasing maturity its õ13C
becomes more positive. Although, the C15* fractions of saturate or aromatic fractions are
much less sensitive to this maturity effect, Sofer (1984) noted that in a single family of oils
maturation differences could cause, a shift of up to 2%o tn the saturate fraction which would
result in a range of 0.2 to 0.25%o in the value of Aft.
Biodegradation and water washing are other factors considered to affect the isotopic
composition of crude oils. The extent of their effect depends on the isotopic composition of
the compound classes removed from the oil. Typically, biodegradation results in a partial or
total removal of n-alkanes, followed by removal of branched alkanes and some cycloalkanes
and later on aromatics. Under water washing, soluble hydrocarbons are selectively exftacted
by meteoric water moving along the oil-water contact. Low molecular weight aromatic
components (e.g. alkylbenzenes) are preferentially removed over less polar counterparts and
saturated hydrocarbons (Tissot and Welte, 1978).
Migration: Silverman (1965b) noted that due to selective adsorption of polar components
during migration, oils show an increase in paraffinity and a decrease in the NSO fraction in
the direction of migration. Accordingly, a decrease in the ôl3C value of the total crude oil was
observed and attributed to the apparent increase in the relative concentration of the l2c-rich
saturates fraction during migration. A wider spread of gasoline ô13C values, athibuted to the
redistribution of low-and high-molecular components during migration of the oils through
poorly permeable rocks, has been reported by Alekseyev et al. (1975). As the ratio of
saturates to aromatics typically increases during migration because of the preferential removal
of the more polar and water soluble aromatic compounds, Fuex (1977) suggested that
whole-oil ôl3C values decrease slightly with migration as reflection of this change in
chemical composition. Hence, selective removal of aromatic compounds which ¿re
13C
commonly richer in than the saturates, results in a concurrent slight decrease in whole-oil
õ13c.
66
Chapter 2
CnHzn*z.They are found in organic matter, oils and sediments where they may either be
directly inherited from biological material or derived from other organic compounds such as
esters, alcohols and fatty acids during early diagenesis. Their main sources are aIgae,
bacteria and land plants. Therefore the relationship between z¿-alkanes in organisms and
sediments is governed by the overlapping distributions of n-alkanes in different classes of
organisms. They are normally eluted in the saturate
fraction during column or liquid
chromatography, where they appear to be dominant compounds. Their distribution and
abundance, which in fact represent a mixing of materials from multþle sources, is usually
demonstrated by their carbon number distribution as determined by gas chromatography.
or C17 (Clark and Blumer, 1967). Non-marine algae and cyanobacteria have a characteristic
Crs-Czo range with a maximum at Ce or C1e although some species have longer chains in
the C21-C37 raîfle (Parker and Leo, 1965; Gelpi et a1.,1970).
Collister et aI. (1994) suggest that in multicomponent mixtures of n-alkanes the C16-C1e
range is likely to be dominated by algal sources; the Czo-Czø range by strong inputs from
bacterial and aquatic plants; and the Czt-Cy range by terrestrial plant waxes. Further more,
of individual extractable z¿-alkanes
these workers found that the variation in the ôl3C content
in the Crc-Czs range indicates a contribution from at least five distinct sources, namely
cynobacterial (C16-C13) with the ô13C value of -31%o;phytoplanktonic (C16-C23) with -32%o;
chemoautrophic bacterial (Czo-Czù with -38%o; phytoplanktonic or heterotrophic bacterial
(Czo-Czù with -30%o; and vascular plants (CzrCzù with-29%o.
67
Chapter 2
are also known as "chemical fossils" (Eglinton and Calvin,1967; Blumer, I9l3), "molecular
fossils and biological markers" or simply "biomarkers" (Peters and Moldowan, 1993).
The main classes of alkanes which have been used as biomarkers include normal, iso- and
anteiso-alkanes, acyclic isoprenoids and cyclic alkanes The saturated biomarker groups
identified in this study are briefly discussed below.
68
Chapter 2
Most crude oils have CPI values around 1.0. This is because the high molecular weight
by hydrocarbons from kerogen
n-alkanes inherited directly from terrestrial plants are diluted
degradation. Some oils, derived mainly or solely from terrestrial organic material which has
been reworked by bacteria and other microbes, contain large amounts of high molecular
weight n-alkanes with a moderate odd-carbon number predominance.
69
a. Regular and irregular isoprenoids
Squalane (C¡oHoz)
Botryococcane (C3a)
b. Terpanes X
X
X
c. Regular steranes
X = H, CH3, C2H5
Isoprenoids with 20 or less carbon atoms are found in sediments where their direct
introduction is possible through animals, particularly zooplankton, which feed on plants.
Several saturated and unsaturated isoprenoid have been reported in zooplankton (Blumer and
Thomas, 1965). Pristane (Crs), phytane (Czo), and smaller regular isoprenoids are primarily
derived from the phytyl side chain of chlorophyll in phototrophic organism through
reduction, oxidation and carboxylation processes (Tissot and Welte, 1984).
In very reducing environments, the ester linkage of the phytyl side chain is hydrolysed to
yield phytol which is hydrogenated to dihydrophytol, followed by a subsequent reduction
into phytane. In a normal oxygenated environment, the phytol is oxidised to phytenic acid,
decarboxylated to pristene and then reduced to pristane. \Melte and Waples (1973) and
Powell and McKirdy (1973b) suggested that phytane or pristane may be predominant
depending on the oxicity of the local depositional environment.
By assuming that phytol is the precursor for both pristane and phytane, it follows that; in
environments where bacterial activity is abundant, phytol is mineralised into CO2 and H2O
and only small fraction is preserved as dihydrophytol. Accordingly, only small amounts of
7t
Chapter 2
Apart from the primary influence of source rocks depositional environment on the pr/n-C1,
ratio, this ratio decreases sharply with increasing maturity. It reaches values below unity
over the most of the oil window (Tissot et al., 1971;Durand and Espitalié, 1973; Connan
and Cassou, 1980). Hydrocarbon expulsion has also been reported to influence this ratio,
whereby within source rock intervals of uniform kerogen type and maturity, the prln-Cs
ratio increases with increasing degree of hydrocarbon drainage into the reservoir or carrier
bed (McKenzie et a1.,1983: Leythaeuser et al.,I984a).
The carbon skeleton of 4B@)-eudesmane has been related to eudesmanol which occurs only
in higher plants (Philp, 1985) and higher plant terpenes (Peters and Moldowan, 1993)
whereas drimanes have structural features and a widespread distribution similar to many
biomarkers of prokaryotic origin (Alexander et al., 1983; Volkman, 1988). Despite its
terrestrial origin, 4B@)-eudesmane is present in only very low amounts compared to the
other sesquiterpanes in terrestrially derived Australian oils. In most cases drimane occur in
higher abundance.
72
Chapter 2
2.6.4 Diterpenoids
Diterpenoids are widely distributed in higher plants, particularly in resins (Thomas, 1969;
Simoneit, 1977). Based on structural similarities, Alexander et aL (1987) classif,red the
diterpanes most commonly found in crude oils and sediments into six families, viz: labdane,
abietane, pimarane, beyerane, kaurane and phyllocladane (Fig. 2.10a). Their relative
abundances in different plant types are shown in Table 2.5.
Table 2.5 Diterpenoid classification and abundance in plants (Alexander et al., 1987)
Bicyclic Tricyclic
Plant Type
Labdane Abietane pimaraneu Beyeraneb Kaurane PhYllocladane
Angiosperm ++ + ++ + +++
(flowering plants)
Gymnosperm ++ ++ +++ ++ +++ ++
(mainly conifers)
Pteridophytes + + ++
(ferns)
Bryophytes + + ++
(mosses and
liverworts)
a. pimaranes include isopimaranes and rimuanes;b. stachenes and hibaenes
+++ , widespread; ++ common; + infrequent; - rare or unreported.
73
11 23 ,1123
23
,ll
I
\
109 a
I
109/ 123
8p(H)-Drimane
1 2 3
ji
I
a 165 'lt Y
ßs/.y
12g 1 179 1 23 137
4 5 6
1137
123 Ìr ,z1o9
1
,,1)
r A
(
-'{ 193 \
logz 137 12g/.
8p(H)-Homodrimane
7 I 9
11 09
179
A m/2123
2
3 I
4
I
5
1 oga 9 10
10
Figure 2.9 Bicyclic sesquiterpanes commonly identified in source rocks and oils
(after Noble, 1986; Peters and Moldowan,1993).
Chapter 2
Noble et al. (1985) observed that diagenetic reduction of phyllocladene and ent-kaurene
produce mainly 16cr(H)-phyllocladane and ent-I6a(H)-kaurane, respectively. These
compounds were identified in low rank coals where the concentration of 16ø(H) compounds
were reported to decrease relative to the thermodynamically more stable 16P(H) epimers with
increasing coal rank.
Both the tricyclic and tetracyclic petroleum diterpanes t4p(H)-19-norisopimarane and IJ-
C1e
nortetracyclanel are probably derived from C2s precursors (Alexander et al., 1987). The C2e
diterpenoid acids such as sandaracopimaric acid and isopimaric acid, which are constituents
of conifer resins in particular, could readily decarboxylate in a manner analogous to the
transformation of abietic acid to fichtelite in the subsurface (Streibl and Herout, 1969;
Douglas and Grantham,1974; Barrick and Hedges, 1981).
toC36havebeen reported in ancient sediments and crude oils. (Seifert et aL,1978; Aquino
Neto ¿r aL.,1982,1983). Extended tricyclic terpanes up to Ca5 were reported in a Califomian
oil (Moldowaî et al., 1983). The structures of several homologues of the tricyclic terpanes
have been proven by synthesis (Ekweozor and Strausz 1982; Heissler et al., 1984; Siena et
al., 1984).
Aquino Neto ¿/ al. (1982) identified three of several short chain tricyclic terpanes as C1e, and
Czo. A C36 precursor was envisaged for this series, namely tricyclohexaprenol which is
formed anaerobically from a universal cell constituent, hexaprenol. The saturated counterpart
of hexaprenol has been reported in petroleum (Albaiges, 1981) and in the lipids of
archaebacteria (Holzer et aI., 1979). The C23 member series was isolated by Ekweozor and
Strausz (1982, 1983) from Athabascan Oil Sand bitumen. Tricyclic carboxylic acids (Cyr
and Strausz, 1983) have also been suggested as possible precursors.
75
247
109 t' \
\
,
- - >123 219
:l
1Og1- I 2 1 Og¿-3
4p(H)-1 9-Nor¡sop¡marane Fichtelite Rimuane
247 245\ 231 l.
191r 1 89,t \ 1 89r \
(=
-^1
23
1231'
4
1æl 1231-
lsopimarane enti-Beyerane Phyllocladane
6 16p(H)
8 16c(H)
231'ì"
g9,t 191,163 1 03Y--
1
I
(=
I
tl
\
247
1231- 1231-
ent-Kaurane Pimarane Labdane
7 16cr(H) 10 11
9 16p(H)
231ì 191,163
'1 r1
191,163
175/
(=
,
233
-ù '. )
)
/ /
1231- 123
123
17-Nortetracyclane Abietane Podocarpane
12 13 14
Figure 2.10a Bicyclic, tricyclic and tetracyclic diterpane structures identified in source
rocks and crude oils, (after Noble, 1986; Peters and Moldowan, 1993)
5
3550m
3
6
CRUDE OIL 6
4
11 78
m/2123
3673m
3
11
2
5 4
3820m 7
I
E
ã
o I
I
I
4039m
I 7
I
¿1536m
4554m
42 47
Relentlon Tlme, Mln.+
Figure 2.10b Bicyclic, tricyclic and tetracyclic diterpane peaks (in mlz 123 mass
chromatograms) identifred in source rocks and crude oils from the Gippsland Basin,
Australia (after Noble, 1986; Peters and Moldowan, 1993); their conesponding structures
are as shown in Figure 2.70a
Chapter 2
High abundances of C2a tetracyclic terpane occur in both carbonate and evaporite source
rocks and in related petroleums (Palacas et aI., 1984; Connan et al.; 1986; Connan and
Dessort, 1987; Mann et al., 1987; Clark and Philp, 1989). This compound has also been
found in Australian oils and rock extracts generated from terrigenous organic matter (Philp
and Gilbert, 1986).
78
Chapter 2
The ratios TsÆm or TsÆs+Tm have used to indicate differences in source rock type and
maturation level (Moldowan et a1.,1986; Hong et a1.,1986). The reliability of this ratio as a
source maturity indicator is best when evaluating oils from a conìmon source of consistent
organic facies. The TsÆs+Tm ratio appears to be sensitive to clay-catalysed reactions. It is
anomalously low in oils from carbonate source rocks compared to those generated from
shales of similarmaturity (McKirdy et a1.,1983, 1984; Rullkötter et al., 1985; Price et al.,
1987).
C2eTs is a C2e compound which elutes immediately after C2s l7u(H)-hopane in the mlz I91,
fragmentogram. It has been identified as 184(H)-30-norneohopane (Moldowan et al.,
1991); and its abundance is related to thermal maturity (Hughes et aI., 1985; Sofer et aL,
1986; Sofer, 1988; Cornford et aL, 1988; Riediger et al., 1990). Peters and Moldowan
(1993) described the detection of C2eTs and another rearranged hopane 17cr(H)-diahopane
(C¡o-) in the mlz l9l fragmentogram of petroleum saturates fractions. An unidentified
compound with the same retention index as C36* was previously reporte d,by Yolk,rnan et at.
(1983a) and Philp and Gilbert (1986). Both reports regarded this compound as a possible
terrestrial markerbecause of its presence in coals and terrestrial-sourced oils. However, its
bacterial origin has now been established (Peters and Moldowan,1993). The structures of
these rearranged hopanes are shown in Figure 2.8b.
79
Chapter 2
Environmental conditions are thought to influence the abundance of C2eTs and C36* such that
the C3¡*/ C2eTs ratios in oils can be used in the assessment of depositional environment of
their source rocks. Therefore, Peters and Moldowan (1993) suggested that oils derived from
shales deposited under oxic-suboxic conditions will show higher ratios than those derived
from source rocks deposited under anoxic conditions. Nevertheless, it has been shown that
17cr(H)-diahopanes are more stable than 18cr(H)-neohopanes, which in turn are more stable
than 17cr(H)-hopanes (Moldowan et aL,1991). Therefore, with increasing maturity the ratio
of 17cr(H)-diahopane to either 18ct(H)-30-norneohopane or 17cr(H)-hopane will also
2.6.8 Methylhopanes
The most prominent series of these compounds to have been identified is the 2cr-methyl-
I7a(H),2lþ(H)-hopanes (Summons and Jahnke, 1992). Each component in this series
seems to elute near the corresponding 174(H)-hopane with one less carbon. A second series
is the 3B-methyl-17cr(H),218(H)-hopanes. These occur in lesser amounts and elute much
later than the 2cr-compounds. Both series comprise Czs-Cze pseudohomologues (Summons
and 'Walter, 1990; Summons and Jahnke, 1992). The likely precursors for these
hydrocarbons are 2p-methyldiplopterol and 3B-methylbacteriohopanepolyols found in
methylotrophic bacteria (Bisseret et a1.,1985; Zundel and Rohmer, 1985). The structures of
the 2u- and 3B-methylhopanes are shown in Figure 2.8b.
2.6.9 Steranes
Steranes (Fig. 2.8c) are formed by the reduction of sterols originally incorporated in
sediments. They are widely distributed in both sediments and crude oils but not in living
organisms. Their link to biota is through their precursors, sterols which are found in both
free and bound forms in organisms. The carbon number distribution of regular (i.e.
4-desmethyl) sterols in young sediments has permitted, to a certain degree identification of
the contributing organisms (Huang and Meinschein, 1979; Mackenzie, 1984; Moldowan ¿/
al., 1985).
For instance, planktonic algae in general contain mostþ C27 and C2s sterols, although
diatoms contain approximately equal amounts of C27, C2s, and C2e sterols (Volkman et aI.,
19S1). The zooplankton, of which crustaceans are a dominant class, often contain abundant
C27 sterols.
80
Biogenic inputs
R R R
R-
+
2
3
HO HO
Â2 Sterenes 3B-Stanols A5 3Þ-Stenols
lt Via
R
steradienes
7 7
\- +
5 5
20R+20S Epimers of
5o(H), 14o(H), 17cr(H), 5a(H), 14oc(H), 17a(H)-
A4+Å5 Sterenes 20R-Steranes and 5a(H),14p(H), 17p(H)-
Steranes
The fate of sterols following the deposition and accumulation of organic mafier is discussed
by Mackenzie et aL (1982) and de Leeuw et al. (1989). The similarity of distributions in the
supposed product and precursor compound classes, and inverse abundance trends in
product-precursor pairs with increasing burial depth, make possible the identification of
specific product-precursor relationships. Simulated diagenesis in the laboratory using
isolated or synthesised individual compounds has also been used to confirm sterol
transformations.
Conversion of sterols into steranes and other products appears to involve many diagenetic
transformations (de Leeuw and Baas, 1986; Peakman and Maxwell, 1988). The subsequent
transformations are summarised in Figure 2.11. The initial steps involve liberation of free
sterols via hydrolysis of steryl esters, followed by reduction of free unsaturated sterols
(stenols) to their saturated counterparts (stanols). These processes occur under anaerobic
conditions in both sediments and particulate matter in the water column (Wakeham, 1989).
The reduction process is then followed by dehydration of stanols to sterenes; like the
preceding processes it is microbially mediated and takes place under anaerobic conditions.
The conversion of sterenes to different products follows divergent reaction pathways of
which the major products usually are steranes (Brassell, 1985). The reduction
(hydrogenation) of sterenes to their fully saturated counterparts, steranes results mainly in a
Scr(H) configuration.
In newly formed steranes the configuration at most of the chiral carbons appears to be
unaffected by diagenetic reactions, and hence is that inherited from the original biogenic
steroids. At the end of diagenesis, steranes therefore have predominantly the
5cr(H),8B(H),9cr(H),108(CH3),13B(CH,),14o(H),17cr (H) 20R configuration.
82
Chøpter 2
believed to play an important role in these processes. At a particular chiral centre (e.g. at
Isomerisation ratios have therefore been used in thermal maturation assessments. For
instance isomerisation at C-20 in the C2s 5cx(H),14ct(H),l7cx,(H) steranes causes the
20S/(20S+20R) ratio to rise from 0 to about 0.5 with increasing maturity (Seifert and
Moldowan, 1986). However, factors other than thermal maturity appear to have an influence
on this ratio. These factors include facies effects; weathering (Clayton and King, 1987); and
in situ oil biodegradation which can cause elevated ratios above 0.55 (Rullkötter and
'Wendisch,
1982;McKirdy et a1.,1983; Seifert et aI., 1984). Isomerisation at the C-14 and
C-17 positions in the 20S and 20R of C2s regular steranes causes an increase in the
crBp/(crBB+o(o(o() ratio from near zero values to about 0.7; equilibrium is inferred at O.67 to
0.71 (Seifert and Moldowan, 1986). In contrast to the previous ratio, it appears to be
independent of source facies and is slower to reach equilibrium than the latter, thus being
effective at higher levels of maturity.
The molecular formula of aromatic compounds is CnH2n-', where p varies with the number
of rings (e.g. benzeno p = 6; naphthalene, p = 12; andphenanthrene, p - 18). The aþlated
derivatives of these types comprise one to three additional carbon atoms and happen to be
major components of crude oil. In accordance with the above formula, the naphthalene type,
(CnHzn-rz) will have a distribution which has its mærimum at Cp or C13 (di- or tri-
metþlnaphthalene), and the phenanthrene type; (CnHzn-rs), a maximum at C16 or C17 (di or
'When
tri- metþl phenanthrene). several structural types are possible as for molecules with
83
Chapter 2
three or more aromatic rings, one of them is favoured in crude oils. Thus alþlanthracenes
are present in small amounts and alkylphenanthrenes are largely predominant (Tissot and
welte, 1984)
Naphthenoaromatics are another type of aromatic compound found in crude oils. They
consist of aromatic rings fused with naphthenic rings, either of which may contain alþl
substituents. These compounds are usually the major constituents of the high boiling
fractions of petroleum. Naphthenoaromatics with one or two aromatic rings are more
abundant than polyaromatics in paraffinic-naphthenic crude oils. Naphthenoaromatics are
more abundant than pure aromatics in young or shallow, immature crude oils.
Diterpenoid resin acids such as dehydroabietic acid (Hanis, 1948; Fig.2.IZb structure I) and
podocarpic acid (Campbell and Todd, 1942; Fig.2.l2b structure II) are among the most
important monoaromatic compounds which occur in plants. Genetic relationships between
diterpenoic acids in fresh kauri resins and various Recent and ancient fossil resins have been
established (Thomas, 1969). Thus, the chemical structure of fossil kauri resins has been
described as a polymerisation product of agathic acid, which is a bicyclic diterpenoid acid
with two exocyclic double bonds (Brooks and Steven, 196l; Wilson et aI., 1984).
Macrophyllic acid (Fig.2.12b structure III) from the heartwood of Podocarpus macrophyllus
(Bocks et aI.,1963) is another example of the vast variety of hydroxy-substituted aromatics
found in the plant kingdom.
84
I
Figure 2.12a Benzenoid aromatic monoterpenoid (I) and sesquiterpenoid (II, ilI)
hydrocarbons isolated from plants
Hgc OH
ll,
OH HO
OH
ITIIIIIV
Fluoranthene Pyrene Chrysene Triphenylene
Figure 2.12c Parent PAH which are more abundant than their respective alkyl-substituted
derivatives in Recent marine sediments (Meinschein, 1959)
Chapter 2
Natural fires are thought to be the major source of sedimentary PAH (Radke, 1987). PAH
derived from incomplete combustion of plant material, when absorbed on soot particles
retain their original distribution during long-range air transportation (Lunde and Bjøseth,
1977; Windsor and Hites, 1979). A lower metþlphenanthrene/phenanthrene ratio than that
typically measured in petroleum is considered to be indicative of anthropogenic combustion
PAH (Blumer and Youngblood, 1975; Prahl and Carpenter, 1983). Hites ¿/ aI. (1971)
observed that in sediments deposted during human history, an upward increase in the total
abundance of unsubstituted PAH, grossly parallels the calculated PAH production from
anthropogenic combustion.
Parent PAH (Fig. 2.12c), such as fluoranthene (I), pyrene (II), chrysene (III) and
triphenylene (IV), are more abundant than their respective alkyl-substituted derivatives in
Recent marine sediments (Meinschein, 1959). This observation contrasts with the situation
in crude oils where typically higher relative concentrations of alkyl-substituted PAH are
evident. Thus, PAH from Recent sediments are not representative of sediments in general.
Many early studies aimed at detecting primary sources of PAH in Recent sediments were
hampered by the unrecognised effects of erosion and successive resedimentation of
sedimentary organic matter inputs from oil seeps; and anthropogenic pollution, by oil spills
and the combustion products of fossil fuels. However, recent investigations of PAH in
marine environments far from industrialised areas have revealed a background of pyrolytic-
parent PAH, such as phenanthrene, fluoranthene, and pyrene (Tissier and Saliot, 1983),
impliying that pyrolytic-like PAH were probably derived partly from natural combustion
processes and partly from diagenesis.
Depth profiles of parent PAH in Recent sediment cores showed an exponential concentration
decrease with depth, in situations where, during the past eighty years, sediments had
received a continuous input of anthropogenic PAH. Deviating trends for individual PAH
of hydrochrysenes and hydropicenes, implies that
such as retene and perylene and for series
these PAH are probably of biogenic origin (Wakeham et aL.,1980b; Tan and Heit, 1981).
86
+
D EHYDROABIE TANE S IMONELLITE
II m
COOH
ABIETIC
I RETENE
VI
DEHYDROABIETIN V
IV
COOH
PIMARIC ACID PIMANTHRENE
VII VIII
Figure 2.13 Suggested precursors and several intermediates in the in situ generation of
retene (IV) and pimanthrene (VIII) (V/akeham et al., 1980b;Radke, 1987)
Chapter 2
Figure 2.13 illustrates the suggested precursors and intermediates in the in situ generation of
retene (VI) and pimanthrene (1,7-dimethylphenanthrene: VIII) (Wakeham et aL.,1980b).
Atthough abietic and pimaric acids are thought to be the respective precursors of retene and
pimanthrene, other resin constituents belonging to the abietane-pimarane skeletal types
(Thomas, 1969; Hanson, 1912) also could be sources by way of diagenetic rearrangement of
their skeletons. It is also possible that retene may be contributed from the combustion of
coniferous wood (Ramdahl, 1983).
PAH have been observed in extracts of oil shales (Anders et aI., 1913) and coals (Hayatsu e/
aL, 19'18; White and Lee, 1980; Radke et a1.,1982b), and in retort oils (Ingramet al.,
1933). These aromatic compounds, however, were found to be more difficult to correlate
with natural precursors than is the case with saturated biomarker molecules. The occurrence
of a substantial concentration of retene, along with fichtelite, in wood stumps sunk in peat
(Skrigan, 1964) suggests that the PAH might have formed from unsaturated terpenoid
88
Chapter 2
Examples of PAH compound types commonly found in crude oils are shown in Figure 2.14.
Anthracene homologues are not normally major components, although a significant
concentration have been encountered in certain oils (Camrthers, 1956). Phenanthrenes
generally dominate over anthracene type compounds (Smith, 1967) and the
phenanthrene/anthracene concentration ratio is about 50 (Grimmer and Böhnke, 1918).
Variations in the concentrations of the prominent PAH in crude oils are suspected to be
related to differences in the maturities of the oils (Grimmer and Böhnke, 1978).
89
IV XV
VIII
/
+
_-+
II
XII
VI
IX
+
ilI XIII
I
X
H2
H2
---->
IV VII XI XIV
Figure 2.14 Aromatic compound types prevailing in crude oils and their structural
relationship. Broken ¿urows point toward PAH types of minor importance (Grimmer and
Bohnke, 1978; Radke, 1987).
1.80
1 I 1 0
I
2 2.12
7 2 2.18
6 3
5 4 65 43
1.96
Naphthalene Phenanthrene
Figure 2.15 Dewar's reactivity numbers, Nr (bold figures) for naphthalene and
phenanthrene (Dewar, 1952). Positions are designated by integer numbers. Sterically
crowded positions where reactivity is lower than indicated by the Nr value are marked
by asterisks (after Radke, 1987).
Chapter 2
The net effect of those reactions is to produce, from a complex mixture of alkyl- and
cycloalkylaromatics, an aromatic fraction that becomes increasingly depleted in hydrogen as
thermal evolution proceeds, and a hydrogen-rich saturated hydrocarbon fraction (Alexander
et aL.,1980). The following discussion draws heavily on Radke (1987).
Table 2.6. Maturity parameters based on aromatic hydrocarbons (after Radke, 1987)
TNR-2
1,3,5-TMN+[ 1,3,6-TMN+ 1,4,6-TMN
Metþlphenanthrene to
MPR-I phenanthreneratios
MPR-2 2-MP
P
MPR-3
MPR-9 9-MP
P
MPR
91
Chapter 2
C15* hydrocarbon yields are an indication that the generation of heavy hydrocarbons from
Type trI kerogen is accompanied by methyl-transfer reactions (Albrecht et a1.,1976).
Provided that phenanthrene has been involved in the metþl-transfer reactions taking place
concomitant with hydrocarbon generation, then the coincident depth trends of 1- and 9-
metþlphenanthrene to phenanthrene ratios (MPR-l, MPR-9) may be explained on the basis
of the very similar reactivities of the C-1- and C-9 positions (Fig. 2.I5). Metþlation may
have involved also possible phenanthrene precursors, such as hydrophenanthrenes. A
conversion of alkyldihydrophenanthrenes to alkylphenanthrenes with thermal evolution has
been reported in the organic extracts of coal macerals (Allan and Larter, 1983).
for crude oil maturity was invented (Radke and Welte, 1983). MPI-I values of the vast
majority of the crude oils fall in the 0.35-l.OOVo range, corresponding to Rç values of 0.61-
l.O)Vo. Different oil types have tentatively been classified into maturity categories with
respect to Rs value (Radke, 1987). Immature oils are those with Rç values <0.7OVo; whereas
mature oils have values of 0.7-0.95Vo, and postmature oil values of >0.957o. From bulk oil
parameters, such as API gravity and gross composition it is evident that thermal evolution of
their respective source beds had probably not passed beyond Ro = l.35%o at the time the oils
were released. Furthermore, the maturity of an oil is likely to be related to the kerogen type
'Welte (1984)
in the source rock from which it was generated. Tissot and related the kerogen
type to the composition of the crude oil it produces, thus inventing analogous oil types.
Accordingly, different oil types may be distinguished based on their Rc values and bulk
composition. It is usually assumed that the fu value of a given oil will correspond to the R6
value of the source rock at the time of expulsion. This will be true provided that the original
distribution of phenanthrene and its metþl homologues was retained during primary and
secondary migration of the oil (Radke, 1987).
92
Chapter2
Type I oils belong to the paraffinic class and commonly have high wax and low sulphur
contents. They are derived from Type I kerogen (sapropelic), which has a liptinite content
>9OVo; hydrogen index > 650 mg HC/g TOC and oxygen index <25 mg COZlg TOC; and
FVC atomic ratio >1.5 and O/C atomic ratio <1.0. The liptinite precursors are largely
freshwater algae and the biodegraded remains of terrestrial plants which, together with
microbial biomass, were deposited in a eutrophic deep-water lacustrine palaeoenvironment.
The Rç distribution of this oil type ranges from 0.57 to 0.7I7o with its mode atO.65Vo.
Alkylphenanthrenes are virtually absent in the oils released from source rocks containing
Type I kerogen because the major proportion of the kerogen is liptinitic with a structure
known to be mainly aliphatic at the immature stage. Secondly, polycyclic aromatic structural
units that are formed during catagenesis tend to remain incorporated in the three-dimensional
kerogen network, and thus do not contribute significantly to oil formation. Therefore, the
available alkylphenanthrenes are generated from a minor vitrinitic component of the kerogen
and hence are similar to those of the soluble organic matter derived from immature Type III
kerogen. Alternatively, they are picked up by the oil on its passage through adjacent Type Itr
kerogen-containing strata.
Type l/il oils are aromatic-intermediate crudes which exhibit lower wax and higher sulphur
contents than Type I oils. They are probably derived from a subtype of Type II kerogen,
defined as Type IIn by Radke et al. (1986). The liptinite content of this kerogen is high, as is
its as initial hydrogen index, whilst the oxygen index may have moderate values (up to 50
mg HC/ g TOC). It is composed almost entirely of biodegraded phyto- or zooplankton
deposited in an anoxic shallow marine environment. The Rç distribution of its derived oils
ranges from 0.66 to 0.\lVo, with the mode around Rc 0.75 + 0.05Vo. The aforementioned
Rs interval is in agreement with the generally accepted view that significant oil generation
and expulsion from source rocks containing Type II kerogen occurs at0.7-0.8Vo Rq' The
presence of alkylphenanthrenes in Type II kerogen is attributed to amorphous liptinite
(Radke et aL,1986).
Type oils are not grouped under any specific class defined by Tissot and'Welte (1984).
II/¡il
The oils are derived from mixtures of detrital Type II and Itr kerogens, with the former being
predominant. Elevated liptinite contents are commonly observed along with elevated
hydrogen index values in the range 100-300 mg HC/g TOC. Oxygen index values are very
variable but generally greater than in the aforementioned types.'The kerogen is commonly
composed the remains of aquatic organisms and terrestrial plants deposited in moderately
anoxic, lagoonal or deltaic environments. The Rç distribution of these oils ranges from
0.75-0.92Vo, with a mode atO.827o.
93
Chapter 2
Type III/il oils arc also not grouped under the Tissot and Welte (1984) classification. These
oils are presumably derived from a mixed Type II/I[
kerogen with a strong terrestrial
organic matter input. Their Rs distribution has a mode at 0.907o and the maximum value at
l.02Vo. This mode, however, may conespond to a second maximum (Radke, 1987: Figure
19) of oil expulsion from the same Type II/III kerogen which have already released Type
II/III oils.
Nevertheless, conflicting results were obtained for immature samples containing low quality
kerogen in that their MPR values showed considerably higher maturities than those obtained
from the MPI values. It was later revealed that these samples contained pyrolytic-like PAH
distributions which are widespread in unpolluted Recent sediments (Tissier and Saliot,
1983). Such distributions are marked by the predominance of phenanthrene, fluoranthene
and pyrene over their respective alþl homologues. Thus, the relatively low MPI values in
of pyrolytic phenanthrene. As in Recent
these sediments are caused by their high abundance
sediments, a background level of pyrolytic-like PAH exists in ancient sediments but here
they are normally outweighed by PAH derived from catagenesis.
The proposed shift of cr-methyl groups to p-positions at elevated temperatures (Radke et al.,
1982b), coincides well with the observed increase of MNR values with rank. However, the
relative increase of 2- over l-methylnaphthalene does not necessarily mean a conversion of
one compound to the other, as there could be other possible precursors. It has been argued
94
Chapter 2
that cr-substitution leads to the greater reactivity of the naphthalene homologues in general,
which may be prompted to enter into other reactions (e.g. condensation) besides
reamangement (Madison and Roberts, 1958). Conversions of dimethylnaphthalenes with
two o(-methyl groups, (e.9. I,4-, I,5- and 1,8-DMN) are likely to be controlled by similar
chemical principles. The abundance of kinetically more stable cr,B- and BP- Dl\rß{ isomers
increases with thermal maturation relative to those with two o-methyl groups.
The selective losses of PAH components from the extractable organic matter of coal has also
been attributed to water washing (Radke et aI., 1982b). A preferential depletion of
components with relatively high solubilities in water, such as mono- and
dimetþlnaphthalenes and benzothiophenes, was reported in the organic extracts of coals
samples derived from water-drained localities. The same samples seemed to be also depleted
in phenanthrene, whilst the methylphenanthrene homologues apparentþ were not affected.
95
Chapter 2
As yet the possible effects of biodegradation on PAH maturity parameters appear to have not
been investigated in much detail. However, it has been reported that biodegradation also
results in preferential loss of individual alkylnaphthalene and alkylphenanthrene isomers
(Volkman et al., 1984; V/illiatns et al., 1986) which likewise would affect the respective
maturity parameters.
96
Chapter 3
Chapter 3
Analytical Methods
3.L Introduction
The ability of the petroleum geochemist to answer geological and other queries which arise
during an exploration program is entirely dependent on results of various geochemical
analyses. Usually these analyses are carried out on rock cuttings, side wall cores, pieces of
outcrop, formation waters drilling muds and fluid hydrocarbons (oil, condensate and gas).
To meet the objectives of this study, rock cuttings and core samples of the Poolowanna
Formation from selected wells in the Patchawarra and Poolowanna Troughs, and crude oils
from adjacent reservoirs of the Eromanga and underlying Cooper and Warburton Basins
were examined (Tables l.I,l.2 and 1.3). The analyses carried out are summarised in Figure
3.1.
9l
Cuttings Oils
Gonventional Cores
oz
Ø<
Rock-Eval Pyrolysis
Solvent Extraction
Column Chromatography
Molecular Sieve
ils on
Gas Chromatography-
Mass Spectrometry (GC-MS)
Figure 3.1 Flow diagram for source rock and petroleum geochemical analyses
Chapter 3
The rock sample is prepared from either rock fragments (cuttings) or core pieces which are
mounted in a block of cold setting resin and then ground and polished to give a flat surface.
The polished block is examined under immersion oil using a reflected light microscope and
the macerals are identified by their relative reflectance, colour, morphology and fluorescence
characteristics. The proportion of each maceral is determined by the point-counting method.
99
Chapter 3
stage. Counts were acquired by traversing the block and alternating the light modes on every
move. At least 100 counts were made on every block. The relative abundance of each
maceral was reported as percentage of the sum of all the macerals identified in the block. In
addition the content of pyrite was reported as a percentage of the total macerals plus pyrite
counts.
100
o 550'c
o
o-
o
-)
(ú Tmax
o
o-
E
.o
300 "c
FINISH START
S2
o
U)
c S3
o
o-
U)
o
É.
o
C)
o
P
o
o
S1
In the first step, hydrocarbons present in a free or adsorbed state (S1) are volatilised at
300'C. The amount of these hydrocarbons is measured by a flame ionisation detector (FID).
In a second step, the temperature is ramped from 300 to 500"C, whereupon hydrocarbons
and hydrocarbonlike compounds (Sz) and oxygen-containing volatiles, i.e. carbon dioxide
(S3) and water, are generated as result of kerogen pyrolysis. The 52 compounds are then
measured through a FID, whilst 53 is measured through a thermal conductivity detector. The
measurement of 53 is limited to a convenient temperaturewindow (300-340"C) so that only
the main stage of CO2 generation from kerogen is included, and to avoid inorganic sources
of CO2. This analysis was also carried out by Geotechnical Services Pty Ltd. Only samples
with TOC values >L)Vo were analysed.
The various parameters measured, including Tr* are: hydrogen index (HI), oxygen index
(OI), production index (PI) and potential yield (S1+S z), are defined in Table 2.3.
V/t EOM
VoEOIII =-x100
Wt Sediment Extracted
Wt EOM (mg)
ppm EOM
V/t Sediment Extracted (kg)
102
Chnpter 3
A small quantrty (ca 50 mg) of each rock extract or oil sample was dissolved in DCM and
then transferred onto dry activated alumina. The mixture of the sample and alumina was then
left open to the air allowing the solvent to evaporate; occasionally the mixture was slightly
heated to less than 40'C to facilitate solvent evaporation. Then, the dry adsorbed sample was
gently poured onto the top of the column.
The excess solvent in the three eluted fractions was removed on a rotary evaporator. The
fraction concentrates were then transferred to labelled, pre-weighed vials where they were
dried in open air and then weighed. The weight of each fraction was used to calculate the
percentage of each fraction in the rock in accordance with the following formulae:
Wt Fraction
VoFruction =-x100
Wt of all Fractions
Wt Fraction (mg)
ppm Fraction
V/t Sediment Extracted (kg)
103
Chapter 3
Diluted samples (ca 1 mg/150 pl) in n-hexane were injected using a split /splitless injector,
operating in split mode. About 1.0 pl of sample solution was injected. The carrier gas was
hydrogen flowing at a linear velocity of 30 cm s-1. Both the injector and the flame ionisation
detector (FID) were maintained at a temperature of 300"C. The oven temperature was held at
104
Chapter 3
40'C for 5 minutes and then programmed from 40 to 300'C at 4"C min-l and then held at
The detector response data were acquired using a Maclab System comprising MacLabl{
hardware in conjunction with Chart software (Analog Digital Instruments, Sydney)
connected to a Macintosh SE computer. Chromatographic data processing including
determination of peak areas and heights were carried out using Peaks software (Analog
Digital Instruments).
The silicalite f,iltration method of Flannigen et al. (1978) and West et aI. (1990) was applied.
Silicalite pellets were ground to powder using a pestle and mortar and then activated by
oven-heating overnight at 500"C. Columns of silicalite were prepared by plugging a Pasteur
pipette with a small piece of cotton wool and adding about 2 g of silicalite powder. A2 mg
sample of saturated fraction dissolved in n-hexane (ca. 2 n'l) was poured onto the column
and allowed to stand for 2 minutes before the column was eluted with at least three bed
volumes of n-hexane into a pre-weighed vial. The silicalite plug was thoroughly washed
with eluting solvent to avoid the retention of acyclic biomarkers. The eluate (non-adduct)
was then dried in air and weighed ready for GC-MS analysis.
105
Chapter 3
trimetþlnaphthalenes were identified from their relative retention times on the reconstructed
ion count (RIC) chromatogram. Quantitation was by measurements of peak heights.
The XRD scans revealed no significant levels of carbonate minerals in any of the rock
samples.
106
Chapter 4
Chapter 4
4.I Introduction
Maceral distributions in coals and the dispersed organic matter (DOM) of associated clastic
sediments in particular the types of maceral present, their relative abundance and their
-
fluorescence characteristics - are useful for qualitative assessment of their source rock
potential.
The total amount of macerals (corrected for inertinite content: Horstman, 1994) gives a good
estimate of source rock richness. The relative proportions of the maceral groups, (viz.
vitrinite, liptinite or inertinite) give an indication of the kerogen type. Thus, a prediction of
whether the source rock is oil or gas-prone can be made.
In this chapter, the organic petrography of samples from selected wells in the Poolowanna
and Patchawarra Troughs is discussed with a view to evaluating the source rock potential of
the Poolowanna Formation.
101
VITRINITE
Field
o Poolowanna
A Tantanna
\
o Sturt
\
\ tr Sturt East
\
\
\ g
o
\ o
oo o
o o
,'p o
I
Âo t
tr o I
tr
o I Â,
o
 0 lo A
oÔ o Â,
A
Ä o
,
t
o
LIPTINITE INERTINITE
Figure 4.1 Distribution and relative abundance of maceral groups in the Poolowanna
Formation
Chapter 4
Examples of telocollinite are shown in the photomicrographs of Plates lA,2A,6G and 10C.
Desmocollinite is shown in Plates 1E, and 2G. Under blue light irradiation it occasionally
fluoresces (Plate 1F). This fluorescence is possibly caused by disseminated liptinite
macerals, such as resinite or bituminite.
109
Table 4.1 Relative abundance of maceral groups and their corresponding maceral contents
Vifinite Liptinite
no. (m) Tc Dc. Cg Vd Total Sp Cu Re I'd Al Bi Ex FlTotal Fu Sf Id I,{a Nfi Sc Total
(Vo) (Vo) (Vo) (Vo) (Vo) (Vo) (Vo) (Vo) (Vo) (Vo) (Vo) (Vo) (Vo ) (vo) (vo) (vo) (Vo) (Vo) (Vo) (Vo) (Vo)
-p
POLI-2 Poolowanna-1 2423 .6 25.2 75 2.6 p 2.6 4.3 5.2* 0.9 t6 0.9 4.3 4.3 -p l0
POLl-3 2438 t9 .8 19.8 ; 40 15.9 1.0 0.1 10.7 0.6 1.5 30 4.7 15.8 10.1 pp 3l
tt
Poolowanna-1
POLl-4 Poolowanna-1 2469 6. 4 33.7 p 40 12.r 1.4 0.7 l2.l 0.7 p I 27 8.5 T2.T 12.4 33
POLl-5 Poolowanna-1 2499 3. 0 19.2 p 22 10.4 r.6 31.5 22.7 p p 66 2.2 6.3 3.0 _ t2
POLl-6 Poolowanna-1 2545 19 .7 17.0 p 37 8.9 t.2 5.9 ll.7 p p 28 ro.2 16.0 9.5 36
POLI-7 Poolowanna-1 2557 9. 4 9.8 19 10.8 1.79.4 17.2 p ; p 39 5.1 10.8 25.9 .;. 42
POL2-8 Poolowanna-2 2496 9. 5 5.8 I 15 7.3 p 44.017.8 p p I 69 3.8 4.5 7.3 pp t6
POL2-9
POL2-10
Poolowanna-2
Poolowanna-2
2524
2542
7. 8
2 .0
26.4
15.4
34
37
5.6 0.6 19.616.1 p
5.9 0.3 r5.7 5.6 p
p
p
p
I
42
28
r.9
6.6
10.9 tr.2
8.4 t9.9 t_l 24
35
POL3-11
POL3-12
Poolowanna-3
Poolowanna-3
2423
2438
t9 .7
T6 .2
15.6
9.9 T
35
26
4.8 1.5 t4.l 3.3 p
8.4 0.5 22.5 Ll.O p
p
I
24
42
1.1
0.5
28.6 rl.2
9.4 21.5 _t_
-p
4t
3I
POL3-13 Poolowanna-3 2505 3 6 5.2 p 9 7.8 2.6 27.r t9.8 0.5 58 1.0 8.3 24.0 33
POL3-14 Poolowanna-3 2551 0 .2 r4.3 p 4 3.0 7.8 4.4 2.4 p p 18 4.0 20.5 13.5 pp 38
I
POL3-15
TAN1-16
Poolowanna-3
Tantanna-1
2560
1795 10.8
.3 tr.4
9.4
p
I
36
20
6.1 r.2 15.3 7.0 p
1. 7 4.3 2.2 18.7 2.2
6. 6 7.t 2.7 25.9 p
p
8.6
p
30
68
72
tt.2
t.4
0.9
8.0 15.5
2.9 7.9
2.7 16.1
_l 35
t2
20
TAN2-17 Tantanna-2 1796 2.7 5.4 8
TAN2-18 Tantanna-2 r799 9.0 10.0 p 19 I 3. 4 2.7 0.2 13.2 6.5 I p 36 22.t 6.1 16.8 p 45
TAN2-19 Tartanna-2 1805 p nd -p I nd 3.9 15.6 80.5 nd
TAN2-20 Tantanna-2 181 1 3.6 7.7 p 11 I 4. 0 5.9 8.6 t7.r p 3.6 p 49 8.1 13.1 r8.5 --- 40
TAN3-21 Tantanna-3 1807 11.5 15.3 p 27 7 I 0.9 0.9 7.1 p p T6 7.4 38.3 11.5 pp 57
TAI:{4-22 Tantanna-4 1807 14.0 15.3 p 29 3 5 0.2 0.5 r.6 p 0.6 I 6 20.6 38.3 5.3 64
TAN4-23 Tantanna-4 T8T4 9.3 12.7 p 22 I 5. 0 1.4 0.0 9.1 r.4 1.7 29 t6.7 7.9 24.6 49
TAN5-24 Tantanna-5 1814 8.0 7.3 p 15 1 2. 3 0.9 2.1 7.0 r.6 1.1 25 27.5 14.6 r7.6 : 60
TANS-25 Tantanna-8 Í1823 76.0 6.3 82 2 I 0.6 0.2 0.6 p 1.0 5 2.7 8.5 t.9 l3
STUl-26 Sturt-1 1865 4.1 t2.0 I t6 I .4 3.2 p 19.0 1.2 0.9 I p 66 4.4 5.5 8.5 -p 18
STUl-27 Stuf-1 t87r 4.2 1.8 p 6 3 .6 1.8 1.8 18.1 1.8 3.6 81 4.2 3.6 5.4 t3
Table 4.1 (continued)
Vitinite Liptinite
no (m) Tc Dc. Cg Vd Total Sp Cu Re Ld Al Bi El( Fl Total Fu Sf Id N{a Nß Sc Total
(Vo) (Vo) (Vo) (7o) (Vo) (7o ) (Vo) (Vo) (Vo) (Vo) (Vo) (Vo) (Vo) (vo) (vo) (vo) (Vo) (Vo) (Vo) (Vo) (Vo)
p p p p
STU2-29 Sturt-2 1856 3.2 p 3 3 3 p 6 327.0 I 6 p p 68 p p 28.6 29
STU4-30 Sturt-4 1859 t2.r 8.3 20 13.7 1.1 I 4.5 0.3 1 6 23 27.4 22.2 7.0 57
STU4-31 Sturt-4 T1881 3.2 18.3 I 52 2.6 0.3
1
2.9 0.9 p p I 7 t4.r 23.6 4.0 p p p 42
Sturt-3 1859 2.8 1 1.3 p t4 16.0 2.8 8.5 18.9 6.6 53 16.0 6.6 10.4 p 33
STU3-32
STU3-33 Sturt-3 1862 11.0 15.9 p 27 5.2 0.8 1.4 5.6 0.2 0.2 I T3 33.2 18.6 7.9 p p 60
STU6-34 Sturt-6 1847 t2.9 18.1 p p 3l t2.2 0.2 1.1 7.4 0.6 0.6 p p 22 25.0 15.0 7.0 47
STU6-35 Sturt-6 Tr865 18.7 19.7 p p 38 1 9 p 1.5 1.9 p p 5 28.1 24.7 3-6 p - p 56
STUS-36 Sturt-8 r862 10.0 19.2 p p 29 3 3 0 1 0.8 1.9 p 0.1 p p 6 37.7 23.6 3.3 65
STEl-37 Sturt East-l 1835 9.4 L4.7 p 24 6 7 p 1.1 8.3 p 0.5 p 17 36.r r6.8 6.4 p p 59
STEl-38 Stuft East-l r84l 1.2p p 1 58 I 0.5 t.0 24.5 3.2 p 87 5.1 1.7 4.7 p 12
STEl-39 Sturt East-l 1.5 8 8 0.0 1.5 26.8 5.4 0.5 83 7.8 2.0 5.9 - p l6
STE2-40 Sturt East-2
184ø.
t829 5.8 4.8 I
1
11 38 ) t.2 1.2 19.4 r.8 0.9 I 63 12.t 6.t 8.5 27
STE3-4I Sturt East-3 1850 2r.6 3.5 p 25 31 7 2.5 3.5 16.1 0.5 0.5 p p 55 5.5 6.0 8.5 20
STE3-42 Sturt East-3 1865 2.3 2.3 p 5 1 2. 0 4.6 3.4 14.9 p t.7 p 37 33.1 10.9 14.9 59
STBl-43 Sturt East-4 1838 3.1 8.0p p 11 9 3.6 2.7 17.8 2.7 0.9 p 60 12.9 r0.4 5.1 p 28
STE4-44 Sturt East-4 T1 r.4 30.2 p p 52 2.0 0.6 1.5 2.6 p 0.1 p p 7 14.0 2t.r 6.5 p 42
STE4-45 Sturt East-4 T1 14.3 16.6 p 3T 0.7 0.2 2.4 1.5 p 5 37.9 2r.t 5.3 64
average abundance of DOM in the Poolowanna Formation of the Sturt and Sturt East
13.7Vo.
Fields contains an unusually high abundance of sporinite (up to 58.lVo in Sturt East-l,
1841 m).
Different shapes and sizes of the spores were observed, possibly indicating contributions
from various plant species. The common microsporinite (or tenuisporinite) in DOM occurs
as miospores, as illustrated in Plates 3G-H and 9C-D. It is translucent in white light, and
therefore can be identified only after blue light excitation. This miosporinite is likely to be
contributed from pollens.
Resíníte
Resinite appears to be quite common in the Poolowanna Formation. Average relative
abundance s of. l4.3Vo and 2.57o were found for the Poolowanna and Patchawarra Troughs,
respectively. An unusually high abundance was noted in several sections of the Poolowanna
field. The highest values recorded were 447o in Poolowanna-2 and3l.5Vo in Poolowanna-l
(Table 4.1).
Rod-like shapes possibly representing resin ducts are noted in a sample from Poolowanna-1
(Plate 1C-D) and in the same section resins appe¿ìr to fill the cell cavities of plant tissue (Plate
lG-H). In the latter case, the host telocollinite is impregnated with resinite thus showing a
tt2
Chapter 4
Cutinite
Cutinite also appears to be ubiquitous throughout the Poolowanna Formation although it is
somewhat less abundant than the aforementioned liptinite macerals. Its average abundance is
2.2Vo tnboth troughs. Most of the identified cutinites are thick and long (Plates 2C-D and
4A-B). A thick cutinite, possibly deformed by syngenetic pyrite as shown in Plate 104-8.
Its compressed cell structures, display a strong orange to brown yellow fluorescence which
is probably enhanced by fluorinite excretion as result of compaction.
Cutinites of medium thickness (Plate 6G-H) and an unusually long, irregularly compacted
leaf cutinite (Plate 7A-B) were identified in DOM from the Sturt field. Thin cutinites are rare.
In the Poolowanna Trough they are associated with lamalginite (Plates lF and 2B); and in
the Sturt Field they occur as isolated stringers (Plate 6A-B).
Alginite
The abundance of alginite is very low compared to the other liptinite macerals. It has an
average abundance of in the Poolowanna Trough and I.7Vo in the Patchawarra
I.5Vo
Trough. I-amnlginite appears to be very rare and was only identified in the Poolowanna
Trough. A relative abundance of 5.2Vo (Table 4.1) was recorded in silty shale from
Poolowanna-l. The general morphological features of lamalginite are illustrated in Plates lF
and 28.
tt3
Chnpter4
scattered through the mineral matrix globules (Plate 6A-B and E-F). Sturt Field appears to
have the highest abundance of fluorinite.
4.1). It is also overall more abundant in the Patchawarra Trough (38.9Vo) than in the
Poolowanna Trough (29.8Vo). This maceral group consists of almost equal amounts of
semifusinite and inertodetrinite (l2.9Vo and l2.8Vo, respectively) while fusinite is the least
abundant of the three macerals (Table 4.1). Inertinite relative abundances )50Vo were
recorded in samples from Tantanna, Sturt and Sturt East (Fig. 4.1).
It commonly occurs in association with the vitrinite and liptinite maceral groups (Plates 54,
6C and 10C). In some carbonaceous shale facies, it occurs as a monomaceralic layer
bounded by liptinite macerals, e.g. miosporinite in a bituminite groundmass (Plates 9C-D
and 4A-B) and telalginite (Plate 4G-H).
Semifusiníte
Semifusinite is similarly widely distributed. It has an overall abundance of l2.9Vo (Table
4.1). In the Poolowanna Trough, semifusinite is more abundant than fusinite (lI.9Vo) and
less abundant than inertodetrinite. In Patchawarra Trough, it is less abundant than fusinite
but slightþ more abundant than inertodetrinite.
rt4
Chapter 4
Semifusinite is identified by its swollen cell walls, a result of humification, and a lower
reflectance than that of coexisting fusinite. The characteristic colours in incident light range
from light grey to white. Like fusinite it commonly occurs in the form of layers (Plates 24,
6C and 10C) and lenses in which the cell cavities are either empty or filled with a wide range
of substances including gelovitrinite, gelo-inertinite, resinite and various minerals.
Sclerotiníte
Sclerotinite is very rare in this part of the Poolowanna Formation. It was sparsely identified
in Poolowanna-2 and in a some Sturt and Sturt East field samples (Table 4.1).
4.2.4 Microlithotypes
A microlithotype denotes the form in which organic matter as macerals appears under the
microscope in a manner analogous to the description of coal bands in hand specimen (i.e.
lithotypes). Under ICCP conventions a microlithotype can only be recorded as such if, on a
polished surface perpendicular to the bedding plane, it has a width of at least 50 microns or
covers a minimum area of 50 x 50 pm. Also, a minimum quantity of 5Vo is required for the
maceral to be incorporated in the microlithotype. Microlithotypes are subdivided into three
groups, viz. monomaceral, bimaceral and trimaceral.
115
Table 4.2 Mac,eralassociation and pyrite contents
84 85 31 "/ifrincrlnrecil
I
POL2-9 Poolowanna-2 2524 76 66 58 2.r3
il
POL1-5 Poolowanna-l 2499 88 78 34 4.45
POL3-12 Poolowanna-3 2438 69 74 58 0.52
POL3-13 Poolowanna-3 2505 67 91 42 7.25
TANI-16 Tantanna-1 1795 88 80 32 Vitrinertosporite nd
I
TAN2-17 Tantanna-2 1796 80 92 28 nd
I
TAN2-20 Tantanna-2 181 I 60 89 51 nd
il
STU1-26 Sturt-l 1865 82 84 34 nd
STU1-27 Sturt-l 1871 87 94 t9 lt
t.78
il
STU2-28 Sturt-2 1829 79 81 40 2.O4
il
STU2-29 Sturt-2 1856 7l 97 32 8.70
STU3-32 Sturt-3 1859 67 86 47 5.36
STE1-38 SturtEast-l 1841 88 99 13 t.45
STE1-39 Sturt East-l 184ø. 84 99 t7 0.49
STE2-40 Sturt Eåst-2 1829 73 89 37 Vitrinertosporite 0.90
il
STE3-41 Sturt East-3 1850 80 75 45 1.00
il
STB1-43 Sturt East-4 1838 72 89 40 o.66
Monomaceral microlithotypes are characterised by not less than 95Vo content of a given
maceral group. They includevitrites which contain not less than 957o vitrinite and not more
than 5Vo of liptinite and/or inertinite; Iiptites containing more than 957o liptinite (e.g. sporite,
resite, liptodetrite, cutite and algite); and inertites containing more than957o inertinite (e.g.
fusite, semifusite and inertodetrite).
Trimaceral microlithotypes are the only ones in which all three maceral groups are present to
extent of more than 5Vo. They are subdivided into:
(a) duroclarite, in which vitrinite is more abundant than liptinite and inertinite;
(b) clarodurite, in which the proportion of inertinite is greater than that of vitrinite plus
liptinite;and
(c) vitrinertoliptite (Marchioni,l976), in which liptinite is predominant.
In these trimacerites the specific origin of the liptinite can be emphasized by adding the
liptinite maceral type, to the description. For example, a'duro-sporoclarite'is a trimacerite in
which vitrinite exceeds the inertinite content whilst sporinite is the main liptinite.
A combination of two maceral groups or bimaceral association (Table 4.2) was obtained by
adding together their relative abundances as determined from point counting (Table 4.1). The
combination of inertinite and liptinite seem to be most abundant with an average value of
7l.8%o, followed by vitrinite and liptinite (65.7Vo), and vitrinite and inertinite (62.4Vo). The
combination showing the highest mean value in the Poolowanna Trough is that of vitrinite
and liptinite (7O.2Vo), whereas in the Patchawarra Trough it is that of inertinite and liptinite
(77.6Vo).
118
VITRINITE
100 %
T v
T
Oa
T
LIPTINITE ¡NERTINITE
a 0.1-1.0 %
V Vitrinertoliptite
C Clarodurite
Figure 4.2Maceral group abundance for some DOM of the Poolowanna Formation,
rep-lotted to illustrate.its rèlationship to microlithotype compositions and pyrite
abundance.
Chapter4
In summary, the source rock facies of the Poolowanna Trough can be characterised by the
vitrinertoresite and duro-sporoclarite maceral associations, whereas in the Patchawarra
Trough the source rocks have vitrinertosporite and claro-sporodurite signatures. The
Poolowanna Trough source facies are strongly influenced by resinite and vitrinite, whereas
in the Patchawarra Trough there is a strong inertinite and sporinite association.
r20
Chapter 4
derived from inorganic matter contained in the coal-forming plants; for example, dispersed
opal (opaline phytolith) and fine crystalline quartz (David et aI., 1984). The second group
comprises all minerals which were washed into the coal swamp as detrital fragments. The
third group consists of inorganic matter precipitated from pore or surface water ('Ward,
1936). The minerals resulting from the second and third groups are referred to as
adventitious (Francis, 1961). However, a clear distinction between the inherent and
adventitious groups of minerals is not always possible. Commonly, both types of mineral
matter occur together. Although the list of minerals found in coal and carbonaceous
sediments is long, only pyrite and to some extent clay and carbonate minerals are commonly
identified.
4.2.5.1 Pyrite
Under the microscope, pyrite is easily identified by its very high reflectance and its yellow to
bronze colour in ordinary white light. It occurs in several forms, including granules to small
nodules in vitrinite (Plates 1C and G) and as precipitated concretions infilling cell cavities
(Plate 1C). Large syngenetic nodules in clastic rocks (Plates 1E and 9E) that deform the
adjacent laminae are also common. Framboids in clastic sediments (Plate 8C) and carbopyrite
(Plate 11C-D) are examples of this feature.
The abundance of pyrite relative to the sum of the associated macerals was determined by
point counting. The general abundance and distribution of pyrite is as shown in Tables 4.2
and 4.3 and Figure 4.2.The highest values recorded were 7 .25Vo in the silty shale from the
Poolowanna Trough and 8.707o in the silty shale at Sturt-2 (1856 m) in the Patchawarra
Trough. Generally, the vitrinertoliptite facies appears to contain the most pyrite, except for a
few vitrinertosporite samples from Tantanna, Sturt and Sturt East which appear to have very
low to low pyrite contents.
The duroclarite facies typically has a low to moderate pyrite content, whereas the clarodurite
facies is the least pyritic. It is very interesting to note that the Poolowanna Formation in the
Tantanna Field is devoid of pyrite, regardless of its organic facies.
t2l
Chapter4
114-B) was noted in one sample (Poolowanna-3, 2560 m). It is tentatively identified as
carbargillite. Other examples of carbonate minerals seem to be very rare, with only one
questionable occurrence shown in Plate llE-F. It displays possible cleavage and is
characterised by a brownish grey colour in white light with a strong yellow fluorescence in
blue light irradiation.
The maceral types of this group are largely related to certain parts of higher plants. For
example, much of the telovitrinite in Euro-American Carboniferous coals originates from the
barks of the Lycophyta and the wood of Gymnosperms and Cycadophytes (Stach et aI.,
1982). Telovitrinites which are bounded by cutinite are derived from leaves or green twigs.
Desmocollinite is derived mainly from the cellulose of soft tissues; and the cellulose and
liginin of disintegrated cell fragments from woody tissues.
The depositional conditions favourable to the preservation of the vitrinite group are also
those which favour the transformation of lignin and cellulose into humic acid. Aerobic
microbiological activity in a weak acid medium is conducive to such transformations (Stach
et aI., 1982). Further, conversion of vegetable matter to vitrinite group macerals continues
below the ground water table. Under conditions of total absence of atmospheric or dissolved
oxygen the vegetable matter is subject to further physical and chemical degradation by
r22
Chapter 4
anaerobic bacteria. Under these anoxic conditions, the precursors of vitrinite group macerals
(i.e. humotelinite, humodetrinite and humocollinite) are subject to compaction and
The relative abundance of vitrinite macerals in a sample is therefore a measure of the relative
input of organic matter from vascular plants under first oxic then anoxic conditions. As
shown in Table 4.1, the Poolowanna Trough depocentre of the Poolowanna Formation
appears to have been the site of a higher vascular plant influx and more anoxic conditions
than was the case in the Patchawarra Trough. Several coals from Tantanna (Tantanna-8,
1823 m), Sturt (Sturt-4, 1881 m) and Sturt East (Sturt East-4, 1850 m) have similar maceral
group abundances and, by inference, depositional settings to those of the Poolowanna
Trough carbonaceous shales.
As already mentioned the presence of inertinite will not necessarily mean oxic conditions at
the site of burial. The inertinite may have been brought in from somewhere else. Different
types of inertinite may sometimes indicate the fusinitisation pathway taken by its precursors.
For instance, most of fusinites consist of fossil charcoal (the most cofirmon type), resulting
from incomplete combustion. Semifusinite is a product of either aerobic biodegradation
during humif,rcation, or oxidation and incomplete combustion of partially humified cell
tissue. However, an accurate identihcation of the various modes of origin is difficult in most
cases because they commonly overlap and rarely proceed in isolation. Important indicators
include the degree of cell preservation which is a measure of the relative effects of
humification and associated biodegradation; and reflectance, which increases with the
amount of oxidation and/or extent of combustion that the maceral has been subjected to.
Wood-derived inertinite typically displays better preserved plant cells than leaf-derived
semifusinite.
t23
Chapter 4
Inertodetrinite consists of remnants of plant tissue, mainly in the form of cell fragments of
fusinite and semifusinite. Sometimes it may result from charcoal formed from burning
vegetation (forest fires), which is then dispersed by wind or water as relatively small
fragments.
As shown in Table 4.1 , the abundance of inertinite in the Patchawarra Trough is greater than
that in the Poolowanna Trough, indicating different degrees of oxic conditions in the two
depocentres.
'While the Poolowanna Trough samples contain of
more semifusinite and
inertodetrinite than fusinite, the Patchawarra Trough samples consistently display more
fusinite than other inertinite macerals. On the basis this observation suboxic aquatic
conditions may be inferred for the Poolowanna Trough whereas the Patchawarra depocentre
appears to have been subject to dry episodes leading to prolonged exposure oxic conditions.
However, the predominance of inertodetrinite in the Tantanna area is an indication of a major
contribution from herbaceous plants.
The inferred oxic conditions are confirmed by high inertinite to vitrinite ratios (W: Table
4.3). It is suggested that W values less than unity indicate the prevalence of anoxic
conditions whereas values greater than unity may indicate suboxic to oxic conditions. Most
of the Poolowanna Field samples have values less than 1 whereas the Patchawarra Trough
in depositional
samples have values greater than 1, verifying the previously noted contrast
conditions for the Poolowanna Formation between the two areas. A couple of samples from
the Sturt East field have anomalously high W ratios (9-13) an indication of strongly oxic
conditions and hence the possible influx of allochthonous inertinite.
As a point of interest, the duroclarite facies has W values of less than unity, whereas the
clarodurite facies has values >1. The vitrinertoliptite facies has mixed values but the majority
are >1, thus showing a strong association between liptinite and inertinite.
Among the liptinite macerals, sporinites are nearly ubiquitous in Mesozoic sediments. Their
124
Chapter 4
occurrence is associated with the existence of the spore-bearing plants (pteridophytes) which
reproduce either heterosporously (e.g. lycopods) or homosporously (e.g. pteropods)
(Collinson and Scott, 1987). Both require a moist environment and free water in order to
fertilise the spores. Pollens on the other hand are produced by seed-bearing plants
(gymnosperms). These plants may survive in wetlands by having shallow roots, or in dry
ground by having deep roots. Therefore, the high abundance of sporinite, particularly in the
Patchawarra Trough samples signihes the prevalence of wetland environments for extended
periods. This situation is well illustrated by Plate 9C-D. In this figure, the fusinite layer
represents dry conditions which alternated with wet conditions represented by two layers of
sporinite. Telalginite is also present and this supports the existence of an aqueous
depositional environment.
The morphological features of sporinite also have implications for the kinds of plant and the
type of the environment in which they grew. For instance, thin-walled tenuispores
(particularly in the form of lycospores) are characteristic of wet, arborescent lycopod
swamps (Smith and Butterwofih, 1967). High abundances of lycospores are reported in
coals with high vitrinite/inertinite ratios in close proximity to palaeochannels (Harvey and
Dillon, 1985). Phillips et aI. (1985) noted that high vitrinite/inertinite ratios are consistent
with maximum preservation of biomass in continuously wet terrains where water covers the
ground for long periods. A similar observation may be made for the Poolowanna Trough
where tenuisporinite is more abundant than crassisporinite. The vitrinite to inertinite ratio
was generally greater than I for the Poolowanna organic facies in this area (Table 4.3).
Further information on the palaeoclimate in which fossil plants grew is provided by their
cutinite. The plants of wet environments are characterised by thin cuticles whereas those
growing in comparatively dry climates (i.e. xenophytic plants) possess relatively thick
cuticles. The cutinite in the Poolowanna Formation from both troughs has a thickness
between 1.6 and 9.6 pm. This thickness range indicates growth conditions that extend from
subaqueous to dry.
The contribution of organic matter from vascular plants is further demonstrated by the
presence of resinite, of which terpenoid resin acids are major precursors (Alexander et al.,
1988). Cunningham et al. (1983) concluded that fossil resinite had been formed by
125
Chapter 4
photolytic polymerisation of resin acids (e.g. agathic acid) after their exudation from the host
tree. The form of resinite encountered in Poolowanna Trough appears to be different from
that in the Patchawarra Trough, thus suggesting that different types of vascular plants grew
in these two areas during the early Jurassic.
Freshwater lacustrine or lagoonal settings are indicated by the presence of structured alginite,
referred to as telalginite. The telalginite that occurs in the Poolowanna Formation is of the
Pilatype which is related to Botryococcus braunli, an extant lacustrine colonial alga (Robert,
1988). It has been found floating as jelly-like masses on the surface of stagnant water near
Salt Creek in South Australia where it is referred to as coorongite (Diessel, 1992).
The Poolowanna Formation in the Patchawarra Trough appears to contain a relatively high
amount of telalginite, particularly in the Sturt East area. Occasionally, it occurs together with
inertinite, indicating that the inertinite is allochthonous. This telalginite is hardly ever seen in
samples from the Poolowanna Trough. Lamalginite, which occurs as thin anastomosing
lamellae formed by algal mats is noted in some samples (Plate 2A-B). This also indicates an
aqueous, probably lacustrine environment.
4.3.4 Microlithotypes
The microlithotypes of coal and dispersed organic matter (DOM) have been attributed to
different environmental settings by various authors (e.g. Hacquebard et al., 1967;
Teichmüller, 1982; Diessel, 1992). Among other microlithotypes, vitrite is generally
considered to be derived from woody tissues like stems, branches and roots and is indicative
of a forest swamp environment. Clarite is usually characterised according to its content of
liptinite. Liptinite-poor clarite originates from strongly decomposed plant litter (wood, bark,
etc.) in a forest moor environment; whilst liptinite-rich clarites originate from reeds and non-
arborescent vegetation which decompose readily, resulting in concentration of the more
resistantliptinite. Durite is usually characterised by its content of sporinite. Thus, sporinite-
rich durites are generally considered to be subaquatic ooze deposits (e.g. Poolowanna
Formation samples at 1841m and 1844 m in Sturt East-l), whereas spore-poor durites are
considered to represent deposition above the water table where extensive oxidation (and bush
fires) have occurred. Only two samples from the analysed section of the Poolowanna
Formation show a bimaceral association characteristic of the latter setting.
As discussed earlier, almost all the Poolowanna Formation samples can be characterised in
terms of diagnostic trimaceral assemblage (Table 4.3).
126
Table 4.3 Depositional environments of Poolowanna Forrnation source rock facies
tl
1796
I
Desmocollinite p nd 2.44
tl il il
181 1 p nd 3.52 Herbaceous wet swamps
1865
il il
I.2 nd 1.15
il
r87l Telocollinite 1.8 1.78 2.20
t829
il I
p 2.04 1.1 I
I
1856
t?
p 8.70 9.00
Sturt-3 lE59 nd 5.36
Sturt East-l 1841
il
3.2 r.45 9.40 Lacustine with
occasional
Sturt East-l 1844
il tl
5.4 0.49 r0.67 drying episodes
Sturt Ea.st-2 1829 Telocolliniæ 1.8 0.90 2.5r (sub aquatic)
Sturt East-3 1850
tl rl
Inertodetrinite 0.5 1.00 0.80
Sturt East-4 1838
il
Desmocollinite Fusinite 2.7 0.66 2.56
il ll
(duroclarite) Poolowanna-1 2423 Telocolliniæ 5.2 6.50 0.13
Poolowanna-1 2438
il
TeUDescollinite 0.6 o.24 o.77
I
Poolowanna-1 2469 Desmocolliniæ Inertodefrinite 0.7 2.42 0.82 Limnotelrnatic
Poolowanna-1 2545 Telocollinite Semifusinite p nd o.97
Table 4.3 (continued)
itt Maceral AI Pvrite Inertinite
facies (m) association Viüinite Inertinite (vo) (Vo) Vitinite
Düro-resinoclarite p 0.69 0.93
(duroclarite) Poolowanna-3 255r Duro-cutinoclarite p 0.13 0.86
Poolowanna-3 2560 Duro-resinoclarite p p 0.97
Sturt-4 1881 Duro-resinoclarite p 0.11 0.81
Sturt East-4 1850 Duro-liptodetroclarite p 0.93 0.81
Tantanna-8 1823 Vitinertite Semifusinite p nd 0.16
p
(claroduriæ) Poolowanna-3 2423 Semifusinite p nd 1.16
Tantanna-Z 1799 Fusinite p p 2.36 Terrestrial
Tantanna-3 r807
I
Semifusinite p nd 2.13 (low watertable)
I rl il
Tantanna4 1807 p nd 2.19
Tantanna-4 1814
I I
Inertodetrinite I.4 nd 2.23
It
Telocollinite Fusinite 1 nd
ll I
Sturt-4 1 859 0.3 0.16 2.77
Sturt-3 I 862 Desmocollinite o.2 0.73 2.22
il
Sturt-6 I 847 Telocolliniæ 0.6 0.37 1.51
rËffiiööäüütö" nd nd 1.47
Sturt-8 t862 I n il
pnd 2.21
Sturt East-l 1835
I
p O.27 2.47 Terrestrial
Sturt East-3 1865
il
Tel/Descollinite
I
p 0.57 12.88 (low water table)
Sturt East-4 r862 Desmocollinite
I
pnd 2.08
Chapter 4
The clarodurite association generally accumulates in areas of low water table and thus under
conditions of increased oxidation. The high abundance of fusinite in the clarite matrix
indicates the terrestrial zone which is affected by periods of dryness. This facies is
characterised by both arborescent vegetation and a low ground water level. Such an
environment may be similar to the one in which some of Poolowanna Formation source rock
facies, particularly those from the Sturt and Sturt East Fields, were deposited. Carbargillite
and other dispersed organic matter typically associated with the clarodurite facies occurs
beyond the fringe of rooted vegetation and is indicative of lakes and ponds which collect
detritus blown or washed in from other parts of the swamp. Indigenous algal-derived
microlithotypes, such as algite, algoclarite, vitrinertoalgite and duroalgoclarite are also
associated with clarodurite. Allochthonous inertodetrinite forms the bulk of the inertinite
component. Lacustrine conditions are indicated for this facies. In the Patchawarra Trough
this facies is clearly developed in the Sturt and Sturt East Fields, whereas in the Tantanna
Field there is less contribution from algal organic matter.
In marine or brackish-water environments, pyrite forms from the reaction of sulphide with
129
Chapter 4
iron, either directly or via a monosulphide precursor (Goldhaber and Kaplan, 1974). Iron is
supplied to the sediments in detrital minerals and as iron oxide coating on grains, mainly in
the fenic form. In order for the bacterial reduction of sulphate to sulphide to take place,
anoxic conditions must prevail since the sulphate-reducing bacteria are obligate anaerobes. In
addition, organic matter is also a prerequisite as it is the material on which the bacterial
heterotrophs feed (Harrison and Thode, 1958; Demaison and Moore, 1980).
The reactions involved are summarised as follows (Fisher and Hudson, 1987):
1. SOoz- +2CH2O 2HCO3-+ H2S @acterial sulphate reduction; the organic matter
->
is represented by carbohydrate)
2. 3H2S + 2 FeO'OH 2FeS + So + 4HzO (bacterially-produced sulphide reacting
->
with hydrated iron oxide, goethite, to form an iron monosulphide, mackinawite),
3 . FeS + So FeS2 (conversion of mackinawite to pyrite)
->
The last reaction may proceed via an intermediate greigite (Fe3S4) phase from which the
characteristic framboidal texture results (Sweeney and Kaplan, 1973).
The formation of greigite from mackinawite is an oxidation reaction, in which every fourth
sulphur oxidised to So' This suggests that the formation of framboids requires conditions
which are not totally reducing (Hudson, 1982; Fisher and Hudson, 1987). In fact, pyrite
generally forms within reducing microenvironments in oxic sediments, for example, disused
burrows, faecal pellets and enclosed voids such as foram tests or ammonite chambers
(Kaplan et aL.,1963).
Based on reaction kinetics, FeS2 can only form in peats and organic muds by bacterial
activity because there is insufficient energy for a purely chemical reduction of sulphate to
disulphides. Therefore, as a prerequisite for the formation of FeS2 in this environment, there
needs be a supply of both sulphur and iron. Sulphur originates from protein (largely
bacterial) or is carried in as sulphate ions by streams and"/or sea water. On the other hand iron
is present wherever silicate minerals weather, or where ground water carries ferrous or ferric
ions in solution (Stach, 1982).
An adequate supply of sulphate is a feature of coastal flood plains where major inundations
130
Chapter 4
could and be linked to influxes of brackish water, banked-up by long-tides and storm waves.
Syngenetic pyrite is occasionally found in the shaly coal facies of the Poolowanna Trough
(Plate lC-D) where it can also be related to a concomitant increase in pH in the tide-affected
marginal portions of the swamp (Anderson, 1976, as cited by Diessel, 1992). Similar but
dispersed syngenetic pyrite is observed in the fluvio-lacustrine, silty shale facies (Plates 8C-
D and 9E-F). The absence or rare abundance of pyrite in the Tantanna area is possibly due to
scarcity of sulphate and"/or highly oxic conditions in which sulphate-reducing bacteria could
not thrive.
The duroclarite facies has relatively moderate abundances of pyrite, with values ranging from
0.1 to 6.5Vo. Once again, this facies in the Tantanna Field appears to be almost devoid of
pyrite. The concentration of pyrite here is similar to that of the vitrinertoliptite facies; except
that the syngenetic pyrite appears to be framboidal (Plate 1lC-D), indicating the lowering of
dissolved 02 to anoxic or suboxic levels. The lower pyrite concentrations may also indicate
the preference of sulphate-reducing bacteria for a particular type of organic substrate.
The clarodurite facies is shown to have the lowest pyrite contents ranging from nil to O.6Vo.
As in the other organic facies, the Tantanna samples contain rare or no pyrite. The lack or
absenceof pyrite in this facies may be attributed to the persistence of oxic conditions. This
conclusion is strongly supported by inertinite/vitrinite ratios >1.0. The presence of alginite in
this facies, indicates perennial aqueous conditions which probably lacked a sulphate source
(e.9. a raised bog). Such settings are reported to suppress bacterial activity (Diessel, 1992).
Alternatively, as argued above, inertinite may be an unsuitable organic substrate for sulphate
reducers.
131
Chapter 4
Table 4.4 Source rock maturation parameters of the Poolowanna Formation based on
vitrinite reflectance
132
NT QLD t a
I z--\
\
--\ --.3r' a
I l/
.J-.L. _.r._. l-.---- ¿-.-
a
SA
tl lfI \
tl POOLOWANNA
ll t, i
It-'rä o' I
I
27.00'
\
STURT/ STURT EAST o
TANTANNA
\ 0.60
-1.
0 60 \
\1
km ---
-t'
29000'
Measured 7o Ro Infened 7o Ro
Figure 4.3 Vitrinite reflectance (7o Ro) at top Poolowanna Formation level
CONDENSATE
o RESINITE
RICH otL s
I
É.
N
GAS CH¿
Ø DENSATE
lrJ LIPTINITE
É. LIGHT
É.
LrJ
VITRINITE GAS
tt
I
¡ Poolowanna Trough
Patchawarra Trough
Figure 4.4 Relationship between organic matter type, expected hydrocarbon product and
thermal maturity for the Poolowanna Formation (based on diagram from from Powell and
Snowdown, 1983)
Chapter 4
The Poolowanna Trough samples have values ranging ftom0.7l%oRs (0.75Vo R1¡¡) at 2496
m in Poolowanna-2 to O.85Vo Rs (0.907o R¡¡) at 2560 m in Poolowanna-3, and are thus
indicated to be within the oil window. A discrepancy is noted in the trend of reflectance with
increasing depth whereby some samples at relatively shallow depth have a higher reflectance
than those below. Coincidentally, those samples showing an unexpectedly low reflectance
are those with relatively high contents of resinite (e.g. at 2499-2496 m and 2524 m n
Poolowanna- 1 and 2542 min Poolowanna-2).
In the Patchawarra Trough, the Poolowanna Formation of the Tantanna Field has values
ranging from}.SSVo Ro (0.61% R¡¡) at 1823 m in Tantanna-8 to O.75 Vo R6 (0.79Vo R¡¡) at
1807 m in Tantanna-4,placing it at early mature to mature stage (or onset) of oil generation.
Here again, there is almost no consistent pattern of reflectance variation with depth and this
is likely to be due to the diversity of organic facies. The formation in the Sturt Field is less
mature than in the Tantanna field, with values ranging from 0.57Vo Ro (0.607o Romax) at
1881 m in Sturt-4 to O.66Vo R6 (0.73Vo R¡¡) at 1865 m in Sturt-6. The onset of oil
generation has been also reached here although again the vitrinite reflectance data behaves
inconsistently with depth. In the Sturt East Field, vitrinite reflectance varies from 0.57o R6
(O.53Vo R¡¡) at 1850 m in Sturt East-3 to O.69Vo Rs (OJ3Vo R¡¡¡) at 1829 m in Sturt East-2.
The same thermal maturation level as in the neighbouring field is indicated. The regional
variation of thermal maturity at the top of Poolowanna Formation is illustrated in Figure 4.3.
As shown in Table 4.4,the /a values for the Poolowanna Trough range from 0.75 to 0.80,
which is equivalent to high-volatile bituminous to medium-volatile bituminous rank. The
Patchawarra Trough samples have somewhat lower values ranging from 0.65 to 0.16 which
are equivalent to high-volatile bituminous rank. Like the vitrinite reflectance data, these
aromaticity values show that the Poolowanna Formation in the Poolowanna Trough is within
the oil window, whereas in the Patchawarra Trough it is at the onset of oil generation.
134
Chapter 4
In the Poolowanna Trough the average maceral abundances show that vitrinite < liptinite >
inertinite, whereas in the Patchawarra Trough their average relative abundance is vitrinite <
liptinite < inertinite It can therefore be suggested that the Poolowanna Trough source rocks
contain Type IIIIII kerogen whilst those in the Patchawarra Trough contain mostly Type
III/II kerogen. The characteristic Type I macerals in the Poolowanna Trough are resinite
and./or lamalginite, whilst in the Patchawarra Trough itis Botryococcus-like telalginite.
Relatively high resinite contents in the Poolowanna Trough impart the potential for early
generation of light naphthenic oil. Because of the considerable abundance of other liptinites,
the Poolowanna Formation here is probably at the peak of waxy oil generation. The high
content of vitrinite suggests that the early stage of gas generation may have been reached. It
is therefore concluded that carbonaceous shales in the Poolowanna Trough are effective
sources for waxy oil and minor gas. The maturation level of the Poolowanna Formation in
the Patchawarra Trough indicates that it is at the early stage of both oil and gas generation,
except where there is a significant resinite content. However, as shown from the maturity
map (Fig. 4.3), this formation in the deeper parts of the trough may be at peak generation for
waxy paraffinic oil or perhaps even within the light paraffinic oil window.
Plate 1
135
IìI,A'I-E
-¿ -'f
I
I 'J\'
',FF':
:/
-- è
- ---
T! f-,r-'1h**
-+<a:- l'r
él 1-t
l=.Þ
I a
tÊÞi-
-å 'I*"*
a
a9t
e
¡ --;
a
¡t
'F44'f'- ;T €
I
t
a.
? '.L'
a
1, t' ?'
"#r.
l'.
f.
Ë
G H
PLATE 1
A: Well: Poolowanna-l
Depth: 2417 m
Microlithotype: Clarite
Rs: 0.75Vo
'Well:
C: Poolowanna-1
Depth: 2499 m
Microlithotype: Monomacerite (DOM in silty shales)
R.s O.797o
E V/ell: Poolowanna-2
Depth: 2542 m
Microlithotype: Trimacerite
Rs: 0.74Vo
Field of view: 136 pm xl97 ¡tm
Description: Desmocollinite, coarse inertodetrinite and liptodetrinite
F: As in E, blue light excitation; fluorescing vitrinite; lamalginite and bituminite.
G V/ell: Poolowanna-l
Depth: 2499 m
Microlithotype: Trimacerite
Rs; 0.797o
Field of view: 136 pm x 797 ¡tm
Description: Resinite in cell cavities (telocollinite); semifusinite band and pyrite
nodules.
As in G, blue light excitation; dull orange resinite fluorescence.
136
j{t
PLATE 2
A B
L
É
c D
E F
G H
PLATE 2
A: V/ell: Poolowanna-2
Depth: 2542 m
Microlithotype: Trimacerite
P.s: O.74Vo
Field of view: 136 pm x 197 pm
Description: Layers of inertinite (fusinite, semifusinite and inertodetrinite); vitrinite
(telocollinite); and a round brownish yellow haze (bituminite ?)
B: As in A, blue light excitation; alginite (lamalginite); fluorescing vitrinite; round
yellowish fluorescing ?bituminite.
C V/ell: Poolowanna-l
Depth: 2417 m
Microlithotype: Trimacerite
R.s 0.75Vo
Field of view: 136 pm x 197 pm
Description: Bands of vitrinite (coarse detrovitrinite); inertinite (semifusinite) and
liptinite (cutinite, ca9.6 ¡rm thick).
D As in C, blue light excitation; thick cutinite, resinite and/or fluorinite stringers.
E V/ell: Poolowanna-2
Depth: 2496m
Microlithotype: Bimacerite (Clarite)
P.s: 0.7lVo
Field of view: 136 pm xl97 ¡tm
Description: Sporangia (sporinite) separated by vitrinite bands; pyrite concretion.
F: As in E, blue light excitation; brownish yellow fluorescence of sporangia.
G Well: Poolowanna-3
Depth: 2560m
Microlithotype: Bimacerite (Clarite)
Re: 0.857o
Field of view: 136 pm x 197 pm
Description: Vitrinite (desmocollinite); liptinite (different forms of macro-sporinite)
H As in G; blue light excitation; sporinite exhibiting a near perfect preservation of the
entire spore and its inner contents; fluorescing vitrinite and alginite.
138
H Ð
- l& r:1
ta t.
q tl.
I t
v
I I
¿\
\.
,
,?
i
.t
I f
x+qt*ìg t
Ò
il
{
t
U
_-
Þ
J 1
4 .t Ç
ó ì
U
.$ )' F*
I
'rt-
i
I
-\t-Þi à. fl
â
t,
;Ë¿$i+''
LI v
, tÀ.'
t- iI.L\"'Id
PLATE 3
A: Well: Tantanna-3
Depth: 1807 m
Microlithotype: Monomacerite
P.s.. 0.62Vo
Field of view: 136 pm x 197 pm
Description: Dispersed organic matter (DOM) liptinite (bituminite, ?sporinite appears
translucent brown).
B: As in A, blue light excitation; bituminite, yellowish groundmass; ?sporinite, yellow
to orange fluorescence.
C: V/ell: Tantanna-l
Depth: 1795 m
Microlithotype: Monomacerite
F.s 0.67Vo
Field of view: 136 pm x 197 pm
Description: DOM; bituminite, brownish translucent liptinite
D As in C, blue light excitation; yellow ?sporinite; brownish orange bituminite and
liptodetrinite.
E Well: Tantanna-8
Deprh: 1823 m
Microlithotype: Monomacerite
Rs: 0.58%
Field of view: 136 pm x 197 ¡rm
Description: Sporangium, reddish brown with partial intemal reflection;
desmocollinite.
F As in E, blue light excitation; yellow ornamented sporangium
G: Well: Tantanna-2
Depth: 1799
Microlithotype: Trimacerite
R.s: 0.67Vo
Field of view: 136 pm x 197 pm
Descrþtion : Fine-grained vitrodetrinite; inertodetrinite and alginite
H As in G, blue light excitation; miosporinite, telalginite and bituminite
140
PLA'IE 4
t
,l
1..,t
A B
c D
t
+rt
t.
f;
t
a.
.û'
tj
ì ¡
I
li t, ¡
'.\ +
E F
ilT
G H
PLATE 4
A: Well: Tantanna-2
Depth: 1799 m
Microlithotype: Trimacerite
Rs O.67Vo
C Well: Tantanna-4
Depth: 1814 m
Microlithotype: Bimacerite (durite)
Fts; O.66Vo
E V/ell: Tantanna-3
Deprh: 1807 m
Microlithotype: monomacerite
R.s: 0.62Vo
Field of view: 136 pm xl97 ¡tm
Description: Telalginite (brownish with intemal reflections) in association with
minor inertodetrinite dispersed in siltshale
F: As in E, blue light excitation; orange fluorescence of telalginite.
G Well: Tantanna-2
Depth: 1799 m
Microlithotype: Bimacerite
R.s: 0.67Vo
Field of view: 136 pm x 197 pm
Descrþtion: Inertinite (deformed fusinite and semifusinite) dispersed in shale
together with; liptinite (telalginite, dark brown bodies; and bituminite, brownish
groundmass)
As in G, blue light excitation; telalginite (orange fluorescence); bituminite (dull
brownish yellow fluorescent groundmass) ; sporinites and liptodetrinite.
142
[,L,4. ['E 5
A B
r* D
-ir,
'üc'
E F
G H
PLATE 5
A: Well: Sturt-l
Depth: 1865 m
Microlithotype: Trimacerite
Pto 0.597o
Field of view: 136 pm x 197 pm
Description: A layer of fluorinite (dark elongated lens) in desmocollinite,
liptodetrinite, inertodetrinite and fu sinite.
B: As in A, blue light excitation; strong yellow fluorescence of fluorinite; fluorescing
liptodetrinite, miosporinite (tenuisporinite) ; and desmocollinite.
C: V/ell: Sturt-8
Depth: 1862m
Microlithotype: Trimacerite
Rs: 0.59Vo
Field of view: 136 pm x 197 pm
Description: Resinite bodies (dark) in association with desmocollinite, semifusinite,
micrinite and liptodetrinite.
D As in C, blue light excitation; strong yellow fluorescence of resinite, miosporinites
and stringers of exsudatinite and fluorinite (strong orange-yellow fluorescence)
E Well: Sturt-6
Depth: 1847
Microlithotype: Trimacerite
R6: not determined
Field of view: 136 pm x 197 pm
Description: Resinite and/ or fluorinite (dark) globules in telocollinite interlayered
with fusinite (displaying bogen structure)
F: As in E, blue light excitation; strong green yellow fluorescence of fluorinite
G V/ell: Tantanna4
Depth: 1807 m
Microlithotype: Trimacerite
P.s: 0.75Vo
Field of view: 136 pm x 197 pm
Description: Layers of telocollinite, semifusinite and dark brown telalginite bodies,
in association with inertodetrinite.
As in G, blue light excitation; showing fluorescence of cutinite and liptodetrinite.
t44
PLATE 6
t' t
A B
C D
'tæ' i*fi {"{ .
E F
lfrr
G H
PLATE 6
A: Well: Sturt-3
Depth: 1859 m
Microlithotype: Monomacerite
Rs: not determined
Field of view: 136 Pm x 197 Pm
Description: Spherical brownish fluorinite, translucent brown cutinite and sparse
vitrodetrinite in siltstone.
B: As in A, blue light excitation; strong green fluorescence of fluorinite globuies;
cutinite stringer showing yellowish orange fluorescence'
C Well: Sturt-4
Depth: 1859 m
Microlithotype: Trimacerite
P.s: 0.64Vo
Field of view: 136 ¡rm x 197 Pm
Description: Alternating bands of vitrinite (desmocoliinite and telocollinite); inertinite
(semifusinite and fusinite) and liptinite (fluorinite - dark layer)
D As in C, blue light excitation; strong greenish-yellow fluorescence of fluorinite; weak
greenish-brown bituminite in desmocollinite ; and liptodetrinite specs.
E V/ell: Sturt-6
Depth: 1865 m
Microiithotype : Trimacerite
Re: 0.667o
Field of view: 136 Pm x 197 Pm
Description: DOM; coa.rse vitrodetrinite and inertodetrinite; translucent liptodetrinite
and light brown resinite in silty shale'
F As in E, blue light excitation; strong yellow fluorescing globules of fluorinite; light
brownish yellow cutinite; resinite body with zonal structures.
G Well: Sturt-4
Depth: 1881 m
Microlithotype: Bimacerite
P.6: 0.57o/o
Field of view: 136 Pm x 197 Pm
Description: Alternating bands of thin cutinite (l'6-3.2 pm thick) and
t46
If l.,A'l'll 7
A Ë
I , ttat?'
C D
E F
Lã H
PLATE 7
A: V/ell: Sturt-l
Depth: 1865 m
Microlithotype: Monomacerite (liptite)
R6: 0.597o
Field of view: 136 pm x 197 pm
Description: Cutinite-rich liptite (leaf cuticles closely compacted) and sparse
inertodetrinite.
B: As in A, blue light excitation; yellow fluorescing cutinite band.
C Well: Sturt-l
Depth: 1865 m
Microlithotype: Bimacerite (durite)
R6: 0.597o
Field of view: 213 ¡tmx 308 pm
Description: Mega-sporinite, brownish spongy texture indicative of oxidation; in
association with fusinite.
D: As in C, blue light excitation; yellow fluorescing mega-sporinite and liptodetrinite.
E Well: Sturt-4
Depth: 1859 m
Microlithotype: Monomacerite
Rs: not determined
Field of view: 136 pm x 197 pm
Description: DOM; telalginite (reddish-brown intemal reflection) associated with
pyrite concretions in silty shale.
F: As in E, blue light excitation; strong orange fluorescence of. Pila colonies
G Well: Sturt-3
Depth: I862m
Microlithotype: Bimacerite (durite)
R.s: 0.64Vo
Field of view: 136 pm x 197 pm
Description: Telalginite in association with macrinite (ikely transformed from
resinite) and inertodetrinite in carbonaceous shales.
H As in G, blue light excitation; strong orange yellow fluorescence of PiIa colonies and
liptodetrinite.
148
I'I ,A-T'E 8
A B
tb#*
t d\t
!
+ "t
Vì.
U D
E F
G H
PLATE 8
A: V/ell: Sturt East-4
Depth: 1838 m
Microlithotype: Bimacerite (durite)
F.s 0.647o
Field of view: 136 pm x 197 pm
Description Botryococcøs-like telalginite (Pila) and sparse inertodetrinite.
B: As in A, blue light excitation; PiIa colonies (strong orange fluorescence).
r50
PI,ATtr 9
ì- (,+
._ P
,$. ,Ü
t
#,t r
J}
*1
2 tî+
A B
c D
E F
n
H
PLATE 9
A: Well: Sturt East-l
Depth: 1835 m
Microlithotype: Trimacerite
P.s: O.66Vo
Field of view: 136 pm x 197 pm
Description: DOM; brownish translucent liptinite associated with fine inertodetrinite,
vitrodetrinite and pyrite nodules in silty shale.
B: As in A, blue light excitation; abundant orange yellow fluorescing microsporinite
(crassisporinite) or miosporinite and bituminite.
C Well: SturtEast-l
Depth: 1841 m
Microlithotype: Monomacerite
Rs: 0.587o
Field of view: 136 ¡tm x 197 pm
Descrþtion: DOM; layers of translucent liptinite and fusinite (deformed cell
structures), overlain by sandstone layer.
D As in C, blue light excitation; liptodetrinite and miosporinite associated with
telalginite in bituminite groundmass.
t52
to orange yellow
153
PI-,ATE IO
I
I
ì¡"
I
\ .t
:' I ¡
I
Õ {.* b \
I irr
A B
:)
?
<¿i^ i.*
' "*
.*
¡. þ *, ,.¡È'
C D
H F
G H
PLATE 10
A: Well: Sturt East-3
Depth: 1865 m
Microlithotype: Bimacerite (durite)
Rs: 0.607o
Field of view: 136 pm x 197 pm
Description: Oblique to bedding section through a thick (ca 10 pm) cutinite, in
association with fusinite, micrinite, sparse vitrodetrinite and some pyrite nodules.
B: As in A, blue light excitation; orange to brown fluorescence of cutinite, showing
compressed cell structures and possibly orange fluorescence of fluorinite excretion
(in the middle).
liptinite.
D As in C, blue light excitation; orange yellow fluorescence of sporinite clusters in
band (?sporangia).
155
H: As in G, blue light excitation; greenish yellow fluorescing oil specs
156
H Ð
I =l
ü J
I V
..q ¡
f
t -/t- I
a
f s
t t, t
.ô:
I t
- d rlr ."1
-i- ùr' tr
,f* t !¡
-l t
t "¡
T,l-VTcl
ff
I I
PLATE 11
A: Well: Poolowanna-3
Depth: 256O m
Microlithotype: ?Carbargillite
R6: 0.857o
Field of view: 136 ¡rm x 197 pm
Description: Clay mineral (?illite) or amorphous organic matter, dark brown to
translucent, bounded by tellocollinite.
B: As in A, blue light excitation; massive fibrous body with yellowish fluorescence.
C: Well: Poolowanna-3
Depth: 2551
Microlithotype: Carbopyrite
Rs: not determined
Field of view: 136 pm x 197 pm
Description: Pyrite framboids (spherical concretions of fine crystallite); in
desmocollinite band.
D: As in C, blue light excitation; deformation orientation of sporinite (dull orange
fluorescence), indicates syngenetic formation of pyrite.
E Well: Sturt-4
Depth; 1881m
Microtthotype:
Rs: not determined
Field of view: 136 pm x 197 pm
Description: ?Carbonate mineral bounded by quartz grain.
F As in E, blue light excitation.
158
Chapter 5
Chapter 5
5.1 Introduction
Source rocks are defined as sedimentary rocks which contain sufficient organic matter of
suitable chemical composition to generate hydrocarbons. The term'source rock' is applied
irrespective of whether the organic matter is mature or not. The presence of kerogen is a
prerequisite for an active or potential source rock. The term 'potential source rock' is used to
describe the organic-rich rocks which have not yet generated significant amounts of
hydrocarbon due to immaturity. Active or effective source rocks are those which have
generated and expelled the hydrocarbons.
Petroleum source rocks are normally recognised by determining their total organic carbon
(TOC) and extractable organic matter (EOM) or bitumen contents. A second step in source
rock identification is the determination of the type of kerogen and the composition of the
solvent extractable hydrocarbons and non-hydrocarbons. The final step is the determination
of the evolutionary of the kerogen, a concept commonly referred to as source rock
stage
thermal maturation. In this case it has been determined by vitrinite reflectance (Section
3.3.1.2) and by chemical indicators based on bitumen and Rock-Eval pyrolysis parameters.
A combination of data derived from these methods allows the evaluation of the hydrocarbon-
generative potential of the source rock and prediction of whether it is gas or oil-prone.
160
Table 5.1 Rock-Eval pyrolysis data and expected hydrocarbon product for the Poolowanna Formation
No. (m) oC
mÐ( 2
mgH(Y mgHCY mgHC/
3
mgHCY Vo mgHC/ mgHC/ Product
rock rock rock rock g TOC TOC
POLI-2 Poolowanna-l 2423 4t 1.4 9.6 0.4 11.0 24.7 0.12 5.1 190 8 Gas
POL1-4"' Poolowanna-l 2M9 47 5.3 22.5 1.1 27.9 2t.2 0.19 rt.8 191 9
"
'
ÞöË3:î i"
"
Þööïö\üäää: j
"
2423 47 3.9 4t.2 1.2 45.1 33.E 0.09 18.3 225
POL3-13 Poolowanna-3 2505 MI r.6 12.4 0.4 14.0 28.7 0.12 5.4 227 8
POL1-1 Poolowanna-l 2417 M7 3.7 38.3 1.1 41.9 33.9 0.09 16.3 235 7
POL1-3 Poolowanna-l 2438 47 7.4 46.3 1.5 53.7 30.3 0.14 19.5 237 8
POL1-6 Poolowanna-l 2545 448 4.6 38.5 1.1 43.t 35.9 0.11 15.4 250 7 Gas and Oil
POL1-7 Poolowanna-]. 2557 M9 3.5 29.3 0.8 32.8 38.1 0.11 tt.2 262 7
POL2-10 Poolowanna-2 2542 451 4.7 52.8 1.0 57.5 55.0 0.08 19.8 266 5
POL3-15 Poolowanna-3 2560 450 4.4 44.3 0.9 48.7 50.3 0.09 16.3 272 5
POL3-14 Poolowanna-3 255r 451tr.2 96.6 2.0 107.8 48.1 0.10 33.5 288 6
1-5 Poolowanna-1
-249q 442 6.4 65.2 1.9 7r.6 34.1 0.09 20.7 315 9
POL2-9 Poolowanna-2 2524 452 8.7 106.4 1.5 115.1 71.0 0.08 30.7 347 5 oil
POL2-8 Poolowanna-2 2496 449 13.5 t46.3 2.5 159.8 59.2 0.08 42.r 348 6
+
No. (m) oC mgH(Y mgH(}
3
mgHCY 7o mgHC/ mgHC/ Product
rock rock rock grock TOC TOC
STU2-28 Sturt-2 1829 436 0.1 3.5 0.8 3.6 4.5 0.04 2.1 165 36
STU3-33 Sturt-3 1862 437 7.6 60.8 2.8 68.4 22.0 0.11 36.t 168 8 Gas
STU4-30 Sturt-4 1859 435 2.1 37.5 1.5 39.6 25.7 0.05 2t.4 175 7
STU6-35 Sturt-6 1865 436 12.2 TIT.7 4.2 123.9 26.8 0.1 61.4 1,82 7
STU6-34 Sturt-6 t847 435 2.2 43.6 1.8 45.8 24.2 0.05 22.r r97 8
STUS-36 Shlrt-8 t862 434 2.8 66.3 3.5 69.r 19.0 0.04 33.4 t99 10
STUl-27 Sturt-1 1871 436 0.8 28.0 1.0 28.8 27.2 0.03 tt.1 239 9 Gas and Oil
STU4-31 Sturt-4 1881 435 15.6 t32.8 4.0 148.4 32.9 0.1 51.9 256 8
STB1-45 Sturt East-4 1862 437 5.1 t15.4 4.9 120.6 23. 80 04 68.1 170 7 Gas
STE1-37 Sturt East-l 1835 434 2.2 47.4 2.1 49.6 22. 30 04 249 190 9
"'
S'iË2 -4ö
"""
Stää'Ë¿îË2"
"""""" "'
î 82t
""
436 1.5 3t.4 t.7 32.9 18.6 0.04 15.6 201 î
STE3-41 Sturt East-3 1850 433 0.6 18.0 0.8 18.6 22.8 0.03 7.9 229 10
STEl-39 Stuft East-l 1844 432 0.7 12.4 1.1 13.1 I 1.1 0.05 5.3 234 2I
STE/-44 Sturt East-4 1850 435 14.8 137.9 4.0 t52.7 34.5 0.1 56.4 245 7 Gas and Oil
STEl-43 Stuft East-4 1838 437 t.3 36.4 1.3 37.7 29.r 0.03 14.2 256 9
STEl-38 Sturt East-l 1841 438 0.9 22.0 0.9 22.9 23.4 0.04 7.8 283 12
A cross plot of HI versus OI for the Poolowanna Formation source rocks is shown in Figure
5.2a. It appears from this diagram that samples from the Poolowanna Trough and some
from the Patchawarra Trough lie to the left of the Type I evolution pathway. These samples
are characterisedby very low oxygen index values which seem to be an analytical artefact.
The only clearly indicated categorisation is that between kerogen Types II and III for a
couple of samples from the Patchawarra Trough. Obviously this plot does not resolve the
kerogentypes adequately.Insteadaplotof HI versus T,nu* (Espitalié et al., 1984) was used
for kerogen classification (Figure 5.2b). It is clear from this figure that most of the
Poolowanna Trough source rock facies contain Type II/III kerogen, although an apparent
Type I kerogen may be presented in samples which have a high resinite content (Table 4.1).
The Patchawarra Trough samples plot mostly in the Type III/tr zone.
163
Vitrinite
TOC Range
a
. 1-10%
r 11 -3O%
I 30 -49To
r 50 - 10O%
a
T
TT
I T ^
a
ra ¿
T
a a T
ao a a
lI a
a
¡ a o
Liptinite lnertinite
100.00
Vitriniûe
* Liptinite
tr
.9 75.00 Inertinite
f
¡¡
tr
8 so.oo
o
.l
(É
co es.oo
0.00
0.00 20.00 40.00 60.00 80.00
TOCo/o
Liptinite
The cross plot of liptinite versus HI (Fig. 5.3a) reveals a weak positive correlation between
these two variables. The Poolowanna Trough samples show a positive linear relationship
with a regression coefficient (r) of 0.52. The Tantanna and Sturt East data have regression
coefficients of 0.48 and0.46, respectivel], whereas a negative correlation is observed in the
Sturt data (r = -0.23). These variations indicate different organic matter inputs to the
Poolowanna Formation in these fields. A negative correlation possibly shows that the
hydrogen in those samples is rarely contributed from liptinite or alternatively is released from
liptinite finally disseminated in other maceral groups (e.g. bituminite and resinite in vitrinite,
or fluorinite in inertinite which were not accounted for in point counting). Another possible
explanation for the lack of a strong positive correlation between liptinite and HI is that it is
due to oxidation effects or the presence of reworked liptinite.
Vitrinite
Again, there is only a weak positive correlation between vitrinite and HI (Fig. 5.3b). A
negative correlation is observed for data from the Poolowanna f,reld where the highest HI
values (>300 mg HC/g TOC) are recorded. This may imply that the vitrinite content in this
area does not significantly contribute to the measured HI values. The Tantanna data display a
weak positive correlation (r = 0.25) whereas in the Sturt data a somewhat stronger positive
correlation (r = 0.5) is observed, implying some contribution by vitrinite to the HI values.
The lack of a correlation in the Sturt East data is a possible indication of a mixed hydrogen
contribution from all the maceral groups.
Inertinite
A negative correlation between inertinite content and HI is shown for all the fields giving an
overall trend with a regression coefficient of -0.49 (Fig. 5.3c). This indicates that inertinite
makes no contribution to the measured HI values.
r65
1000 OIL PRONE
o Poolowanna
A Tantanna
800 o Sturt
tr Sturt East
o
o
F
600
ctt
o
ctt
E
400
200
o
ilt
0 GAS PRONE
0 50 100 150 200
900
olo
L'
o.-
800
.\S'
/
700 I
I Field
I O
x I
Poolowanna
HO 600 I A Tantanna
=o O Sturt
zo,
tr Sturt-East
UÞ 500
PË
og
400
.o
oô
300
À
o
200 ooo
\
100 \o I
\
0
380 400 420 440 460 480 500 520
TmaxoC
Figure 5.2b Kerogen classification based on hydrogen index and T-u*
Chapter 5
Saturated hydrocarbons as a function of the total EOM range ftom 7Vo in silty coals from
Tantanna and Sturt East to 35Vo in coal from Sturt. Across all the fields saturates comprise an
average of l5%o of the EOM. Aromatic hydrocarbons range ftom23Vo in coal from Sturt-6 to
37Vo in silty shale from Poolowanna-l. It is noted that where the liptinite (sporinite) content
is relatively high the aromatic content is relatively low. Poolowanna Trough extracts tend to
have larger aromatic fractions (average 34Vo) of EOM than do those from fields, in the
Patchawarra Trough. Almost all samples have aromatic to saturate ratios (A/S) greater than
unity, except one coal sample from Sturt-6 which has a value of 0.65. This sample appears
to be unusual as most coals have higher A/S values than do non-coal lithologies of the same
rank or maturity. Coincidently, another coal sample (Sturt-4, 1881 m) has a relatively low
A/S value (1.22) and thus is likely to be of higher maturity than indicated by its vitrinite
reflectance (0.57Vo Re), or contain migrated hydrocarbons rich in saturates. Liptinite-rich
silty shales from Sturt East-1 (1841 and 1844 m) and Tantanna-2 (1799 m) also have
relatively low A/S ratios (1.09-1.3).
The macromolecular structure of Type III kerogen comprises condensed polyaromatic units
'Welte,
and oxygenated functional groups, with minor aliphatic chains (Tissot and 1984).
Aromatic and naphthenic rings are common in Type II kerogen, whereas Type I kerogen
consist of cross-linked aliphatic chains and a few aromatic nuclei.
Therefore the high NSO contents (average = 53Vo), and generally high aromatic to saturate
ratios of the Poolowanna source rock extracts appear to be typical of mixed Type IIlIl
kerogen.
A comparison of hydrocarbon yield (i.e. the sum of saturate and aromatic fraction: Table
5.2) and maceral group abundance (Table 4.1) shows that where vitrinite content is relatively
low the hydrocarbon yield is also low (< 5000 ppm). Liptinite is the highest contributing
maceral in these samples. This situation might be explained by an oxidation event, whereby
resistant liptinites survived while the vitrinite and the contained hydrocarbons were readily
oxidised. In samples with intermediate hydrocarbon yields (5,000 to 10,000 ppm), there is
an increase in vitrinite content. However, inertinite abundance increases more noticeably than
both vitrinite and liptinite. It is well known that inertinite does not contain significant
amounts of extractable hydrocarbons. The observed increase in these samples is probably
due to disseminated liptinite (Chapter 4).
168
a) Liptinite c) lnertinite
Í=0.21 O o oo r = -0.49
o o
A
o o À
Å Á.
o
o
F e
o) ô o
o tr9" ot
Ê¡ + qtr
o o@
I Ä
(') u o
E s oo o o o æ
tr sd o o
=
1
A å.A A. AA
1 trô o
o
4
0 25 50 75 100 0 10 20 30 40 50 60 70
Liptinite (%) lnertinite (%)
b)Vitrinite
r = 0,19
oo
o
^å o
g Á,
o tro o o Poolowanna
o + o
ct) À Tantanna
o Re' o
I a8 otr oo o Sturt
o)
E 1 4a tr Sturt East
= 1
o
0 25 50 75 100
Vitrinite (%)
Figure 5.3 Influence of maceral groups on hydrogen index values. The regression line in
each panel is that for the whole sample population.
Table 5.2 Rock extract yield and composition data for the PoolowannaForrnation
Sample Depth EOM Saturates Aromatics NSO Saturates Aromatics NSO* msEOMlqgsat lqg$ry 4qSHC Arcmatic
No. (m) (ppm) (Vo) (Vo) (Vo) gTOC gTOC Saturate
* NSO polar compounds eluted from the column during liquid chromatography. The balance of the non-hydrocarbon fraction of the EOM (mainly
=
asphaltenes) was retained on the column
Chapter 5
Several samples ptot off trend in the early mature zone. The cause of this inconsistency is
either variation in kerogen type or contamination by migrated bitumen. For instance, the
coaly siltstone facies at 1811 m in Tarúanna-2 has an anomalously high hydrocarbon yield
while not yet fully mature. A possible explanation is either that it contains migrated
hydrocarbons or its recorded vitrinite reflectance is too low as a result of bitumen or resinite
dissemination in the vitrinite. The PI value (0.13) of this sample reasonably indicates early
172
AROMATIC HC
80 20
60 40 o Poolowanna
r Tantanna
o Sturt
40 60 o Sturt East
20 80
Figure 5.4 Gross cornposi..ion of Poolowanna Formation source rock extracts a.s
demonstrated by the relative abundance of their hydrocarbon fractions (saturate and
aromatic) and NSO compounds. The latter fraction does not include asphaltenes.
Chapter 5
The relationship between kerogen type, extractable hydrocarbon yield and maturity was
checked by plotting the hydrocarbon yield against hydrogen index (Fig. 5.6). Here, three
categories of positive correlation were identified based on maturity zones. ZoneI represents
the 0.50-0 .697o Ro maturity range. Two trends were observed in this zone. The first trend
shows a wide spread of HI values with little or no variation in hydrocarbon yield (ca.20 mg
IFrCIgTOC). This suggests that, regardless of the type (or quality) of the kerogen, it has not
acquired a level of maturity suitable for hydrocarbon generation. Most of the samples from
Sturt East and Sturt and a few from Tantanna, belong to this category. A second subordinate
trend is marked by an increase of HI with hydrocarbon yield. Such a trend may indicate that
at this relatively low maturity level, at least one kerogen type is able to generate
hydrocarbons. Most of the Tantanna samples plot within this trend which indicates that they
contain a kerogen type capable of hydrocarbon generation at early maturity. A similar but
stronger trend is noted in the 0.70-0.79Vo Re maturity zone. All but one of the samples in
this zone are from the Poolowanna Trough, reflecting hydrocarbon generation from its
resinite-rich organic matter.
The 51+ 52 values of the Poolowanna Formation in the Poolowanna Trough range from 11
mg HC/g in silty shale (Poolowanna-1, 2423 m) to 160 mg HClg in shaly coal
(Poolowanna-2,2496 m). Very good to excellent generative potential is indicated. In the
Patchawarra Trough, the values at Tantanna range from 9 mg HC/g in shale (Tantanna-1,
1795 m) to I54 mg HC/g in coal (Tantanna-8, 1823 m) and indicate good to excellent
hydrocarbon generative potential.
174
0.4
lmmature
Early
0.6 $oo mature
A ^
i¡l-/-t-Í.Í.Í t ¡Í¡/i/tt/¡
KrritttÍ!¡i/t./.t
òq
o o o Mature
É.
A
oo
Peak oil
0.8 P P,,,,,,,,
generation o
oo
Í,! / -/ / / / / ¡ / / ¡ ! / / / /,/ /,/ -/ ./ ./ -Í / -/ .t -t -Í .t .Í -Í -Í -Í -/ -Í -l J -Í -/ .f.l .rr ,l -/
Late mature
0 10 20 30 40 50 60
mgHC/gTOC
Field
o Poolowanna o Sturt
A Tantanna o Sturt East
350
I
300 I
A\
O
o
I
trl
t- I AI
o) I
tr ¿l
o
I
o)
E
250
I
I
I
#-i
ul
I I A\
= 200 I I
I I o
K
\
150
A,z
\ \.
-¿ -/
100
0 20 40 60 80
Hydrocarbon yield
mgHC/gTOC
o Poolowanna
0.50 - 0.69 %Ro
A Tantanna
0.70 - 0.79 %Ro
o Sturt
0.80 - 0.89 %Ro
tr Sturt East
A good to very good generative potential is demonstrated in the Sturt field where the 51+ 52
values range from 4 mg HC/g in siltstone (Sturt-2, 1823 m) to 148 mg HC/g in coal (Sturt-4
1881 m). At Sturt East the values range from 6 mg HC/g in siltstone (Sturt East-3, 1865 m)
to 153 mg HC/g in silty coal (Sturt East-4, 1850 m). It should be noted that in both troughs
the highest values were recorded in the coal facies.
Like the 51+ 52 parameter, the extractable hydrocarbon yield (ppm) has values that indicate
very good to excellent generative potential. The Poolowanna Trough samples have values in
the range 1972-15394 ppm whilst those of Patchawarra Trough have ranges of 1827-11226
ppm at Tantanna; 2429-13157 ppm at Sturt and 1018-11443 ppm at Sturt East. In both
troughs the highest hydrocarbon yields are those of the coal or shaly coal facies. The
outcome of this assessment is that the coal facies of the Poolowanna Formation is capable of
generating oil.
Similarly, a positive correlation is shown for vitrinite (Fig. 5.7b). In the Tantanna and Sturt
East Fields a strong contribution from vitrinite is demonstrated by positive correlations with
a regression coefficient of 0.88. An even stronger correlation (r = 0.92) is recorded in Sturt
field. The influence of inertinite is shown to vary between these fields (Fig. 5.7c). While an
apparent negative regression is displayed for the Poolowanna and Tantanna Fields a positive
regression is shown at Sturt and Sturt East. The positive correlation is likely due to
disseminated liptinite (fl uorinite and resinite).
These observations are consistent with the presence of Type II/III kerogen in the
Poolowanna Trough and Type III/II kerogen in the Patchawarra Trough.
117
a) Liptinite r = -0.49 c) lnertinite r = 0.16
l¿
o
o
L o g
o) € ^o
o Otr
I ô o o
o) o o
E 1 o
¿¿
c! f
v)
o
ft
+
U)
I ìù- o + __ß-Ê-
tr
5 o Õ
ô
ô o o
É qt
A ^
"tt AN
025 50 75 100 0 10 20 30 40 50 60 70
Liptinite % lnertinite %
b) Vitrinite r 0.53
l¿ o
o g o Poolowanna
o
L
1
^
o)
qcç Ä Tantanna
O
I o
o)
1
o Sturt
E
C\I tr Sturt East
ct)
+
r
U)
o/
o
0 25 50 75 100
Vitrinite %
Poolowanna General
trend trend
Stained
E
o- 1 ,t____r
o- o
ooo
c Fo
o o À tr Excellent
P
c oô
o o
o
o
c
-ra --T-
o---+tr-tr
q
-o tl tr ry Good
(d
C) 1 : -.t----r
I
o :rr ''-f Good
ìo
I Fair
õ Shale, and Silty,shale Carbonaceous Clastic Coal
P
1
10 50 100
TOC "/"
o Poolowanna À Tantanna O Sturt tr Sturt East
Figure 5.8 Source rock richness based on organic carbon content and hydrocarbon yield
Chøpter 5
In every category the Poolowanna Trough samples appear to have higher hydrocarbon yields
than those in Patchawarra Trough. This is because the Poolowanna Formation in the
Poolowanna Trough is thermally more mature than the sections in the Patchawarra Trough.
A comparison of the Patchawarra Trough sections shows that the Poolowanna Formation is
more mature at Tantanna than in the other fields. As mentioned earlier, hydrocarbon yield
increases with maturation level. Oil staining of these rock samples is not indicated.
The type of hydrocarbon (either gas or oil or both) anticipated to be generated from potential
source rocks is normally assessed using the Rock-Eval parameters, hydrogen index and
S2/S3, and kerogen atomic IVC ratios (Espitalié et al., 1977; Peters, 1986). The atomic FVC
ratios are obtained from elemental analysis (not carried out in this study). The Rock-Eval
data and the anticipated type of hydrocarbon product are shownin Table 5.1. Almost all the
S2/S3 values are >10, and indicative of oil-prone source rocks. Hydrogen index values
forecast the existence of gas, gas and oil and rarely oil-prone source rocks (Table 5.1).
The oil-prone source rock facies is developed in the middle section (2499-2524 m) of the
Poolowanna Formation in the Poolowanna Trough. The gas and oil-prone source rock facies
occurs in the lower section and happens to be the most predominant, whereas the gas-prone
facies is confined to the upper part of the formation.
In Patchawarra Trough, only the gas and oil-prone facies and the gas-prone facies were
identified. In the Tantanna Field silty shales and coals of the upper (1795-1807 m) and
lower (1811-1832) parts of the section are oil and gas-prone whilst the gas-prone facies is
the carbonaceous clastic middle section (1807-1814 m). At Sturt, the gas-prone facies
appears to be predominant and is identified in the carbonaceous clastics and coals of the
upper section (1329-1865 m) of the formation, whereas the carbonaceous clastics and coals
of the lower section (1865-1881 m) are gas and oil-prone. In the Sturt East Field, the gas
and oil-prone facies appears to be the most predominant. It occurs in the upper section
(1829-1850 m) with an occasional, gas-prone interval. The lower section (1862-1865 m) is
a gas prone facies.
180
Chapter 6
Chapter 6
6.1 Introduction
It is well established that autotrophic biosynthetic processes like photosynthesis favour the
lighter isotope 112C¡ over the heavier isotope (13C) during the incorporation of carbon into
live organic tissue (Section 2.5.1). Accordingly, this l3C-depleted isotopic signature is
preserved in sedimentary organic matter (Cranwell et a1.,1987; Poynter et aL.,1989). During
the diagenesis and catagenesis of kerogen, this isotopic signature is passed on, to the
generated hydrocarbons which therefore contain isotopic clues to their source materials.
Aromatic hydrocarbons appear to be prevalent in humic organic matter and highly mature oils
and source rock bitumens. As mentioned in Section 2.7.3., the low-molecular-weight
compounds may be preferentially removed from oils by water-washing.
181
Table 6.1 Bulk composition of crude oils from the Poolowanna and Patchawarra Troughs
Hence, as for the saturates fraction, the carbon isotopic signature of the aromatic fraction is
the average of the isotopic signatures of its constituent compounds, but may be modified by
various fractionation and disproportionation processes.
As shown by their stable carbon isotope type-curves (Galimov, 1973; Stahl, 1978; Chwg et
al.,l9ïl),oils in 13C for fractions
and rock extracts (bitumens) display a general enrichment
of increasing polarity and boiling point. Peters and Moldowan (1993) point out that the
whole oil or bitumen generally shows an isotopic composition between that of its saturates
and aromatics fractions because of mass balance considerations. The bulk carbon isotopic
signatures of these two hydrocarbon fractions in source rocks from the Poolowanna
Formation and a suite of oils from adjacent reservoirs are summarised in Table 6.2a,b.
Table 6.2a Carbon isotopic composition of saturated and aromatic hydrocarbon fractions of
selected source rock extracts from the Poolowanna Formation.
Canonical variable (CV) = -2.53 õ13Cs a¡+ 2.22 ôt'Cu.o - 11.65 (Sofer, 1934).
183
Table 6.2b. Carbon isotopic composition of saturated and aromatic hydrocarbon fractions of selected oil samples
from the Poolowanna Formation and adjacent reservoirs in the western Eromanga Basin.
The range of isotopic variation within the fields is 1.9%o at Poolowanna;2.3%o at Tantanna;
2.6%o at Sturt; and2.8%o at Sturt East. These variations imply that the organic input is more
heterogenous in the Patchawarra Trough than in the Poolowanna Trough.
It is worth noting that samples with the highest üptinite contents have lighter isotopic
signatures, e.g. aresinite-rich sample from Poolowanna-2, (2524 m) has a õ13C value of -
27.5%o; and sporinite-rich samples from Sturt East-l (1841 and 1844 m) have values of -
29.1%o and -28.4%o, respectively. The more mature samples of the Poolowanna Trough
show slightly more positive signatures within a naffow range than do early mature ones from
the Patchawarra Trough.
average value of -26.5%o (Table 6. 2b). Although from reservoirs of different ages, almost
all the oils have very similar isotopic signatures. The waxy Poolowanna-l oil (Fig. 6.2) is an
exception (ôt'Csat = -29.7%o). Its lighter isotopic signature reflects its derivation from a
distinctly different source rock facies.
t'C.at
The Poolowanna-reservoired oils have an average ô value of -26.5%o, and a range of
3.6%o. When the Poolowanna-l, DST 2 sample is excluded from the suite because of its
anomalously negative signature, the isotopic variation is reduced to 0.3%o. The
corresponding average ô13Cs¿1 values (and variation) for the oils in other Eromanga
reservoirs are as follows: Namur, -27.0%o; Birkhead, -26.3 (0.8)%o; and Hutton, -26.7
(0.1)%o. The two oils from Patchawarra (Permian) reservoirs have almost identical ôt3Csat
values of -26.5 and -26.4%o whilst those from the Mooracoochie Volcanics (Cambrian) have
values of -26.3 and -26.5%o. With the exception of the Poolowanna-l crude all oils in
Jurassic reservoirs have isotopic signatures similar to those of the oils in Permian and Pre-
Permian-reservoirs. This implies a conìmon origin, or generation from source rocks with
similar organic matter inputs. Mixing of oils as a result of migration may also be a cause of
similar isotopic signatures.
185
Well: Poolowanna-1 Well: Tantanna-2
Denth:2557 m Deoth: 1811 m
ô l3Csar: -26.6Voo ô f3crut: -25.2%o
Pr/Ph: 6.63 Pr/Ph:8.31
C 15
C I7
Cn
czs
ÁJgaI(597o)
Aleal (687o)
Well: Sturt-l
Well: Sturt-4 Denth: 1871 m
Deoth: 1859 m ô f3crut, -28.6%o
ô 13C.ut' -26.IVoo Pr/Ph:5.11
Pr/Ph: 5.97
czs
c Cn
Figure 6.1 Influence ofdifferent precursorbiota on the z-alkane profiles ofsource rocks
from the Poolowanna Formation. Interpretation based on diagnostic ca¡bon number
ranges from Collister et al. (1994).
czz
Well: Sturt-7
Cn Form.: Poolowanna
DST 5
ô l3Csat: -26.3Voo
PrlPh: 6.75
c 12
Well: Taloola-2
Form.: Hutton
DST 2
ô l3Csat: -26.57oo
Pr/Ph: 5.33
Figure 6.2 Source affinity of selected Eromanga Basin oils as indicated by their ¿-alkane
profiles
Chnpter 6
Oils from multiple stacked reservoirs in the same field show a very narrow range of isotopic
variation. For instance, in Tantanna-l there are three stacked oil pools of which the
Poolowanna has a ôt'Csat value of -26.4%o,while both the Hutton and Birkhead pools have
values of -26.8%o.In Tantanna-2,the Poolowanna and Hutton oil pools differ by O.5%o, and
in Sturt-3, the Poolowanna and Birkhead oils differ by only 0.2%o. In Sturt-6, the
Patchawarra and Cambrian oils have the same value which is 0.5%o lighter than that of the
Birkhead oil. In Sturt-7 the variation between the Patchawarra and Poolowanna oils is
0.2%o, whereas in Taloola-2 the Namur oil is lighter than both the Poolowanna and Hutton
crudes by O.5%o. Therefore either the bulk isotopic signatures of the saturated hydrocarbons
are incapable of differentiating Jurassic from Permian and Cambrian oils, or the oils have
common source.
The overall average ôt'C*o value for the entire Poolowanna Formationis -25.0%o. As is the
case for the saturated hydrocarbons, the 2.6%o variation may indicate the heterogeneity of
source input and possibly maturity differences between the Poolowanna and Patchawarra
Troughs. The indicated source input heterogeneity supports the observations based on
maceral group distributions, which showed the samples to be very diverse especially in the
Patchawarra Trough (Section 4.2: Fig. 4.1).
188
-23
Waxy
-24 o
oo Poolowanna
o
-25 Poolowanna
o --Ã Patchawa¡ra
o -9_
e
èe Patchawarra
9 -26 Non-waxy
O
ca)
râ -2i
-28
-29
-30
-30 -29 -28 -27 -26 -25 -24 -23
õ r3cru¡%o
O Poolowanna O Sturt
Figure 6.3a Waxy and non-waxy character of Poolowanna Formation source rock
extracts as demonstrated by carbon isotopic composition of saturated and a¡omatic
hydrocarbons. The oblique line separating the wCIry and non-waxy extracts corresponds
to CV = 0.47 (after Sofer, 1984).
-23
-24
Waxy
-25
e
èR -26 o Non-waxy
!
U
ca)
.o -27
-28
-29
-30
-30 -29 -28 -27 -26 -25 -24 -23
6 r3cru¡%o
a Namur X Patchawarra
Birkhead * Mooracoochie
I Hutton
O Poolowanna
Figure 6.3b Isotopic composition of saturated and aromatic hydrocarbon fractions of oils
from Jurassic and non- Jurassic reservoirs
Chapter 6
It is noted that the isotopic difference between the saturated and aromatic fractions of source
rocks in the Poolowanna Formation ranges from 0.4%o at Tantanna-z (1811 m) to 3.I%o aI"
Sturt-l (1871 m) with an average value of 1.7%o. This shows that the aromatic fractions are
always isotopically heavier than the saturates. The exact cause of this relationship is not
known; but it is most likely related to the fact that the organic precursors of the aromatic
13C
compounds are formed via processes that result in less discrimination against the heavier
isotope.
It should also be noted that the õt'Caro-sat values of the source rock extracts increase as the
saturates become more waxy (based on GC n-alkane profiles). This is illustrated by
comparingthe¡¿-alkaneprofiles of the Tantanna-2 (1S11m) extract (non-waxy; ôl3Caro-sat
= 0.4 %o) withthat of the Sturt-l (1371 m) extract (waxy; ôt'C*o-rut = 3.1%o: Fig. 6.1).
This tells us that the waxy C1¡¡ n-alkanes are isotopically lighter than the lower molecular
weight alkanes. The difference between the ôl3Csat values of these two samples is 3.4%o;
and only 0.6%o in the case of ôt'Curo values. Clearly, the isotopic composition of the
saturates fraction is less constant than that of the aromatic fraction.
The ôr3C¿¡6-s¿¡ values of the oils range from 0.99%o in the Sturt-3 (DST 18, Birkhead)
crude to 3.60%o in the Poolowanna-l (DST 2, Poolowanna) sample, with an average of
1.5%o. Not suprisingly Poolowanna-l oil is very waxy (Fig. 6.2). This oil might have been
generated from a source rock with characteristics similar to that at Sturt -1 (1871 m),
Other possible causes of large isotopic differences between the saturated and aromatic
hydrocarbon fractions of crude oils are fractionation processes attributable to water washing
and migration (Fuex, 1972; Alekseyev et a\.,1915).
The ô13Cs¿1to õl3C¿¡s ratio does not show much variation in either oils or source rocks. It
varies between 1.02 and l.l2 in source rocks and between 1.04 and l.l4 in oils. As for the
õt'Ca.o-sat parameter, the waxy samples have the highest values and non-waxy samples the
lowest. Because the wax content is routinely attributed to land plant precursors, the observed
isotopic signature in waxy source rock and oil samples also may be confidently assigned to
organic matter inputs of higher plant origin.
t9t
Chapter 6
The relationship between the isotopic compositions of these saturated and aromatic fractions
was further evaluated by applying the canonical variable (CV) method devised by Sofer
(1984). Values for Poolowanna Formation source rocks range ftom -2.94 in siltstone at
Tantanna-2 (1811 m)to 4.I4 in carbonaceous clastic at Sturt-l (1871 m). No systematic
differences were observed between the Patchaw¿ura and Poolowanna Troughs (Fig. 6.3a).
Waxy and non-waxy source rocks occur in both troughs. In each source rock category, the
samples from the Poolowanna Trough tend to be isotopically heavier, as is discussed below.
According to Sofer (1984), CV values less than 0.47 are indicative of strong marine
influence.'Whether the non-waxy source rocks (CV < 0.47) in this study (Table 6.2a, Fig.
6.3a) have any marine affinity is questionable because no other marine features have been
reported for the Poolowanna Formation in the Patchawarra Trough. However, some marine
influence has been recorded elsewhere in the Poolowanna Trough (Wiltshire, 1978).
McKirdy et al. (1986b) attribute CV values in the non-waxy range to algal and bacterial
precursors, not necessary of marine origin.
The CV values of the oils range from -1.41 (Sturt-3, Birkhead) to 5.53, (Poolowanna-1,
Poolowanna). Almost all the Jurassic oils have values much less than O.47 and are thus
indicated to have an algal or bacterial affinity. The Poolowanna-l, (Poolowanna) and
Tantanna-l, (Hutton) have CV values of 5.53 and O.73 respectively, thus signifying their
strong terrigenous affinity. All the Permian oils and one of the two Cambrian
(Mooracoochie) oils also have CV values much greater than 0.47 and are obviously derived
mainly from terrigenous plant material.
These results show that the Jurassic-reservoired oils have a different origin from those in
Permian and Cambrian reservoirs. Regardless of source affinity two families of oils and
source rocks are recognised based on their CV values (greater or less than 0.47). Both oil
families could have originated from Poolowanna Formation source rocks, as discussed in
more detail below.
192
-24.
Sturt-6 Sturt-7
-24.5 x
(n=2)¡
o
a
-25.00
*
-25.50
-27.50 -27.00 -26.50 -26.00
-24.00
Tantanna-1
I
%\ -24.50
ä
o
P o Reservoir
.ê
O Namur
-25.00
A Birkhead
A I Hutton
o Poolowanna
-25.50
X Patchawarra
-24.00 ¡ Mooracoochie
Taloola-2 (Cambrian)
Waxy
-24.50 Non waxy
25.00
o
T
-25.50
-27.50 -27.00 -26.00
ô 13Osat %o
Figure 6.4. Comparison of the isotopic signatures of oils in selected stacked reservoirs
Chapter 6
Both waxy and non-waxy organic inputs appear to characterise the Poolowanna Formation
source rock facies of the study area. In the waxy zone (non-marine) the Poolowanna Trough
samples plot as isotopically more positive than the Patchawarra Trough samples. A Similar
separation is observed in the non-waxy (marine) zone (Fig. 6.3a). This offset towards more
positive (i.e. heavier) values is most likely to be caused by the greater maturity of the
Poolowanna Trough samples.
The distinction between Jurassic oils and those from Permian and Carnbrian reservoirs is
quite subtle, as shown in Figure 6.3b. With the exception of one Canrbrian (Mooracoochie)
sample, the Permian and pre-Permian oils plot in the waxy zone, which is in agreement with
their terrestrial plant affinity. Most of the Jurassic oils have a non-waxy, bacterial signature.
The Poolowanna-l (Poolowanna) crude, which happens to have a very strong waxy
signature, is an obvious exception. Here, land plant lipids clearly outweigh the bacterial
input to the rock.
On the basis of the isotopic evidence, the Poolowanna source rocks are possible sources for
the Eromanga-reservoired oils. Generally, the non-waxy source rock facies appear to have
played a more important role in the generation of these oils than the waxy facies. The waxy
Poolowanna-l (Poolowanna) oil is likely to have been generated by a local waxy organic
facies with characteristics similar to those of carbonaceous silty shale at Sturt-l (1871 m). Its
position in the cross-plot, apart from organic matter input, might have been influenced by
water washing (Sofer, 1984).
Sturt-6 The waxy Patchawarra (Permian) and Mooracoochie (Cambrian) oils plot in the same
spot indicating their common origin, probably from an intra-Permian source. The non-waxy
Birkhead oil is well separated and thus of different origin (probably Jurassic).
Sturt-7 Here the Permian (Patchawarra) oil plots in the waxy zone, whilst the Cambrian
(Mooracoochie) and the Jurassic (Poolowanna) oils plot in the non-waxy zone. In this case
the Mooracoochie oil plots together with the Poolowanna oils and is likely to be derived from
a similar non-waxy Jurassic source.
Tantanna-l These three oils are of similar origin, and plot close to the line (CV = 0.47)
which separates waxy from non-waxy crudes. On balance, they are probably of Jurassic
origin. However, it is not possible to determine from their isotopic composition alone
whether the source is the Poolowanna or the Birkhead Formation.
t94
Table 6.3a Gas chromatographic data on saturated hydrocarbon fractions of rock extracts, Poolowanna Formation.
35 t5,25
STEl-38 1841 nCrr-nC* C,,,,, 7.25 0.52 0.08 1.26 1.01 r.12 1.25 29 45 26
STEl-39 t84,/¡ nCrr-nC' C,, 5.38 0.80 0.16 1.35 0.99 1.29 r.23 34 46 20
STE2-40 1829 nCrr-nC' cr.,.r., 3.6r 0.62 0.17 1.27 1.01 1.11 r.27 29 47 24
STE3-41 1850 nCrr-nC' cru,r, 6.02 0.73 0.12 1.31 0.99 r.l4 1.30 25 50 25
STE4-43 1838 nCrr-nC' cru,r, 5.81 0.62 0.10 1.23 0.97 t.t2 1.24 25 49 26
STH-4 1850 nCr-nC' cr" 8.28 0.69 0.10 t.o2 0.98 1.08 1.24 50 40 10
STB+45 1862 nC,^-nC"" Cro r. 8.44 1.85 0.2r 1.18 0.99 1.08 1.33 32 53 15
I
*c27 * + +
CPI =
2 Cz+*cz6* cz8 * czo*c3z
t
-t)t -
OEP =
Algae
Field
o Poolowanna
Â
r Tantanna
o
€ o Sturt
æ
a o Sturt East
A
tr
E
o 8q
o
Algae 16-19
Bacteria and
aquatic plants 20-26
Higher plant 27 -3t*
waxes
x Certain freshwater algae (e.g. Botryococcus) may also contribute to this range
Figure 6.5a Source rock affinity to organic matter inputs based on n-alkane distributions,
Poolowa¡na Formation
Chapter 6
Taloola-2 Like in Tantanna-l, these three oils have very similar carbon isotopic signatures
for both the saturated and aromatic hydrocarbon fractions. It is probable that they are of the
same origin. Their source rock is almost certainly of Jurassic rather than Permian age (i.e.
B irkhead and/or Poolowanna).
The above observations suggest that the oils from Jurassic reservoirs have a different origin
from that of the Permian oils. One of the Mooracoochie oils (Sturt-6) is of Permian origin,
and the other (Sturt-7) is clearly of Jurassic origin.
The relative abundance of these three diagenetic-alkane groups for the Poolowanna
Formation source rock extracts are presented in Table 6.3a, and for the crude oils in Table
6.3b. The resulting implications for the major organic matter inputs to the source rock are
illustrated by a ternary diagram (Fig. 6.5a). The corresponding diagram for the oils (Fig.
6.5b) demonstrates their source affinity.
A strong contribution of organic matter to the Poolowanna Formation by algae, bacteria and
aquatic plants is indicated for the Poolowanna Trough. Here, the relative abundance of the
algal input ranges from33%o at2423 m in Poolowanna-3 to 597o at 2557 m in Poolowanna-
1. An algal presence would also account for the lamalginite noted in this area (Chapter 4).
The Patchawarra Trough samples are characterised by a stronger input from bacteria and
aquatic plants: 2349 Vo atTantanna;4Ç537o at Sturt East. However, an anomalously high
algal contribution (68Vo) is also noted in Tantanna (Tantanna-2, 1811 m).
r97
Table 6.3b Ga.s chromatographic data on saturated hydrocarbon fractions of crude oils from the Poolowanna and Patchawarra Troughs
s13 Sturt-4 Poolowanna 1 1872-1880 nCro-nC' cro 6.97 0.84 0.12 t.2l 1.01 1.03 t.t1 4l 50 9
s14 Sturt-4 Poolowanna 2 1883-1892 nCro-nC* Cro 6.69 0.87 0.13 r.15 0.98 1.02 1.07 40 50 10
s15 Sturt-5 Poolowanna 1 r858-1866 nCro-nCru C,, 5.38 0.97 0.18 LI4 t.o2 1.01 t.t3 36 50 t4
s16 Sturt-6 Mooracoochie 6 t9t4-t919 nCro-nCro C,, 6.01 0.79 0.14 1.18 1.01 1.04 1.06 40 50 10
s17 Sturt-6 Patchawarra 3 1884-1898 nCro-nC' C,, 6.3r 0.72 0.r2 1.22 0.99 1.03 l.r7 45 48 8
s18 Sturt-6 Birkhead 1 1883-r892 nCro-nC, c,., 6.79 1.06 0.16 1.19 0.98 1.04 1.15 44 48 8
sl9 Sturt-7 Patchawarra I 1923-1938 nC,o-nCro C,, 5.7r 0.74 0.13 1.15 0.99 1.06 1.08 39 50 11
s20 Sturt-7 Mooracoochie 2 1946-2021 nCro-nC, c,n 6.80 0.73 0.11 r.20 0.99 1,.07 1.26 4 47 9
s21 Sturt-7 Poolowanna 3 l87t-1876 nCr-nC* cß 6 75 0.87 0.14 1.1 1 1.02 1.04 1.11 39 50 t2
s22 Sturt-7 Poolowanna 5 1885-1889 nCr-nC, C,, 6 75 0.85 0.13 1.1 7 0.73 1.06 r.t2 53 39 9
s23 Sturt-8 Poolowanna 1 1880-1884 nCn-nCro c14, cl s7 2t 0.86 0.12 1.1 1 1.00 t.02 r.04 39 51 10
SF'24 Sturt East-2 Poolowanna 1 1845-1899 nCr-nCro c t2 7.55 0.85 0.11 1.1 8 0.99 1.03 1.08 4l 49 10
TN,25 Taloola-2 Poolowanna 1 1829-1836 nCr-nC* C t2 5.01 0.3s 0.07 I 1 9 1.00 1.07 t.r4 4l 50 9
T¡J-26 Taloola-2 Hutton 2 t789-1793 nCr-nC' c t2 5.33 0.33 0.06 1 24 1.01 3.05 r.o2 34 6t 5
TN-27 Taloola-2 Namur 3 r384-1397 nCo-nC"" C 1a 7.40 0.73 0.10 1 25 1.00 1.07 1.19 40 52 9
Algae
Reservoir
a Namur
Birkhead
T Hutton
a Poolowanna
x Patchawarra
x Mooracoochie
o
A
_ir.
Figure 6.5b Oil source affinity based on z-alkane distributions, Poolowanna and
Patchawarra Troughs. Key as for Figure 6.5a.
Chnpter 6
The contribution of higher plant waxes is shown to be relatively lean in all fields. However,
there is an app¿ìrent trend of increasing higher plant input from Poolowanna to Sturt East
(Fig. 6.5a). The presence of telalginite, especially Botryococcus-like alginite, in samples
from the Patchawarra Trough (Chapter 4), would tend to enhance this trend.
A strong affînity with algal, bacterial and aquatic plant precursors is inferred for almost all
the oils based on their n-alkane group distributions (Fig. 6.5b). The highest algal input
(537o) is recorded in the Poolowanna-reservoired oils from Sturt-3 and Sturt-7 whereas the
highest bacterial and aquatic plant contribution (6IVo) is recorded in the Hutton oil from
Taloola-2. The higher plant wax contribution is relatively low in all the anaþsed oils.
The highestvalue of 27Vo is recorded in the Poolowanna-l (Poolowanna) oil whilst the rest
of the oils have values ranging ftom 57o to l4Vo.
The inferred source aff,rnity of oils plotted in Figure 6.5b is remarkably uniform. Of all the
reservoir units sampled, the Poolowanna Formation and, and to a lesser extent, the Hutton
Sandstone display the greatest variability in oil source affinity. This is attributable to ranging
degrees of in situ alteration of the oils in the latter reservoirs by water washing.
Finally, the n-alkane-based compositional fields of the source rocks (Fig. 6.5a) and the
crude oils (Fig. 6.5b) overlap. Thus, this method cannot be used to distinguish oils of
Jurassic and Permian origin.
200
a) Algae b) Bacteria and aquatic plants
70 /^ 60
r = 0.54 / r = -0.20 o
/ o
òs
60 o / o
o o/ / 50 ÀÂ A
o
o
É @ ¡__.o__g
€s0 o o
o oo o
o
)
-o
o
A 40
3¿o / oo o
þ o
tt(€ / ÂÀ
tr
/ 30
.9 ¡o /
/ À
o/ o À
20 20
30 -29 -28 -27 -26 -25 -29
l3Csat%o õ l3Csat%o
õ
\o r = -0.71
25
À
o
Field
èe
o \\o A o
o Poolowanna
cl 15 o\ tr
Ë
o
o ¡ Tantanna
-o
C\I 1 ù,o o
Ä
o Sturt
C)
o
o o Sturt East
(d 5
c)
oo
ú
0
-30 -29 -28 -27
õ L3CsatVoo
Figure 6.6 Relationship between source biota and saturates isotopic signature in
Poolowanna Formation extracts
a) Algae b) Bacteria and aquatic plants
60 65
*è
a 60
5- 50
o A
o 55
É
(ll T f*^ a
T
-B+o
(d
a &\ 50 a ¡
*î
o) o
'tr T 45 ^
(d
õ30
ú 40
oa
o
20 35
-29 -28 -27 -30 -29 -28 -27 -26 -25
6l3Csat7oo 6l3Csut%o
Ëzo A Birkhead
É I Hutton
rs
-ã
(d
o)
a O Poolowanna
'Ë 10 X Patchawarra
a X Mooracoochie
ú
0,)
{å
5 I
0
-30 -29 -28 -27 -26 -25
õ l3Ctut%o
Figure 6.7 Relationships of saturates isotopic signature to the infened source affinity of
crude oils in the Poolowanna and Patchawarra Troughs
Chapter 6
From the above illustrations, it can be deduced that the saturates isotopic signature, of the
Poolowanna Formation in the Poolowanna Trough is related to a freshwater algal input
whereas in the case of the Sturt and Sturt East samples higher plants exert a major influence.
The bacterial influence is relatively minor.
It is concluded that the isotopically lighter signatures of the saturates fraction result from a
progressive increase of the higher plant wax input, whereas the isotopically heavier
signatures coincide with strong algal inputs. These two observations agree with the
interpretation of the CV data.
In this study, the maturation effect was examined by comparing the õl3C signatures of the
saturated and aromatic hydrocarbon fractions with measured vitrinite reflectance (Rs) in the
caseof source rocks and calculated vitrinite reflectance (Rç-1 and Rs-2)in the case of oils and
source rocks. Rs-l is the calculated reflectance proposed by Radke and Welte (1983) and
Rç-2 is the one suggested by Boreham et al. (1988).
Cross-plots of the õt'C*o signatures versus R6 and Rc are shown in Figure 6.8. The
ôt3Csat of the rock extracts showed only a weak positive correlation with maturity, yielding
regression coefficients of 0.59 for Rç- 1; 0.60 for Rs-2 and 0.33 for Rs. In contrast, no such
correlation was observed in the oils. This suggests that the effect of maturation on the ôl3C
signature of the saturates is slight to non-existent.
The ð13C signature of the aromatic fraction of the Poolowanna rock extracts showed positive
correlationswithmaturity (r = 0.65 for Re and 0.61 for both Rs-l and Rc-2).The general
trend for the whole Poolowanna source rock population is illustrated in Figure 6.8a.
203
0.9
sd
o
0.7
/
èe
A
o //
o /À
ú
./
0.6 / ./ A
tr o
tr -/Q
/ /
e*
0.5 -
0.4
-27 -26 -25 -24
6l3Caro%o
0.90
Poolowanna : Ro> 0.6 Vo Field
r =0.93
o Poolowanna
0.85 o o A Tantanna
o Sturt
o tr Sturt East
Ba
0.80
& o
0.75
o
0.70
-25.5 -25.0 -24.5 -24.O
õ r3Cuo%o
Figure 6.8b Variation of a¡omatic õ 13C signature with maturity (Ro) in Poolowanna
Trough source rock
Rc-l =0.378 ô13C*o +10.18; r=0.72
úo //a a
A
0.5
A
I
V
A
0.3
-26.5 -26.0 -25.5 -25.0 -24.5 -24.0
õ r3CæoVoo
Reservoir
a Namur X Patchawarra (Permian)
a Poolowanna
Figure 6.8c Variation of aromatic isotopic signature with oil maturity based on calculated
reflectances (Rc-1 and Rc-2)
Chnpter 6
Once again it is shown that the source rock in the Poolowanna Field is more mature than
those in neighbouring fields of the Patchawarra Trough. A very strong positive correlation (r
= 0.93) was observed when the mature Poolowanna Trough samples were plotted alone
(Fig. 6.8b). This observation suggests that at high maturation levels (>O.6Vo R6;, the õr3C
signature of aromatic fraction become more positive with increasing maturation, particularly
in similar lithological facies.
The total suite of oil aromatic fractions also showed a positive correlation with maturity (r =
0.34 for both Rç-l and Rç-2). Because of the extreme ôt'C*o value (-26.I%o) of the
Poolowanna-l oil, this sample was excluded for the purposes of correlation (Fuex, 1977).
After its exclusion, a stronger positive correlation resulted (r = O32 for Rs-l, and 0.71 for
Rc-2). In both cases, it is shown that there is an apparent positive increase of ô13C signature
of aromatic fraction with increasing maturity. A more pronounced positive correlation is also
observed in the case of calculated reflectance values >0.6Vo Rc Gig. 6.8c).
It is clearly shown by Figure 6.8c that the Permian and Canrbrian oils are isotopically more
positive and thus more mature than the overþing Jurassic oils. It can also be deduced from
this figure that the Tantanna-l (Poolowanna) oil originated from a Cooper Basin source
rock. The Jurassic oils from Birkhead Formation, Namur Sandstone and Hutton Sandstone
and the rest of those from the Poolowanna Formation, are shown to be quite separated from
those in Patchawarra Formation and Mooracoochie Volcanics because of their lower
maturity. The Sturt-7 (Mooracoochie) oil plots in the same maturity domain as the bulk of the
Jurassic oils. However, biomarker evidence (Chapter 7) suggests that it may not be of
Jurassic origin.
206
ChapterT
Chapter 7
7 .l Introduction
Saturated hydrocarbon biomarkers are part of the suite of molecular fossils derived from
once living organisms. The group of molecular fossils dealt with in this chapter are normally
encountered in source rock extracts (bitumen) and oils and include a range of acyclic normal
alkanes, acyclic isoprenoids, and cyclic alkanes (viz. sesquiterpanes, triterpanes and
steranes).
The relative distribution and abundance of these biomarker alkanes is used to identi$ the
type of biota (algae, bacteria and land plants) which contributed organic matter to the source
rocks. In the case of oils, the biomarkers help determine their source affinity by providing
information on the organic facies and depositional environment of the source rock. Source
rock and oil maturity assessment is also possible using biomarkers (Peters and Moldowan,
1993). Oils can then be grouped into families of the same source and maturity and each
family be correlated with a potential source rock.
207
nctz
Well: Poolowanna-2
Depth: 2542m
Pr/Ph: 3.81
Suboxic to oxic
ncn
nc:o
nCt¡
Well: Tantanna-2
Depth: 1799 m
PrlPh: 7.63
Mixed inpul oxic
nCzz
nCzs
Well: Sturt East-l
Depth:1844 m
PrlPh: 5.38
Oxic
Pr
n Ct¿ Well: Sturt East-4
Depth:1862m
nCzz Pr/Ph: 8.44
nC3g Very oxic
nC 0
Figure 7.1 Gas chromatograms of alkanes in Poolowanna Formation source rock facies:
implications for source biota and depositional environment
Chapter 7
This feature is possibly caused by an unusual mix of bacteria, aquatic plants and higher
plants.
The regular isoprenoids, pristane (C1e) and phytane (Czo), are clearly identified (Fig. 7.1).
Pristane is commonly predominant over phytane. The pristane to phytane ratio (prlph) ranges
from2.29 in coaly siltstone (Tantanna- ,1807 m) to 8.44 in coal (Sturt East-4, 1862 m).
The averagepr/ph value for the Poolowanna Trough is 4.56 and that for the Patchawarra
Trough is 6.21, implying differences in depositional environment and/or organic matter input
between these troughs. An anomalously high pristane peak is noted in the Sturt East-4 (1862
m) coal sample (Fig. 7.1). Such an abundance may indicate more than one source for the
pristane (Goosens et a1.,1984), although another explanation is more likely (see below).
Pristane to n-heptadecane ratios (prln-Cs) vary from 0.15 in a Poolowanna Trough siltstone
(Poolowanna-2,2542 m) to 1.85 in the coals of the Patchawarra Trough (Sturt East-4, 1862
m). The average value for pr/n-Cs in the Poolowanna Trough is 0.35, whereas that in
Patchawarra Trough is 0.71. According to the criteria suggested by Lijmbach (1975), the
Poolowanna Trough source rock facies appear to have been deposited in relatively open
water environments with abundant bacterial activity, whereas the Patchawarra Trough source
rock facies were deposited in environments with alternating peat swamp and open water
conditions. This observation supports the organic petrographic data which shows inertinite
(inertodetrinite) occurring in association with liptinite (telalginite) (Chapter 4). The high
value observed at 1862 m in Sturt East-4 (prln-Cs >1.0) clearly indicates the persistence of
peat-swamp conditions in that area.
Phytane to n-octadecane ratios (ph/n-C13) vary from 0.04 in the Poolowanna Field to 0.21 in
Sturt East Field. The relative lower values shown in the Poolowanna Trough are probably
indicative of the higher maturity of the Poolowanna Formation in this area. However, it is
possible that this ratio also might have been influenced by depositional conditions.
Carbon preference index (CPI) and odd-even predominance (OEP) trends are as shown in
Table 6.3a. The CPI values in both the Poolowanna and Patchawarra Troughs ile very
similar with average values of 1.14 and L24, rcspectively. These values normally are
indicative of mature source rocks (Bray and Evans, 196l). As was noted from the vitrinite
reflectance data, the Patchawarra Trough samples are somewhat less mature, and this would
account for their slightly higher CPI values.
(-l)t* t
The OEP values, centred at i=16, i=22 and i=26 (Scalan and
Smith, 1970) show only very slight differences between the two troughs.
209
nc14
Tantanna-l, DST 3
Birkhead/Hutton
nCtz Pr/Ph: 8.9
nctt
nczg
Taloola-2, DST 2
Hutton
PrÆh:5.3
Pr
Sturt-8, DST I
Poolowanna
PrlPh 7.2
Sturt-7, DST 2
Mooracoochie (Cambrian)
PrlPh: 1.3
Figure 7.2a Ga,s chromatograms of alkanes in selected oils from Jurassic and non
-Jurassic reservoirs in the Patchawara Trough
Chapter 7
The average OEP values for the Poolowanna Trough are 0.99, 1.04 and 1.08, respectively;
and for the Patchawarra Trough they arc 0.99, 1.10 and 1.27 (Table 6.3a). As pointed out
previously, the samples from the former depocentre are slightly more mature than the latter;
and this is entirely consistent with values shown in the Patchawarra Trough source rocks.
A tentative quantification of source biota inputs, based on the relative abundance of three ¡¿-
alkane groups (Collister et aI., 1994: Fig. 6.5a), shows that the Poolowanna Trough on
average is characterised by almost equal contributions from algae (467o), and bacteria(457o),
and lesser contribution from higher plant waxes (9Vo). The Patchawarra Trough is
characterised by high average contributions from bacteria and aquatic plants (47Vo), followed
by algae (34Vo) and higher plant waxes (l8Vo). Although both areas have low contributions
from higher plant waxes, the input to the Poolowanna Trough is half that to the Patchawarra
Trough. The higher algal input to the Poolowanna Trough is possibly of the lamalginite type
(Chapter 4).
a number of the oils have n-Cmax between Cp to C1a. Although primarily controlled by
source biota these maxima are likely to be modified by cracking of medium molecular weight
n-alkanes (Crs -Crz). Such thermal effects are only likely to be significant in the more mature
(i.e. Permian) oils. The unique Poolowanna-l (Poolowanna) oil (Fig. 6.2) has the carbon
range tp to n-C3s and a peak at n-Cy, thus signifying strong bacterial and land plant inputs.
In all the oil samples analysed in this study no clear distinction can be made between Jurassic
and Permian/pre-Permian oils on the basis of their n-alkane profiles.
The pristane to phytane ratios of all but two of the oils range from 4.98 in the Poolowanna-l
(Poolowanna) sample to 8.88 in the Tantanna-l (BirkheadÆIutton) crude suggesting that
their source rocks were deposited in peat-swamp conditions. The two exceptions are the
Tantanna-2 (Poolowanna) and Stut-7 (Mooracoochie) oils with prþh values of 2.29 and
1.25, respectively (Table 6.3b). The latter oil may be of pre-Permian (possibly Cambrian)
origin (see below), whereas the former sample could be a mixture of the pre-Permian and
Jurassic oils.
2tl
ncn
Birkhead - Jurassic
DST 1
Pr/Ph:6.8
Pr Pr/nC17: 1.06
ncro
nCzs
Patchawarra - Permian
DST 3
PrÆh: 6.3
PrlnCg:0.72
Mooracoochie- Cambrian
DST 6
Pr/Ph: 6.0
PrlnC1¡: 0.79
Figure 7.2b Gas chromatograms of alkanes in oils from stacked reservoirs at Sturt-6,
Patchawa.rra Trough
ChapterT
The average prlph ratio of 4.56 In the Poolowanna Trough source rocks almost coincides
with that of the Poolowanna Trough oil (4.98) whereas the average prþh ratio value 6.2I for
the Patchawarra Trough source rocks falls well within the observed range of oil values in
this trough. This is one piece of evidence that some of the oils in these troughs were
generated from Poolowanna Formation source rocks.
Thepr/n-C17 values vary from O.2l in the Poolowanna Trough oil (Poolowanna-1, DST 2;
Fig.6.2) to a maximum of 1.06 in the Patchawarra Trough oils (Sturt-6, DST 1; Fig. 7.2b).
A value of O.2l is indicative of a source rock deposited in an environment of open water
sedimentation. This value falls within the observed range of source rock values in the
Poolowanna Trough (0.15-0.64). Values between 0.5 and 1.0 are indicative of source rocks
deposited in environment with alternating swamp and open water conditions, Most of the
Poolowanna Formation source rocks in the Patchawarra Trough have prln-C17 values in this
range. The high value (prln-Cn = 1.06) shown by the Sturt-6 (Birkhead) oil indicates that
this oil was derived from a source rock deposited under peat-swamp conditions. It can be
correlated to the Sturt East-4 source rock sample (1862 m) which has a prln-Cs value of
1.S5 (Fig. 7.1). k is also noted that the Permian and Cambrian oils have Prln-C17 values
between 0.12 and 0.79 whilst most of the Patchawarra Trough Jurassic oils have values
>0.80.
ThePhln-C13 values range from0.04 in the Poolowanna-l oil to 0.19 in the Patchawarra
Trough oils with the one exception of the Sturt-7 (Mooracoochie) oil which has a value of
0.60. Once again there is good agreement between the oils and Poolowanna Formation
source rocks in the two troughs. The source of the Sturt-7 (Mooracoochie) crude is likely to
be quite different from those of the Jurassic and Permian oils and the other Cambrian oil.
The CPI values of the oils range from 1.02 in the Tantanna-2 (Poolowanna) oil to 1.48 in the
Tantanna-l (BirkheadÆIutton) oil This spread of values reflects a range of affinities and
maturities (see below). The OEP values =1 indicate mature oils. A high OEP value (3.05)
centred ati=22 in the Taloola-2 (Hutton) oil supports the previously suggested indication of
a high bacterial contribution to the source organic matter (Fig. 6.2). An even-carbon-number
predominance (OEP <1) is shown by the Sturt-2 (Poolowanna) oil. This might be suggestive
of an unusual source rock facies.
A tentative quantifîcation of source biota based on the relative abundance of certain ranges of
n-alkanes (Fig. 6.5b) suggests that most of the alkanes are derived from bacterial and algal
organic precursors. The contribution of land plants is shown to be very low. However, these
data need to be treated cautiously because oil n-alkane profiles are also influenced by their
maturation level and by the possibility of oil cracking in the reservoir.
2t3
Patchawarra TFough
8 Sturt East-4
1862m
m/2123
9 3
1
10
4 7
2
6
5
Poolowanna-1
Poolowanna Tlough 8 2417 m
3
mlz 109 1 \ 4 5
910 2 I 6
mlzl23
á á, -.-,-¡ -- i_^
mlz 137
mlz I79
Time
No m C,o Cro cÚ C,, C* C,, C,, Cru C,u Cru Dr(3) Bi-Sesq.*xx Bi-Sesq.***
Vo Vo 7o Vo 7o Vo Vo Vo Vo Vo
7
POL1-3 Poolowanna-l 2438 3.77 2.83 7.55 3.77 15.09 3.77 1.89 5.66 1.89 53.77 3.56 o.26 0.08
POL1-5 Poolowanna-l 2499 4.04 2.02 8.08 4.04 12.12 7.07 0.00 6.06 0.00 56.57 4.67 0.10 0.28
POL2-9 Poolowanna-Z 2524 2.84 2.27 17.05 t0.23 17.05 7.39 4.55 4.55 r.70 32.39 1.90 0.25 0.18
POL3-13 Poolowanna-3 2505 1.89 2.83 9.43 5.66 r1.32 3.77 0.94 7.55 2.83 53.77 4.75 o.25 0.16
TAN2-18 Tantanna-2 1799 7.87 3.15 6.30 3.94 15.75 6.30 3.15 4.72 r.57 47.24 3.00 0.74 o.32
TAN2-20 Tantanna-2 1811 0.00 0.00 5.66 s.66 15.09 8.49 2.83 7.55 3.77 50.94 3.58 2.10 0.47
TAN3-21 Tantanna-3 1807 2.38 1.59 9.52 4.76 15.87 1 1.11 3.t7 7.14 0.00 44.44 2.80 0.39 0.16
TANS-25 Tantanna-8 18231 6.t4 6.74 12.36 5.62 15.73 tt.24 2.25 4.49 3.37 31.46 2.00 0.28 o.26
STU1-27 Sturt-l 1871 3.06 1.02 4.08 3.06 10.20 6.t2 3.06 7.14 3.06 59.18 5.30 T.T6 0.42
STU4-30 Sturt-4 1859 8.77 2.63 6.t4 4.39 12.28 5.26 r.75 4.39 t.75 52.63 4.29 1.64 0.37
STU4-31 Sturt-4 1881t 8.26 5.50 12.84 7.80 20.64 8.26 4.r3 4.59 2.75 25.23 t.22 0.26 o.l2
STU6-35 Sturt-6 18657 9.03 5.16 7.74 5.81 19.35 7.74 1.29 5.16 t.29 37.42 r.43 o.76 0.27
STE1-38 Sh¡rtEast-l 1841 4.30 1.72 6.02 5.16 13.77 6.88 4.30 6.02 0.17 51.64 3.75 5.99 0.81
STB1-45 Sturt East-4 18621 15.52 6.90 10.34 5.17 13.79 6.90 2.87 3.45 1.72 33.33 2.42 0.16 0.38
* CroÞ hopane 'c* > Crr-Cn steranes x*x > Bicyclic sesquiterpanes
1-10 refer to peak assignments in Figure 7.3; peak 3 = SÞ(H)-drimane; peak 8 = 8Þ(Ð-homodrimane
t coaly facies
Chapter 7
range of 0.08-0.28 (average = 0.16) in the Poolowanna Trough source rocks, and a range of
0.16-0.81 (average = 0.36) in the Patchawarra Trough. Once again, the higher relative
abundance of bicyclic sesquiterpanes in the rocks from the Poolowanna Trough is consistent
with more intense bacterial reworking of detrital organic matter here than in the Patchawarra
Trough. However, the Patchawarra Trough coal facies (Tantanna-8, 1823 m; Sturt-4, 1881
m; Sturt-6, 1865 m) show values similar to those of the carbonaceous shales in the
Poolowanna Trough.
2r6
8
7
mlzl23
Patchawana
mlz I79 DST 3
^
mlzl23
mlz I79
Mooracoochie
DST6
-r{.
mlz t09
mlzl23
Poolowanna
DST 2
mlz 137
mlz 179
(Time)
no no m cro Cro C* C,, Cß C,, Cl5 cru cru C,u Dr (3) Bi-Sesq. Bi-Sesq.
Vo Vo Vo Vo Vo Vo 7o Vo Vo Vo
0.0 0.0 y--T-7.4 20T l.+ 0.9 9.3 0.9 50.0 2 45 1.51 0.59
T2 Tantanna-1 1 1800-1814 Poolowanna 2.3 4.O r7.l 9.1 15.4 10.3 2.3 6.3 L.7 31.4 2.04 0.16 0.07
T3 Tantanna-1 2 1635-1642 Hutton t.7 0.8 9.2 3.4 16.8 12.6 3.4 5.0 0.0 47.1 2.80 o.26 0.20
T4 Tantanna-1 3 t347-1359 Birkhead/Hutton 0.0 0.0 7.9 5.9 11.9 7 .9 2.O 7.9 1.0 55.4 4.66 0.45 0.33
s11 Sturt-3 1A 1848-1855 Poolowanna 0.0 0.0 7.8 6.2 23.3 7.8 2.3 7.8 2.3 42.6 1.83 1.70 0.94
s12 Sturt-3 1B t654-t662 Birkhead 3.0 2.3 10.5 7.5 18.0 9.0 2.3 4.5 2.3 40.6 2.26 0.59 0.23
s13 Sturt-4 | 1872-1880 Poolowanna 2.8 2.8 9.2 5.7 22.7 8.5 4.3 6.4 0.0 37.6 r.66 o.64 0.35
s16 Sturt-6 6 1914-1919 Mooracoochie 3.7 4.3 12.3 8.0 20.4 I 1.1 2.5 4.9 0.0 32.7 1.60 0.38 0.22
s17 Sturt-6 3 1884-1898 Patchawara 0.7 1.4 10.1 8.0 r 8.8 8.7 2.2 7.2 4.3 38.4 2.04 0.71 0.36
s18 Sturt-6 I 1883-1892 Birkhead 1.8 2.7 7.1 6.2 15.9 8.0 2.7 6.2 1.8 47.8 3.00 0.48 0.27
s19 Sturt-7 I 1923-1938 Patchawarra 2.8 3.5 11.1 8.3 19.4 8.3 4.9 5.6 0.0 36.1 1.86 0.73 0.29
s20 Sturt-7 2 1946-2021 Mooracoochie 1.2 2.4 t2.l 10.3 24.2 7.3 3.6 4.8 3.6 30.3 r.25 o.75 0.32
s21 Sturt-7 3 1871-1876 Poolowanna 2.6 4.0 tr.9 7.9 17.2 9.3 4.0 6.6 1.3 35.1 2.04 0.36 o.22
s23 Sturt-8 1 1880-1884 Poolowanna 0.0 0.0 7.9 7.9 19.3 8.8 3.5 7.0 0.0 45.6 2.36 3.46 0.75
SE24 Sturt East-2 I 1845-1899 Poolowanna 1.9 2.5 8.7 6.8 19.9 17.4 2.5 5.6 I.2 33.5 1.68 0.39 0.20
TL25 Taloola-2 I 1829-1836 Poolowanna 1.7 1.7 10.0 6.7 15.0 8.3 4.2 7.5 0.0 45.0 3.00 0.44 0.23
TL26 Taloola-2 2 L789-1793 Hutton 2.1 2.r 12.8 8.5 15.6 9.9 4.3 6.4 0.0 38.3 2.46 0.30 0.19
TL27 Taloola-2 3 1384-1397 Namur 1.7 t.7 ro.2 6.8 13.6 6.8 1.7 7.6 3.4 46.6 3.43 1.20 0.31
Assuming that the bicyclic sesquiterpanes of the drimane series arise from microbial
degradation of triteqpanes (Alexander et aL, 1983), the hopane to bicyclic sesquiterpane ratio
in the Poolowanna and Patchawarra Trough oils may be used to compare the extent of this
process in the organic matterof their respective source rocks. The values of the ratio range
from 0.16 in the Tantanna-l (Poolowanna) oil to3.46 in the Sturt-8 (Poolowanna) oil (Table
7.lb). The Permian and Cambrian oils show a relatively narrow range of values (0.38-0.75)
compared to that in the Jurassic oils (0.16-3.46) of the Patchawarra Trough. The
Poolowanna-l (Poolowanna) oil in the Poolowanna Trough has a relatively high value of
1.51. Suchhighvalues (>1) probably indicate less intense bacterial reworking of the oil-
source organic matter. With the exception of the Sturt-8 (Poolowanna), Sturt-3
(Poolowanna) and Taloola-2 (Namur) oils, the Patchawarra Trough oils appear to have been
derived from source rocks deposited in environments which favoured pervasive bacterial
degradation of pre-existing prokaryotic and higher plant triterpenoids. The oils with very low
values (e.g. Tantanna-l, Poolowanna, 0.16; and Tanatanna-l, Hutton, 0.26) originated
from source beds in which the microbial degradation of triterpanes was extreme.
The sterane to bicyclic sesquiterpane ratio values range from 0.07 in the Tantanna-l
(Poolowanna) oil to 0.94 in the Sturt-3 (Poolowanna) oil. This range show that the
contribution of microbially reworked triterpenoids is greater than that of eukaryotic sterols.
Once again, the source rocks of some of the Poolowanna oils (Poolowanna-1, Sturt-3 and
Sturt-8: Table 7.1b) appear to have been subjected to a different type or lesser intensity of
microbial degradation whereas, at the other extreme, the Tantanna-l (Poolowanna) oil
records evidence of very strong bacterial reworking.
Poolowanna Trough.
2t9
Table 7 .2a Terpaneratios (mtz l9l) in selected rock extracts, Poolowanna Formation
27 Cr,cÞ Crrdþ
(m) Ts CrrTs Cro* Ts 225 225 pa Po Crotefia crùoP
(Ts+Tm) Crn+CrrTs CrrTs 225+22R225+22R d c[ CrrhoP
POL1-3 Poolowanna-l 2438 0.19 0.17 t.4l 0.09 0.58 0.53 0.20 0.15 0.24 2.t7
POL1-5 Poolowanna-I 2499 0.30 0.t7 t.29 0.20 o.54 0.57 0.r3 0.16 0.05 r.18
PO/-2-g Poolowanna-2 2524 0.52 0.41 1.16 0.39 0.51 0.52 o.2r 0.r7 0.34 1.85
POL3-13 Poolowanna-3 2505 0.32 o.20 2.43 0.2r 0.59 0.58 0.2r 0.19 o.43 1.43
TAN2-18 Tantanna-2 1799 0.08 0.13 0.75 0.06 0.63 0.48 0.25 0.24 0.18 1.52
TAN2-20 Tantanna-2 1811 0.16 0.15 0.56 0.13 0.64 0.67 0.22 0.30 0.20 t.07
TAN3-21 Tantanna-3 1807 0.08 0.13 o.42 0.07 0.57 0.59 0.20 0.29 0.15 1.08
TANS-25 Tantanna-8 18231 0.11 nd nd 0.09 0.59 0 58 0.15 0.23 0.16 o.76
STU1-27 Sturt-l 1871 0.10 nd nd 0.07 0.54 0 53 0.24 0.31 o.t4 0.97
STU4-30 Sturt-4 1859 0.05 nd nd 0.05 0.61 0 57 0.26 0.39 0.r7 0.83
STU4-31 Sturt-4 1881t 0.06 nd nd 0.06 0.59 0 59 0.29 o.45 0.12
0.r7
1.05
r.t6
STU6-35 Sturt-6 18651 0.06 nd nd 0.06 0.58 0 58 0.24 0.37
1.10
STEI-38 Str¡rt East-l 1841 0.r3 0.11 0.67 0.09 0.65 0 58 0.25 0.31 0.10
l3 t.r7
STE4-45 Sturt East-4 18 62+ 0.03 nd nd 0.03 0.59 0 58 0.36 0.46 0.
t coaly facies
Chapter 7
The Patchawaffa Trough samples have more consistent values, in the range 014.2 (average
In the Poolowanna Trough, this ratio increases from 0.16 at2411 m in Poolowanna-l to
O.52 at2524 min Poolowanna-2. This trend appears to be consistent with increasing thermal
maturation, although this ratio may also be affected by organic facies variation (McKirdy er
al., 1983, 1984). In the Patchawarra Trough, this ratio increases from 0.03 at 1862 m in
Sturt East-4 to 0. 16 at 18 1 1 m in Tantanna-2 whereby an increase of the ratio with increasing
maturity is apparentþ demonstrated. However, the coaly facies tends to have lower values
than the shaly facies (Fig. 7.4b). Comparison of the Poolowanna Formation in these two
troughs shows that the shaly facies in Poolowanna Trough is more mature, and thus
characterised by higher Ts/Ts+Tm ratios, than the early mature shales in the Patchawarra
Trough.
The Poolowanna Trough oil (Poolowanna-1, DST 2: Table 7.2b)has the TsÆs+Tm value of
0.44 . This value is within the range indicated for the Poolowanna Formation source rocks in
this area (0.16-{.52), consistent with a local origin for this oil.
221
oil
Poolowanna-1
DST 2
c¡o aÞ
Poolowanna
Rc =O.7l7o
czgo,þ
C29 Ts
C24tetra Tm
Ts
czzo,þ
C2s dia
Figure 7 . 4a Tnterpane mlz 191 signatures of the Poolowanna oil and selected
Po-olowanna sourcé and non-source facies in the Poolowanna Trough
oil
Tant¿nna-1 C¡o crÞ
DST 3
Birkhead/ Hutton
Rc = 0.487o
czsq,þ
Tm czzg,þ
Ts
C:o
c¡o*
Ts
C24tetsa
Non- source
Tantanna-8
1823 m
TOC: 5l.0Vo
Ro: 0.587o
C3o* = diahoPane
Figure 7.4b Comparison of triterpane mtz I9l signatures of a Jurassic oil and selected
soúrce and non-soirrce facies of Põolowanna Formation, Patchawarra Trough
Chøpter 7
In the Patchawarra Trough, the Jurassic oils have values ranging from 0.19 (Sturt-East-2,
DST 1) to 0.33 (Taloola-2, DST 1). These values are somewhat higher than those shown by
the Poolowanna source rock facies, an indication that the oils are unlikely to have originated
from these sections of the Poolowanna Formation sampled. However, these values may be
influenced by factors other than thermal maturation. The Permian and pre-Permian oils have
values similar to those of the Jurassic oils, suggesting either shaly source rocks of equivalent
organic facies and maturity or coaly source rocks of higher maturity.
The rearranged C2e hopane, 1Sa(H)-30-norneohopane (C2eTs: Fig. 2.8) was identified in all
the rock samples from the Poolowanna Trough but it is absent from the coal facies of the
Patchawarra Trough. Its abundance relative to C2el7u(H)-hopane (Czsap) is expressed
using the ratio C2sTslC2eaB+C2eTs. In the Poolowanna Trough, the values of this ratio fall
in the range 0.17-0.41 and apparentþ increase with thermal maturity. Like Ts [18c¡¿(H)-
trisnorneohopanel, its relative abundance is highest at 2524m in Poolowanna-2, being
influenced by both maturity and also the oxicity of the depositional environment (Peters and
Moldowan,1993).
Again, the Poolowanna Trough oil (Poolowanna-1, DST 2) has the highest
C2eTslC2saB+C2eTs value (0.25: Table 7.2b). The Jurassic oils in Patchawarra Trough
havevalues ranging from 0.12, Sturt 3 (Birkhead) to 0.19 Tantanna-l, (BirkheadÆIutton).
A similar range is observed for the Cambrian (Mooracoochie) and Permian oils. Therefore,
this parameter appear not to be suitable for distinguishing between Jurassic and
PermiarVpre-Permi an oils.
Another rearranged hopane, 17ct(H)-diahopane(C36*: Fig. 2.8) was identified in the source
rock extracts and crude oils. Its abundance relative to C2eTs is shown in Table 7.2a. The
Poolowanna Trough rock samples have higher values (1.16-2.43) than those from the
Patchawarra Trough (0.42-0.75). Both maturity and depositional environment have been
suggested to influence the C36*/C2eTs ratio (Kolaczkowska et al., 1990; Peters and
Moldowan , 1993). Thus, this parameter confirms that the Poolowanna Trough source rock
facies are relatively more mature than those in Patchawarra Trough, but also hints at possible
differences in their depositional environments.
The oil in the Poolowanna Trough has higher C36*/C2eTs value (0.72) than do the
Patchawarra Trough oils (0.82-1.33), without any clear distinction between Jurassic and
Permian/pre-Permian oil s.
224
Table 7.2b Terpane ratios (mtz l9l) in crude oils from the Poolowanna and Patchawarra Troughs
I 1800-1814 Poolowanna o.32 0.17 1.18 0.16 0.54 0 51 o.r4 0.15 0.18 r.66
T2 Tantanna-1
T3 Tantanna-1 2 t635-1642 Hutton 0.32 0.17 1.00 0.r7 0.62 0 60 o.l2 0.15 0.25 r.52
Tantanna-1 3 1347-1359 Birk/Ilutton 0.27 0.19 0.82 0.13 o.57 0 58 0.27 0.26 0.14 t.70
T4
1A Poolowanna 0.20 0.16 0.90 0.12 0.57 0 56 0.22 0.24 0.15 r.44
sl1 Sturt-3 1848-1855
0.20 0.t2 0.19 t.52
s12 Sturt-3 18 1654-1662 BirkÍread o.25 0.t2 1.33 0.13 0.53 0 60
s13 'Shrrt-4 I 1872-1880 Poolowanna 0.23 nd nd o.l2 0.59 0.61 0.18 0.r7 0.16 1.55
s16 Sturt-6 6 1914-Igl9Mooracoochie 0.26 nd nd 0.15 0.59 0.61 0.11 0.15 0.22
0.18
1.49
t.67
s17 Sturt-6 3 1884-1898 Patchawarra 0.31 0.15 t.20 0.16 0.58 0.59 0.13
l5
0.18
I.r9
s18 Sturt-6 1 1883-1892 Birkhead o.24 0.13 0.96 0.16 0.58 0.55 0. o.20 0.26
s19 Sturt-7 I L923-1938 Patchawarra 0.20 nd nd 0.t4 0.58 0.59 0.19 0.18 0.16
o.t7
1.59
s20 Sturt-7 2 1946-2021 Mooracoochie 0.22 0.t4 0.80 0.13 0.59 0.60 o.22 0.26 1.48
s21 Stuf-7 3 1871-1876 Poolowanna o.2r 0. r3 1.50 0. l3 0.58 0.54 0.21 0.20 0.18 1.64
s23 Sturt-8 1 1880-1884 Poolowanna 0.2L nd nd 0.09 0.59 0.57 0.24 0.22 0.10 1.98
SE24 Shrrt East 2 I 1845-1899 Poolowanna 0.19 0.15 1 08 0.10 0.59 0.61 0.16 0.21 0.22 1.46
t.70
TL25 Taloola-2 1 1829-1836 Poolowanna 0.33 0.18 I 25 0.14 0.61 0.60 0.12
0.t4
0.16 0.20
0.t7
TL26 Taloola-2 2 1789-1793 Hutton 0.31 0.18 1 t9 0.14 0.62 0.59 0.19 1.58
r.63
TLN Taloola-2 3 1384-1397 Namur 0.32 0.18 1 04 o.t2 0.60 o.54 0.15 0.18 0.11
Chnpter 7
High ratio values (i.e. high C¡o*), apart from being indicative of terrestrial organic matter,
have also been suggested to be related to bacterial hopanoid precursors that have undergone
oxidation in the D-ring and rearrangement by clay-mediated acidic catalysis (Peters and
Moldowan, 1993).
identified in both source rock extracts and crude oils. Their concentrations relative to their
corresponding C2ecrB- and C36oB-hopanes are shown in Table 7.2a for source rock extracts
of the C2epu/crB ratio in the
and Table 7.2b for oils. Whilst there is no significant variation
source rock extracts, appreciable variation is shown by C36Bo/crp. The latter ratio in the
Poolowanna Trough source rock facies has values of 0.15-0.19 whereas in the Patchawarra
Trough the values are in the range 0.23-0.46. During catagenesis the moretanes decrease
rapidly in relation to the crB-hopanes (ten Haven et al., 1992). Therefore, the lower
moretane/hopane ratios of the Poolowanna Formation in the Poolowanna Trough are further
evidence of its more advanced level of catagenesis.
The moretane/tropane (C3sBa/o(B) for the oils have a range of 0.12-0.26. Assuming that
of their source rocks at the time of expulsion, it appears that
these values are close to those
the Poolowanna Formation source rocks are most likely the source for the Poolowanna-l oil.
Also, in the Patchawarra Trough, the Poolowanna Formation source rocks analysed seem to
be not quite mature enough to generate the oils found in this trough. The Sturt-7
(Mooracoochie) oil and Tantanna-l (BirkheadlHutton) (with the C36Bcr/crB value of 0.26)
appeff to have been generated at an earlier maturity stage than the other Permian and Pre-
Permian oils.
226
Reservoir
Sturt-6
DST 1 c¡o oÞ Birkhead
Ts/(Ts+Tm):0.24 Middle Jurassic
C zz aþ 225I (225+22R) : 0.5 5 czguþ
Tm
c¡z oF
C2 tetracyclic Ts c¡oÞc
R
Sturt-7
DST 3 Poolowanna
Ts/(Ts+Tm):0.21 Early Jura.ssic
Czz aþ 225 I (225+22R) : 0.54
Sturt-6
DST 3 Patchawarra
Ts/(Ts+Tm):0.31 Early Permian
Czz aþ 225 I (225+22R) : 0.59
Sturt-7
DST 2 Mooracoochie
Ts/(Ts+Tm):0.22 Early Cambrian
czzg,þ 22Sl(225+22R): 0.60
FigureT. 4c Triterpane mtz I9I signatures of oils in stacked reservoirs of the Sturt field,
Patchawarra Trough
Table 7.3a Distribution of methylhopanes (r/z 205) in selected rock extracts, Poolowanna Formation.
crg Cro
No. (m) 2þ 2a 3Þ 3þt2a 2þ 2a 3þ* 2u (þo) 3þt2u 2a 3p mk 3þtza klm
Vo Vo Vo Vo Vo Vo Vo Vo Vo Vo Vo
3.8 9.8 8.4 0.85 7.7 12.9 7.r 0.0 0.55
POL1-3 Poolowanna-l 2438 0.0 9.3 tr.7 r.26 9.7 12.5 7.3 0.0 0.59 18.6 8.6 12.3 9.9 o.47 0.81
poll-s pegls¡yanna-l 2499 0.0 0.0 10.1 nd il.6 13.1 10.1 0.0 0.77 19.1 10.8 15.8 9.5 o.57 0.60
POL2-9 Poolowanna-2 2524 0.0 0.0 26.9 nd 0.0 18.6 0.0 0.0 0.00 20.8 10.5 11.6 11.6 0.51 1.00
POL3-13 Poolowanna-3 2505 4.7 8.5 17.l 2.W 0.0 15.5 7.0 0.0 0.45 19.4 8.5 11.6 7.8 0.44 0.67
TAN2-18 Tantanna-2 1799 4.8 15.3 6.3 0.4L t4.4 18.1 9.2 3.5 0.51 t9.7 8.7 0.0 0.0 o.44 nd
TAN2-20Tantanna-2 1811 0.0 14.4 7.2 0.50 0.0 t6.7 11.7 13.5 0.70 24.3 12.2 0.0 0.0 0.50 nd
TAN3-21Tantanna-3 1807 0.0 13.3 t0.2 0.77 0.0 24.0 14.6 3.3 0.61 25.4 9.2 0.0 0.0 0.36 nd
TANS-25 Tantanna-8 18231 6.4 rt.s 7.0 0.61 0.0 25.5 8.3 7.3 0.33 24.2 3.2 3.5 3.2 0.13 0.91
STU1-27 Sturt-l I87I 6.3 11.3 9.9 0.88 0.0 22.5 10.6 4.9 o.47 26.t 8.5 0.0 0.0 0.32 nd
STU4-30 Sturt-4 1859 6.0 r7.5 11.1 0.63 0.0 20.3 12.4 5.1 0.61 18.3 9.4 0.0 0.0 0.51 nd
STU4-31 Sturt-4 1881f 5.1 r8.2 6.5 0.36 0.0 18.5 9.5 4.0 0.51 29.8 6.5 0.7 1.1 0.22 1.50
STU6-35 Sturt-6 18657 5.5 19.9 6.2 0.31 0.0 15.8 8.2 4.1 0.52 34.7 5.5 0.0 0.0 0.16 nd
STE1-38 Sturt East-l 1841 5.3 9.1 7.6 0.84 15.0 13.8 10.6 4.t 0.77 23.5 10.9 0.0 0.0 0.46 nd
STB+-45 Sturt East-4 18621 4.8 19.4 7.7 0.40 8.7 15.5 11.0 5.5 0.71 20.3 7.1 0.0 0.0 0.35 nd
t coaly facies
ChapterT
The C3dC2ecrB values for the oils vary from 1.35 in the Poolowanna-l (Poolowanna) crude
to 1.98 in the Sturt-8 (Poolowanna) oil of Patchawarra Trough. By comparing these oil
ratios to those of the source rock extracts, most of the Poolowanna Formation extracts from
the Patchawa:ra Trough can be shown to be non-sources; and most of those from the
Poolowanna Trough to be potential sources. Only the Sturt-6 (Birkhead) oil, (with a value of
1.19) could be generated from a local Poolowanna Formation source rock facies. A general
overview of the mlz l9l mass fragmentograms of the oils from the Sturt field look very
similar (Fig.7 .4c). However, closer inspection of their various terpane ratios reveal subtle
differences that show these oils in multiple stacked reservoirs of the same field do not have a
common origin.
The (C2e-C31) 2o-compounds appear to be ubiquitous. The 2cr-isomer is the most abundant
of the C31 metþlhopanes, with a relative concentration of 18.6-2l.3Vo in the Poolowanna
Trough and 18.3-34.77o in the Patchawarra Trough. The 2cr-isomer is also the most
abundant of the C2s and C36 methylhopanes, except in the Poolowanna Trough. A C3g 2u-
metþlmoretane was tentatively identified in the Poolowanna Formation from the
Patchawarra Trough, but not in the more mature samples from the Poolowanna Trough.
The C2e-C31 3p-metþlyhopanes were also identified and no significant variation in their
distribution was noted between the two troughs. The 3þl2u ratio is <1 in all cases except for
the C2s compounds in two samples from the Poolowanna.Trough. This may be partly due to
increasing maturity, although this ratio is also influenced by organic facies. The C31 3þ/2u
ratio is relatively low (0.13-0.22) in the coal facies of the Patchawarra Trough.
229
Poolowanna Trough c¡t oil:
Poolowanna-l, DST 2
2a Early Jurassic
c¡o \
c¡o C3y3þl2u:0.27
2a C:r
tÞt 3Þ(Me
3Þ (Me)
2o (Fo)
m Source rock
Poolowanna-1
\ k
2499 m
C3¡ 3þl2u:0.57
Sturt East-1
1841 m
C3y 3þl2a:0.46
czg
czs 3p (Me)
2þ 2a
Tantanna-8
1823m
C3¡3Bl2u:0.13
Poolowanna
Early Jurassic,
czg Sturt-7, DST 3
C3¡ 3þl2u: O.ll
2a
2p
Mooracoochie
Early Cambrian,
Sturt-7, DST 2
C3y 3þl2u:0.46
m k
Figure 7. 5b Metþlhopane mlz 205 signatures of oils in stacked reservoirs of the Sturt
Field, Patchawarra Trough
Tabte 7.3b Dstribution of methylhopanes (m/z 205) in crude oils from the Poolowanna and Patchawarra Troughs
crn cro 3t
Two unidentified C31 methylhopanes were noted in the mlz 2O5 chromatograms (Fig. 7.5).
One compound (m) elutes just before and the other (k) just after the C31 2cr-metþlhopane.
These compounds are quite cofirmon in the Poolowanna Trough, but very rare in the
Patchawarra Trough where they occur only in two coal samples. The k/m ratio ranges from
0.60 to 1.00 in the Poolowanna Trough and 0.91 and 1.5 in the Patchawarra Trough.
with increasing carbon number. The C3lmethylhopanes have low 3þl2u values which range
from 0.10 in the Tantanna-l (Hutton) oil to 0.46 in the Sturt-7 (Mooracoochie) oil. The 3B-
compounds seem to be of lower concentration in the oils than in the source rock extracts.
Different values of this ratio may reflect variations in the bacterial input to the source rock.
The unidentified C31 methylhopanes (peaks k and m) noted in some of the source rock
extracts are surprisingly present in all the oil samples. The k/m ratio varies from 0.62 in the
Tantanna-l (DSTl) to2.25 in the Sturt-3 (DST 1A) oil, both from Poolowanna reservoirs.
However, no consistency is evident in this ratio. Nevertheless, the presence of these
compounds may be applied in oil-source correlation. The source rocks containing these
compounds are most likely to have generated these oils. As indicated by other parametets,
most of the Poolowanna Trough source rock facies identified in this study are likely to be
effective sources of Jurassic oil (see below).
233
oil
Poolowanna-1
DST 2, Poolowanna
Early Jurassic
Czs ÞF/(ÞÞ+oa):0.55
(R
czl cza S
q,C[ R
Non-source facies
czg Poolowanna-1
13p, I /ü-Olasrcranes f r 2417 m
TOC:16.3 Vo
Czs ÞÞ/(ÞÞ+oo): 0.43
czg
czl
Figure 7. 6a Steran e mlz 217 signatures of the Poolwanna oil and selected source and
non-source facies in the Poolowanna Trough
Table 7.4a Relative distribution of steranes and C, isomerisation ratios in PoolowannaFormation source rocks
t coaly facies
Slggr" =2 Crn sterane
Hopane C, op-hopane
Bo (20R) =Czg Dtast"t*e
ocr (20R) sterane
The various isomer ratios of 24-ethylcholestane are given in Table 7.4a. These three ratios
were in order to assess the influence of source rock facies on sterane isomerisation.
Isomerisation at the C-14 and C-17 positions of the C2e sterane, represented by the
14P(H),17P(H)/(crBB+üo(o) ratio, has values ranging from 0.43 to 0.56 in the Poolowanna
Trough. An increase of the ratio with depth was noted in Poolowanna-1, reflecting the
previously noted increase in thermal maturity. In the Patchawarra Trough, values range from
0.33 to 0.43. The rocks are indicated to be less mature than those in the Poolowanna
Trough. This ratio seems to be independent of the TOC content of the source rock (Fig.
1.7).
The crBp(20R)/qacr(20R) ratio varies from 0.79 to 1.20 in the Poolowanna Trough and
from 0.34 to 0.61 in the Patchawarra Trough, highlighting in a similar fashion to the
øBB/Bp+ucrcr ratio the lower maturity of the latter samples'
236
czg
aaR
qaS
R
Tantanna-2
cza
czt 1799m
TOC:78.17o
\ Czs ÞÊ/(ÞÞ+c¡cr):0.36
Tantanna-8
1823 m
TOC:5lVo
Czs FÞ(ÞÞ+oo):0.39
Sturt-4
1859 m
TOC:2I.4Vo
czs FÞ/(ÞÞ+ca):0.33
,'1, ,\-J¡*--¡^-rt-
czs
13p, l7c-diasteranes Sturt East-1
1841 m
TOC:7.\Vo
2 Czs ÞF/$F+oo,): 0.34
czt
Figure 7. 6b Sterane mlz 2L7 signatures of potential source facies in the Poolowanna
Formati on, Patchawarra Trough
ChapterT
A wider range of values, from 0.28 in carbonaceous shale (Sturt-4, 1859 m) to I.52 in
siltstone (Tantanna-z,1811 m), is observed in the Patchawarra Trough. Since there is very
little difference in maturity between these samples, this wider range presumably reflects
variation in their lithological character and depositional setting.
As reported by McKirdy et aI. (1953) and Rullkötter et a/. (1985) low diasterane/sterane
values are indicative of anoxic, clay-poor depositional conditions. Thus, the anomalously
high value (1.52) in the Tantanna-2 (1S11 m) sample may signify a higher clay content. The
same argument may be applied to the less mature Poolowanna-l (2417 m) sample in the
Poolowanna Trough which has a much lower value than other samples in that trough. Its
low value may also partly reflect less oxic conditions in the upper part of the formation. The
coals are also characterised by a wide range of diasterane/sterane values (0.36-0.89). Like
'other
samples from the Patchawarra Trough, the coals do not differ very much in maturity,
and therefore the variation in their diasterane/sterane ratio reflects the different surface
catalytic properties of their kerogen matrix and/or differences in their mineral matter content.
The influx of mineral matter, including clays, may explain the higher diasterane content of
the Tantanna-8 (1823 m) coal. This is consistent with its relatively low TOC content (50.IVo:
Table 5.1). In both the Poolowanna and Patchawarra depocentres, the Poolowanna
Formation source rocks are indicated by such variations to be of paludal to fluvial-lacustrine
origin.
The abundance of steranes relative to hopanes was evaluated using the ICzs sterane/C36crB-
hopane ratio. This ratio is used to compare the input of eukaryotic biota (i.e. algae and higher
plants) to that of prokaryotic organisms (bacteria). In the Poolowanna Formation from the
Poolowanna Trough the values are mostly in the range 0.18-0.44, indicating high bacterial
inputs. The unusually high ratio (1.7) observed at2499 m in Poolowanna-l may indicate
diminished bacterial reworking of the higher plant contribution to its preserved organic
matter.
In samples from the Patchawarra Trough the sterane/hopane values range from 0.07 to 0.45.
Coaly facies are shown to have relatively high values, consistent with their high terrestrial
organic matter contents. Very low values (0.07) shown by the Tantanna-2 (1799 m) and
Sturt East-l (1841 m) shales indicate higher pH and stronger microbial activity in their sub-
aquatic depositional environments. The strong bacterial influence indicated by this parameter
supports earlier observations based on the n-alkane distributions of these samples (Table
6.3a).
238
Table 7.4b Relative distribution of steranes and C, isomerisation ratios of oils in the Poolowanna and Patchawarra Troughs
Key: as forTabteT.4a
ChapterT
As in the source rock extracts, the C2e homologue is the most abundant sterane, ranging
from 3'7Vo in the Tantanna-1 (Hutton) oil to 53Vo in the Poolowanna-l (Poolowanna)
sample. The C27 sterane appears to be second in abundance (average = 29Vo), whereas the
C* sterane is the least abundant (average = 26Vo).However, the C27 steranes are subject to
interference by co-elutingC2e diasteranes (Fig. 7.6).
The 2OS/20S+20R isomerisation ratio has values ranging from 0.44 in the Sturt-8
(Poolowanna) oil to 0.53 in the Tantanna-l (Poolowanna) sample. Like the previous
isomerisation ratio, it does not differ much among these oils. For instance, in oils from
stacked reservoirs at any one location the difference in this parameter is between 0.01 and
0.06. Thus, the implication is that these oils in stacked reservoirs have roughly equal
maturation levels. However, the average 2OS|2OS+20R values in the Poolowanna Trough
(O.52) and Patchawarra Trough (0.50) source rocks are slightly higher than those of the
adjacent oils.
The diasterane/sterane values in the Poolowanna and Patchawarra Trough oils (Table 7.4b)
suggest highly mature oils (Peters et al., 1990). They range from 1.06 in the Sturt-7
(Mooracoochie) oil to 2.55 in the Tantanna-l (Hutton) oil. This is in contrast to the
Poolowanna source rock extracts which in all but one case have values <1 (Table 7 .4a). The
higher values in the oils may be attributed to a geochromatographic effect during primary
migration. Peters et al. (1990) reported increased diasterane/sterane ratios in expelled oils,
compared to the source rock bitumen, during hydrous pyrolysis experiments.
Oils from Permian and Cambrian reservoirs tend to have lower diasterane/sterane ratios
(1.06-1.68; average = 1.38) than those from Jurassic reservoirs (1.13-2.55; average =
1.70) in the Patchawarra Trough. Given that the Permian and Cambrian oils are more mature
than the Jurassic reservoired oils (Chapter 6), it can be asserted that the source rocks for
most of the Jurassic oils were richer in clays than those of the former oils. The Poolowanna-
1 (Poolowanna) oil from the Poolowanna Trough has a diasterane/sterane ratio of 1.31
Hopanes appear to be more abundant than steranes in all but two of the oils. This is shown
by sterane/hopane ratios mostly in the range 0.03-0.39 (Table 7.4b).
240
czg Birkhead
Middle Jurassic
czs Sturt-6, DST 1
BP @+s) Czs ÊF/GF+cra):0.59
Czt
acrS acrR
I
Poolowanna
Early Jurassic
Sturt-7, DST 3
czs FÊ/(ÞÞ+aø):0.61
\-*tàÉl,'lrÐfr¡4J
Patchawarra
Early Permian
Sturt-6, DST 3
czs FÞ/(ÞÞ+acr):0.55
--r a++r$,**s\,q
czg
Figure 7. 6c Sterane mlz 2I7 signatures of oils in stacked reservoirs of the Sturt Field,
Patchawarra Trough
ChnpterT
From this ratio the Taloola-2 (Namur) and Tantanna-l (BirkheadlHutton) oils appear to have
the strongest contribution from bacterial lipids. The Tantanna-l (Hutton) and Sturt-8
(Poolowanna) oils show anomalously high sterane/hopane values indicative of minimal
bacterial reworhng of the organic matter in their respective source rocks.
7.3.1.1 n-Alkanes
As discussed previously (Sections 6.3 and 7.2.1) the n-alkane profiles of source rock
extracts represent multi-component mixtures from different source biota. Normal alkanes
between C15 and C1e are the most prevalent and are usually assigned to algae deposited in
either marine or in this case lacustrine environments. A strong bacterial contribution is also
indicatedby the abundance of n-alkanes in the Czo-Czø range, whereas land plant inputs
show up in the Czt-Cy range (Table 6.3a).
CPI and OEP values are used not only as maturity indicators but also for organic matter input
verification. High CPI values (above 1.5) always signify immature source rocks. Low CPI
values, however, do not necessarily mean higher maturity. They can also imply a lack of
Cya n-alkanes stemming from tenestrial inputs. The Poolowanna Formation source rocks
have CPI values between I.02 and 1.39, which is indicative of them being mature or instead
having strong inputs of planktonic algae and/or bacteria that dilute the higher land plant
contribution.
The bimodal profiles, comprising both medium and high molecular weight n-alkanes in the
Patchawarra Trough rock extracts imply contributions from a wider range of biota. A
maximum atn-C25is frequently shown in the higher molecular weight range of the bimodal
profiles (Fig. 6.1 and 7.1). According to CPI, measured vitrinite reflectance and other data,
it seems that most of the bimodal and unimodal profiles with a maximum at n-C25 ma!
indicate bacterial rather than higher land plant influence. Coincidentally, such samples (Fig.
6.1) also show strong contributions from bacteria based on the relative abundance of z¿-
242
ChapterT
bimodal profiles indicates a fluvial depositional setting wherein various types of organic
matter are inter-mixed.
Terrestrial organic matter deposited under suboxic to oxic conditions is therefore inferred for
the Poolowanna Formation source rock since its prþh ratios are >2. The source rocks in the
Patchawarra Trough seem to have experienced a wider range of oxicity (prlph = 2.3 to 8.4)
than those in Poolowanna Trough (prlph =3.2 to 6.6). A strong terrestrial organic matter
input is further demonstrated in the plot of prln-Cs versus ph/n-C13 (Fig. 7.8). The
Poolowanna Trough source rocks are shown to have been deposited in less oxidising
conditions and to be of higher maturity than those in the Patchawarra Trough. However,
by other
since there is a possibility of a strong bacterial contribution (as previously indicated
parameters), the prþh ratios in this case are likely to be influenced by pristane derived from
bacterial lipids (Volkman and Maxwell, 1986).
7.3.2.1 Terpanes
Several homologous series of terpanes including bicyclic (drimane), tricyclic, tetracyclic and
pentacyclic compounds found in source rocks and crude oil are believed to originate from
bacterial membrane lipids (Ourisson et al., 1982; Alexander et a1.,1983; Volkman, 1988).
The presence of these compounds in the Poolowanna Formation source rocks is discussed in
previous sections of this chapter. Though not very specific, their presence is indicative of
bacterial inputs. The heterogeneity of depositional conditions between source rocks of the
Poolowanna Trough and those of the Patchawarra Trough is well illustrated by non-uniform
ap-hopaneibicyclic sesquiterpane ratios (Table LIa). The variation in depositional
conditions is further demonstrated by the unusually low C3çlC2e hopane ratios (<1) in some
sections, and by the absence of the unidentified methylhopanes peaks (k and m) in the
243
1.00
0.75
o
t)
É
cl
ñl
È 0.50
>ì
Ð
l-i
(É
0.25
0.00
0 20 40 60 80
ToTOC
Figure 7.7 Compaison of C29 sterane isomerisation ratios with non-biomarker maturity
paiameters, and their relationship to organic facies based on TOC
ChapterT
7.3.2.2 Steranes
The presence of steranes in the source rocks reflects inputs of eukaryotic organisms mainly
higher plants and algae. Therefore, the abundance of steranes relative to hopanes indirectþ
reflects the proportional inputs of eukaryotic and prokaryotic organisms. The C2e sterane/C3¡
hopane ratios (Table 7.4a) show that the input of prokaryotic organic matter was
predominant over that of eukaryotic origin. A similar conclusion is suggested by the
steranes/drimane ratios (Table 7. la)
The relative abundance of C27, C2s and C2e steranes plotted in a ternary diagram (Huang and
Meinschein, 1979) was used to categorise different depositional environments and the type
of contributing organisms. A mixed depositional regime of lacustrine and terrestrial
environments appears to have been prevalent during the deposition of the Poolowanna
Formation source rocks (Fig. 7.9). Neither zooplankton nor phytoplankton are indicated to
be significant contributors to the source biota, but an enhanced contribution by higher plants
is shown for coals from the Sturt and Sturt East Fields. Otherwise a mixed non-marine
depositional environment (i.e. lacustrine to meandering fluvial and paludal) is indicated.
7.3.3.1 n-Alkanes
The n-alkane range in the crude oils suggests a strong contribution from terrestrial organic
matter precursors, particularly cuticular and leaf waxes which are known to have a carbon
range number up to n-C3 5 or n-C37 with a maximum between n-Czt and n-Cy (Eglinton er
at.,1962). However, the credibility of a strong terrestrial input is doubtful due to the lack of
mærima in the higher plant range. Most of these oils have a maximum between n-Cp and
n-Ct¿. It is likely that the original maxima have been altered by migration processes or
thermal cracking. There is no evidence of microbial activity (i.e. biodegradation) within the
reservoirs of the Poolowanna and Patchawarra Troughs. Nevertheless, terrestrial and
bacterial inputs are implicated in Poolowanna Trough oil by its n-alkane range of Ce-C3e and
maximum atCy.
The CPI values of the oils (0.79-1.25) appear to be within the normal range for mature
source rocks and oils. Values >1.1 possibly signify terrestrial plant inputs to the source rock
of the oil. No significant differences are evident between the Jurassic, Permian and
Cambrian oils as far as their CPI and n-alkane ranges are concerned. Therefore these
parameters can not be used positively to identify oil source affimty in the study area,
although the bacterial affînity of the oils is clearly indicated by their high relative abundance
of C2ç-C26 n-alkanes.
245
1.0
o
OI
L o
È
IL
0.1
o Poolowanna
r Tantanna
O Sturt
o Sturt East
0.01
0.01 0.1 1.0
Ph/n-C18
Environment
Fields
o Poolowanna
,rlllll
À Tantanna
o Sturt
tr Sturt East
Higher
'*otîiÌiifijl plant
czt czg
Figure 7.9 Source biota and depositional environments of Poolowanna Formation source
roõks based on regular steranes. Ternary plot modified from Huang and Meinschein
(re7e)
Chøpter7
7.3.4.1 Terpanes
Bacterial inputs to the oils are demonstrated by a number of terpanes series. The drimanes
are quite abundant in many oils. The hopane (C¡ocrþ) to bicyclic sesquiterpane ratio has
valuesrangingfrom0.t6to3.46 (Table7.1b) indicating large differences in the types of
bacterial input. Different values may indicate different biota or depositional conditions.
Based on this ratio it can be seen in several fields that the oils in multþle stacked reservoirs
do not have common source (Table 7.Ib). For instance, in Tantanna-l well, the Poolowanna
oil has a very much higher abundance of bicyclic sesquiterpanes than the BirkheadlHutton
oil; and the Namur oil seems to have a very different source from the rest of the oils in
Taloola-2. The Sturt-7 (Mooracoochie) sample is unique among this collection of oils
because its source rock appears to have been deposited under highly reducing conditions
(prlph = I.2)
7.3.4.2 Steranes
A derivation from source rocks deposited in terrestrial, lacustrine or estuarine environments
is indicated by the relative abundances of C27, C2s and C2e steranes in the oils (Fig. 7 .ll).
None of major biota (higher plant, zooplankton and phytoplankton) is indicated to be
dominant.
247
1 0
C)I
c
O
fL
0.1
0.01
0.01 0.1 1.0
Ph / n-C1g
Reservoir
a Namur x Patchawarra (Permian)
a Birkhead r Mooracoochie
Volcanics (Cambrian)
r Hutton
. Poolowanna
Figure 7.10 Oil source affinity, maturity and biodegradation based on isoprenoid to
n-alkane ratios, Poolowanna and PatchawÍura Troughs.
Depositional Environment
m open marine
Lacustrine
-lÍÍilii'td Estuarine/bay
N Terrestrial
czz czg
Figure 7.11 Oil source affinity based on the distribution of regular steranes, Poolowanna
anã Patchawarra Troughs. Ternary plot modified from Huang and Meinschein (1979).
(Key as in Figure 7.10)
ChapterT
However, a mixture of higher plants and phytoplankton are the most likely eukaryotic
precursors. It is interesting that the Sturt-7 (Mooracoochie) marine oil plots within the
compositional field of other non-marine oils.
7.4.2.1 Terpanes
The thermal maturity applications of the triterpane biomarkers are based on the fact that the
biologically produced hopane precursors catry a 22R configuration which is gradually
converted to a mixture of 22F. and 225 diastereomers. Thus, under normal circumstances,
the proportions of 22R to 225 in the C31 to C35 homologues will give a relative thermal
maturity of the source rock.
The 225/(225+22R) ratios for the C31 and C32 l7c-homohopanes derived from the
Poolowanna Formation source rock extracts are shown in Tabke 7.2a. Values in the
Poolowanna and Patchawarra Troughs are in the range 0.51-0.59 and 0.54-0.65,
249
ChapterT
respectively for the C3l hopanes (both with an average value of 0.56 that is indicative of the
early mature stage). However, since the equilibrium value of 225/(225+22R) is 0.60, the
indicated maturity level is under estimated, particularly for the Poolowanna Trough source
rocks which are indicated by vitrinite reflectance data to be mature. On the other hand, the
maturity indicated for the Poolowanna Formation in the Patchawarra Trough agrees with
vitrinite refl ectance data.
7.4.2.2 Steranes
Thermally driven reactions undergone by steranes include equilibration between the
biological epimer 20R and the geological epimer 20S 5cr(H),14cr(H),17ct(H) (Section
2.6.9) which is expressed in the 20S/(20S+20R) ratio. Isomerisation at C-20 in the C2e
5c(H),14cr(H),17cr(H)-sterane causes this ratio to rise from 0 to about 0.5 with increasing
maturity (Seifert and Moldowan, 1986). The source rocks in the Poolowanna Trough have
values ranging from 0.49 to 0.58 (Table 7.4a), which are very similar to those of the
Patchawarra Trough source rocks (0.48-0.56). These values are indicative of early mature to
mature source rocks.
Isomerisation at the C-14 and C-17 positions in the 20S and 20R C2e sterane, causes an
increase in the c¡¿ÞÊ/(crÞÞ+acrcr) ratio from near-zero values in immature sediments to about
0.7 which is equivalent to peak oil generation. The Poolowanna Formation source rocks in
the Poolowanna Trough have values ranging from 0.43 to 0.56 (Table 7.4a) and are thus
indicated to be within the oil window. In the Patchawarra Trough, the source rocks have
values ranging from 0.33 to 0.43 and are thus again shown to be less mature than those in
Poolowanna Trough.
250
1.75
o Poolowanna O Sturt
1.25
Øl É.
olo
C\I I ôI
1.00
ËlËro
tÕl
0.75
0.50
0.25 0.50 0.75 1.00 1.25
14þ,17p (20R)
5c (20R)
Key:
First Order Kinetic Conversion (FOKC)
ffi Shale/
Ca¡bonaceous
(Seifert and Moldowan, 1981)
The shaly-carbonaceous clastic source rocks from the Poolowanna Trough plot along a
separate trend from that of the coaly Patchawarra Trough samples. This has important
implications for oil - source correlations (see Section 7.4.4)'
Theprln-C17 and pWn-Cß values of most of the oils (Fig. 7.10) overlap those of the
Poolowanna Formation source rocks (Fig. 7.8). Suggesting that they are of similar maturity.
The one exception is the marine oil from the Mooracoochie Volcanics in Sturt-7. This oil
appears to be appreciably less mature than the other oils, although its isoprenoid/n-alkane
ratios may have been enhanced by an early phase of biodegradation when the Carnbrian
reservoir was invaded by meteoric water.
The C2e moretane/hopane ratios have values ranging from 0.09 to 0.24 for Jurassic oils and
from0.l I to0.22 forPermian and Carnbrian oils. The conesponding ranges for the C3s
isomers are 0.12 to 0.26 and 0.15 to 0.26, respectively. Generally in stacked reservoirs,
oils from the younger reservoirs have higher values than those from older reservoirs (e.g. at
Sturt-6 and Taloola-2). Thus, the latter oils are indicated to be more mature than the former.
TsÆs+Tm values appear to be too inconsistent to be employed for maturation assessment.
This inconsistency is attributed to variations in organic facies.
252
1.50
o
x
É.
o
ôl a
Ë o
e
U)
1 00
A
t
o
ôt a
Ë
lr) o .tl o
a
x
0.50
0.50 1.00 1.50 2.OO
Reservoir
a Namur X Patchawarra(Permian)
A Birkhead
x Mooracoochie
¡ Hutton Volcanics (Cambrian)
O Poolowanna
Sturt-6 x Sturt-3 A
1.1 0.9
o
1 0.8
0.9 o.7
0.8 0.6
x
o.7 0.5
1.1 1.15 1.2 1.25 1.3 1.35 1.4 50. 5 1.25 1.5
É.
o 1 1
ôt Sturt- 7 x Taloola-2 o
ð
ìa¡
U)
o
c\¡ 0.95 1 1.,
ð
rñ
0.9 0.95
a
0.9
0.9 1 1.1 1.2 1.3 1.4 1.5 1.25 1 .3 .35 1.4 .45
1 1 1 .5
t4p,17þ (20R)/sa (20R)
1.1
Tantanna-1 o
1.05 Reservoir
1
a Namur
Birkhead
T Hutton
0.95
o Poolowanna
x Patchawarra (Permian)
0.9 Mooracoochie
L/ x Volcanics (Cambrian)
0.85
.9 1 1 .2 .3 1.4
FOKC - curve
Translated FOKC
The oils are indicated to be early mature to mature oils by their C2e sterane 20S/(20S+20R)
and ocBB/crpB+crocr isomerisation ratios. The 20S/(20S+20R) values for Jurassic oils range
from 0.44 to 0.51; for Permian oils from 0.47 to 0.51; and for Carnbrian oils from 0.48 to
0.49. The corresponding oBp/oBB+oaa ratios arc 0.49 to 0.61 for Jurassic oils; 0.55 to
0.56 for Permian oils; and 0.51 to 0.57 for Cambrian oils. These sterane maturity parameters
fail to support the impression of significantly lower maturity indicated for the Stun-7
(Mooracoochie) oil by its isoprenoid/n-alkane ratios (Figs. 7.10 and 7 .I3a). However, they
do show that in the Patchawa:ra Trough the Permian oils a¡e more mature than the Carnbrian
oils and most of the Jurassic oils.
In Sturt-6, the Mooracoochie and Birkhead oils plot close to the FOKC line (and thus may
have the coÍtmon origin); but are clearly different from the Patchawarra oil. In Sturt-7,
however, the Mooracoochie, Patchawarra and Poolowanna oils all plot on different
maturation trends and are genetically unrelated. Finally, comparison of Figures 7.I2 and
7.13 rcveals that none of the oils originated from the coal facies of the Poolowanna
Formation.
255
Chapter I
Chapter I
S. L Introduction
Aromatic compounds are major constituents of crude oils and source rock extracts (Radke,
1987). These compounds include aromatic hydrocarbons of the benzene, naphthalene,
phenanthrene and anthracene, pyrene, benzanthracene and chrysene type (Section 2.1);
naphthenoaromatic hydrocarbons; and aromatic sulphur compounds, of which
benzothiophene derivatives are the most coÍtmon.
Their use as biomarkers has been based on their structural similarities to certain naturally
occurring biogenic compounds, as first noticed by Mair and Martinez-Pico (1962). As an
example, laboratory experiments by Ruzicka and Hosking (1930) showed that aromatisation
of the bicyclic resin acids and alcohols produced agathalene (1,2,5-trimethylnaphthalene)
whereas the tricyclic components gave rise to pimanthrene (1,7-dimethylphenanthrene) and
retene (Fig. 1.7). The presence of polycyclic aromatic hydrocarbons (PAH) in petroleum are
thought to be products of complex transformations of naphthenic and olefinic biological
precursors, because they are not synthesised in significant quantities by living organisms
(Grimmer and Düvel, 1970; Grimmer et aI.,1972; Hase and Hites, 1976).
258
TMN
1
DMN
r,6
1,3+1,7
1,4+2,3
2,6+2,7
1,4,6+I,3,5
1,2
2,3,6
1
1,3,6
\
7,3,7
DMN - Dimethylnaphthalene
TMN - Trimethylnaphthalene
mlz I78
t Methylphenanthrenes (MP)
9
mlz 192 2
J
mlz206
(x) ì-\
13+3,9+2,10+3,10
t,6+2,9+2,5
1,9+4,9
1,8
h
Retene (R)
rnlz2l9
depositional environment of the source rocks; the source aff,rnity of the oils; and the thermal
maturity of the source rocks and the oils.
261
AROMATIC HC
1OO "/o
Reservoir
80 20
a Namur
Birkhead
60 40
I Hutton
O Poolowanna
40 60 X Patchawarra
tß Mooracoochie
20 80
Oil Classifrcation
P Paraffinic
PN Paraffinic-naphthenic
Al Aromatic - Intermediate
(aromatic HC> lOVo)
I Intermediate (aromatic }JCclOTo)
In the Poolowanna Trough 1,2,5-TMN/1,3,6-TMN values range from 0.81 at 2505 ffi, h
Poolowanna-3 to 3.04 at24l7 m, in Poolowanna-I. The present maturation level of this
section, is 0.72-0.827oR6. A strong araucarian resin input is clearly indicated at the top of
the Poolowanna Formation and decreases towards the base of the unit (Fig. 8.3a). In the
Patchawana Trough the values range from 0.21 at 1881 m in Sturt-4 to 12.5 at 1841 m in
Sturt East-l. Here, a strong araucarian signature is again evident near the top of the
Poolowanna Formation in the Tantanna (Fig. 8.3b) and SturlSturt East Fields (Fig. 8.3c). It
appears that in both the Patchawarra and Poolowanna Troughs a dominant Araucariaecean
flora became established near the end of the deposition of the Poolowanna Formation. In
early Poolowanna time, the major conifer group was the Cheirolepidiaceae (Alexander et al.,
1988).
Another araucarian marker, l-metþlphenanthrene (l-MP), is derived from abietic acid and
sandarocopimaric acid+ype natural products (Fig. 1.7). It has also been suggested to form
during catagenesis by methyl-transfer reactions which favour the most reactive cr-positions
of phenanthrene (Albrecht et al., 1976; Allan and Larter, 1983). Radke et al. (1986)
suggested that the thermal stability of 9-methylphenanthrene (9-MP) is similar to that of
1-MP because in both the methyl group occupies an o-position in the phenanthrene molecule
(Fig. 2.15). Therefore the ratio I-MP/9-MP should be largely independent of thermal
evolution. Rather, as suggested by Alexander et ø/. (1988), the l-MP/9-MP values >1 are
likely to indicate strong organic matter inputs from Araucariaecean flora. Such high values
are therefore another marker for terrestrial organic matter of Middle Jurassic to Middle
Cretaceous age.
The l-MP/9-MP ratio ranges from 0.93 to 1.24 in the Poolowanna Trough samples, and
from 0.83 to 1.43 in the Patchawarra Trough. Samples with l-MP/9-MP values less than
unity come from low in the Poolowanna Formation where the Araucariaceae were not yet a
major part of the local flora. As in the case of the 1,2,5-TMN/1,3,6-TMN ratio, strong
Araucariacean inputs are indicated near the top of the Poolowanna Formation in both
troughs.
263
Table 8.la Aromatic maturity and source-related biomarker parameters of source rock extracts, Poolowanna Formation
Rc-t Rc-z
1,3,6-TMN Vo Vo 9-MP X 9-MP P
POLl-3 Poolowanna-1 2438 3.20 0.96 0.82 1.55 0 65 0.79 0.68 0.93 0.89 0.02 0 06
POLl-5 Poolowanna-1 2499 3.35 0.96 0.71 0.84 0 63 0.18 0.66 1.01 r.02 0.05 0 00
POL2-9 Poolowanna-2 2524 2.60 0 77 0.75 1.24 0 6t
0.76 0.65 1.20 r.26 0.03 0 00
POL3-13 Poolowanna-3 2505 3.85 1 00 0.80 0.81 0 65 0.79 0.61 0.98 0.82 0.02 0 00
TAN2-18 Tantanna-2 1799 1.65 0 68 0.75 7.4r 0 36 0.61 0.41 1.43 3.41 o.32 0 08
TAN2-20 Tantat'na-2 181 I nd 0 97 0.86 1.09 0 68 0.8r 0.70 0.83 0.72 0.16 0 07
TAN3-21 Tantanna-3 1807 3.93 0 98 0.83 1.22 0 61 0.77 0.65 0.96 1.10 0.15 0 t4
TANS-25 Tantanna-8 1823 3.61 0 97 0.89 1.34 0 61 0.71 0.65 t.t6 0.95 o.29 0 l5
STUl-27 Sturt-1 t87t 1.55 0 47 0.49 8.70 0 43 0.66 0.52 1.09 1.79 0.31 0 t4
STU4-30 Sturt-4 1859 1.16 0 78 0.10 9.74 0 49 0.69 0.56 t.2r 2.83 0.21 0 T9
STU4-31 Sturt-4 1881 3.77 0 85 0.81 2.07 0.45 0.67 0.54 1.1 1 1.00 0.5s 0 15
STU6-35 Sturt-6 1865 nd 1 00 0.85 2.68 0.19 0.87 0.77 1.22 1.03 0.04 0 42
STEl-38 Sturt East-1 1841 1.30 0 49 0.57 12.50 0.41 0.65 0.51 0.96 1.54 0.35 0 15
STE4-45 Sturt East-4 r862 nd 0 94 0.83 3.75 0.6r o.7l 0.65 1.09 1.24 0.05 0 45
TMN
NON-SOURCE
Poolowanna-1
Depth: 2417 m
Poolowanna
MPI: 0.55
P
n DMP
MP l-l
SOURCE
Poolowanna-2
Depth: 2524 m
Poolowanna
MPI:0.61
OIL
Poolowanna-1
DST 2
Poolowanna
MPI:0.70
OIL
Tantanna-1
DST 3
Birkhead/Hutton
DMN MPI: 0.37
P
t"l
MP
DMP
R
il,1
SOURCE
Tantanna-2
Depth: 1799 m
Poolowanna
MPI: 0.36
J"
NON-SOURCE
Tantanna-8
Depth: 1823 m
Poolowanna
MPI:0.61
DMN
OIL
Sturt-6
DST 1
Birkhead
MPI: 0.43
P MP
DMP
SOURCE
Sturt East-l
Depth: 1841 m.
Poolowanna
MPI: 0.41
_rl
NON-SOURCE
Sturt-6
Depth: 1865 m
Poolowanna
MPI:0.79
MA
A
/
J
The retene/g-MP ratio is another molecular indicator of the relative input to the Poolowanna
Formation by Araucariacean flora provided that the main source of retene is the abietic acid
(Alexander et a1.,1988). The ratio has values ranging from 0.02 to 0.1l in the Poolowanna
Trough samples, and from 0.04 to 0.55 in those from the Patchawarra Trough (Table 8.1a).
The, higher abundance of retene in the latter samples is possibly related to higher inputs of
araucarian plant remains in the Patchawarra Trough than in Poolowanna Trough.
Once again, the highest values of an araucarian marker are shown near the top of the
Poolowanna Formation in the Poolowanna and Tantanna fields, whereas the lowest values
occur near the base of the unit in Sturt-6 (1865 m) and Sturt Easç4 (1862 m). However, the
highest retene/9-MP value of 0.55 is shown by the deepest sample (Sturr4,1881 m) in Sturt
field. In this case the enhancement of retene is probably not related to Araucariacean resin
input.
268
o.75
Jurassic
0.5
il ilI
fL
I
o, o.25
fL
I
A
o)
J
o d Otr
ao
o
o
I
I
0 9¡
A
I
Permian I I IV
I
-0.25
-1.5 -1 -0.5 0 0.5 1 1.5
Log(1,2,s-TMN/1,3,6-TMN)
Figure 8.4 Aromatic biomarker signatures of Poolowanna Formation source rocks based
on isomeric methylphenanthrenes and trimethylnaphthalenes
o Poolowanna o Sturt
il UI
0.5
Jurassic
fL 0
I
o
E(¡)
co -0.5 -l=- qo -r
É.
o to
qa
c', o
o -1
I 9tr
IV
ô
-1.5 Permian
p o
I
-2
-1 -0.5
0 0.5 1 1.5
Log(1,7-DMP/X)
Figure 8.5 Aromatic biomarker signatures of Poolowanna Formation source rocks based
on retene, and isomeric dimethylyphenanthrenes (Key as for Fig. 8.4)
Chapter I
The peak labelled 'X'(Fig. 8.1b) in the mass fragmentograms of alkylphenanthrene
homologues represents an unresolved mixfure of 1,3-,3,9-,2,10- and 3,10-
dimetþlphenanthrene isomers (Radke et al., 1986) which are thought to be products of
thermal maturation. Therefore, the l,7-DMPD( ratio may also be used to determine the
relative input from Araucariacean flora (Alexander et a1.,1988).
Like the preceding parameters, the l,7-DMPD( ratio shows its highest values near the top of
low values near the base of the unit (Table 8.1a).
the Poolowanna Formation and relatively
The values range from 0.82 at 25O5 m, in Poolowanna-3 to 1.91 at 2417 m in
Poolowanna-1 in the Poolowanna Trough; and from0.72 at 1811 m in Tantanna-2 to3.4l at
1199 m in the same well in the Patchawarra Trough. This parameter happens to be most
consistent in the Patchawarra Trough where, in every well, the values decrease with
increasing depth, hence signifying the relative increase of Araucariacean-related organic
matter up section in the Poolowanna Formation.
In the log-log cross-plot of the retene and 1,7-DMP-based biomarker parameters (Fig. 8.5)
most of the samples plot in quadrant IV. Only one sample plots in the Permian zone and
several Patchawarra Trough samples plot in the Jurassic zone designated by Alexander et aL
(1988). It is apparent from Figures 8.4 and 8.5 that the Poolowanna Formation occupies a
transitional zone between the Permian and Jurassic compositional fields (quadrants I and III,
respectively). Only those samples from the upper part of the formation (i.e. of early Middle
Jurassic age and deposited after the first appearance of the Araucarian flora) display the
characteristic Jurassic biomarker signatures.
270
Table 8.lb Aromatic maturity and source-related biomarker parameters of oils from the Poolowanna and Patchawarra Troughs
1 Rc-
No No 1,3,6-TMN 9-MP X 9-MP P
P 1
T2 Tantanna-1 Poolowanna 7 1800-1814 nd 1.05 0 93 1.03 0.81 0.88 0 .18 0.8 2 0.65 0.38 0 08
T3 Tantanna-1 Hutton 2 1635-t642 nd 0.1r 0 80 10.00 0.4t 0.64 0 .50 I 49 1.01 3.94 0 02
T4 Tantanna-1 BirkheadÆIutton 3 1341-1359 nd 0.52 0 63 5.06 0.37 0.62 0 .48 I 08 r.20 2.76 0 04
s1l Sturt-3 Poolowanna 1A 1848-1855 1.84 0.48 0 62 4.64 0.48 0.69 0 .56 I 15 1.1 1 r.o4 0 18
s12 Sturt-3 Birkhead 1B 1654-1662 2.06 0.22 0 37 9.00 o.23 0.54 0 .38 5 .00 2.86 5.43 0 08
s13 Sturt-4 Poolowanna 1 1872-1880 nd 0.59 0 69 1.9r 0.60 0.76 0 .64 0 .99 0.87 0.62 0 15
s16 Sturt-6 Mooracoochie 5 19t4-19r9 nd 0.84 0 86 r.26 0.88 0.93 0 .84 0 .78 0.5 r 0.54 0 07
st7 Sturt-6 Patchawa:ra 3 1884-1898 nd 0.89 0 9l 0.07 o.94 0.96 0 .88 0 .84 0.53 r.20 0 o4
s18 Sturt-6 Birkhead I 1883-1892 nd 0.59 0 69 4.33 o.43 0.66 0 .52 I .58 13.20 1.13 0 05
s19 Sturt-7 Patchawa:ra I 1923-1938 nd 0.14 0 87 0.92 0.83 0.90 0 .80 0 .79 0.49 0.20 0 o4
s20 Sturt-7 Mooracoochie 2 1946-202r nd 0.51 0 65 0.78 o.7r 0.83 0 .72 0 .66 0.39 0.31 0 06
s21 Sturt-7 Poolowanna 3 t87t-1816 nd 0.64 0 69 2.35 o.57 0.74 0 .62 I .22 o.t9 0.93 0 09
s23 Sturt-8 Poolowanna 1 1880-1884 t.2t 0.59 0 70 2.00 o.5l 0.74 0 .62 0 .99 0.87 1.88 0 23
SE24 Sturt East-2 Poolowanna 7 1845-1899 3.20 o.7t 0 18 1.75 0.58 0.75 0 .63 I .45 o.99 0.95 0 2t
TN,25 Taloola-2 Poolowanna 1 1829-1836 nd 0.62 0 1l t.t] 0.68 0.81 0 .10 1 .10 o.75 0.60 0 l6
TN-26 Taloola-2 Hutton 2 1789-1793 nd 0.60 0 73 1.95 0.57 0.74 0 .62 1 .13 o.67 0.65 0 2t
TN-21 Taloola-2 Namur 3 1384-1397 nd 0.52 0 6t 4.23 0.37 0.62 0 .48 I .45 1.30 2.56 0 04
The 1-MP/9-MP values range from 0.66 in the Sturt-7 (Mooracoochie) oil to 5.0 in Sturt-3
(Birkhead) oil. The Permian oils appear to lack any significant Araucariacean input
(1-MP/9-MP = 0.66-0.84), whereas most of the Jurassic oils are indicated to have an
Araucariacean source affinity which ranges from weak in the Tantanna-l (BirkheadÆIutton)
oil (1-MP/9-MP = 1.08) to very strong in the Sturt-3 (Birkhead) crude (l-MP/9-MP = 5).
The Tantanna-l (Poolowanna) crude is the only oil from a Jurassic reservoir with a
l-MP/9-MP value <1 . This indicate either that it is of mixed Permian and Jurassic origin, or
that it was derived entirely from the lower Poolowanna Formation source facies. The
distinction between oils of Permian and Jurassic origin is clearly illustrated in Figure 8.6. All
the Permian and Cambrian-reservoired oils plot in quadrant I; while most of the Jurassic oils
plot in quadrant III. Oils of possible mixed source affinity plot in the borderline region
between these two quadrants. These oils are all from Poolowanna reservoirs, making it more
likely that they originated from source beds of late Early Jurassic age in the lower
Poolowanna Formation.
The 1,7-DMP/X values range from 0.39 in the Sturt-7 (Mooracoochie) to 13.2inthe Sturt-6
(Birkhead) crude. It is again demonstrated that the Permian and Cambrian oils are
characterised by low values (0.39-0.51) and are therefore genetically unrelated to the
Araucariaceae. Most of the Jurassic oils have 1,7-DMP/X values >1, but quite a number
from Poolowanna reservoirs and the Taloola-2 (Hutton) oil show values <1. These low
values may indicate either a mixed oil charge from adjacent Permian and Jurassic source
rocks, or just derivation from, pre-Araucariacean Jurassic source rocks.
The retene lg-MP values of the oils range from 0.20 in the Sturt-7 (Patchawarra) crude to
5.43 inthe Sturt-3 (Birkhead) sample. The Permian and Cambrian oils appear to have values
at the low end of the range, with the exception of the Sturt-6 (Patchawarra) crude which has
a value of 1.20. This enhanced value may be due to a local alternative source of retene,
possibly phyllocladane. It is significant that this oil is the most mature sample analysed (MPI
= O.94: Table 8.1b) It is also shown that most of the Jurassic oils show values >0.3,
consistent with the lower limit of the Jurassiczone in Figure 8.7. In this cross-plot the
Jurassic oils are more poorly discriminated from Permian and Cambrian oils than in Figure
8.6; and the Poolowanna oils once again tend to plot in between those in the Permian and
younger Jurassic reservoirs.
272
Chapter I
Table 8.lc Abbreviations of aromatic compounds and definitions of aromatic maturity
parameters (also see Table2.6)
TMN Trimethylnaphthalene
P Phenanthrene
A Anthracene
I\44 Methylanthracene
MP Methylphenanthrene
DMP Dimethylphenanthrene
X 1,3-DMP + 3,9-DMP +
2,10-DMP + 3,10-DMP
It is apparent shown from this plot that oils from different reservoirs within the same well
have different biomarker signatures, an indication of them not having a common source or
origin. On the other hand, oils from the same reservoir unit in different wells display similar
molecular signatures (Table 8.2).
Tantanna-l Oils from the Poolowanna, Hutton and Birkhead reservoirs (Fig. 8.9a) are
represented in this well. The Poolowanna oil plots in the Permian zone and is quite different
from the Hutton and BirkheadÆIutton oils which are clearly of Jurassic origin. In the well
correlation diagram for this fietd (Fig. l.4b) it can be seen that the Poolowanna oil was
recovered from near the base of Poolowanna Formation. Therefore it is almost certain that
this oil has migrated from an underlying Permian source particularly in view of its relatively
high maturity (MPI = 0.81). The other oils were generated from Jurassic source rocks, most
likely from within the upper Poolowanna Formation (Fig. 8.3b).
273
Chapter I
Table 8.2 Source-dependent biomarker signatures in oils from the Sturt Field,
Patchawarra Trough
Sturt-3 The oils recovered from the Birkhead and Poolowanna reservoirs in this well both
plot in the Jurassic zone of Figure 8.8. Although they are both of Jurassic origin it is
obvious that they have different sources. Because the Birkhead oil has a much higher
|-MP/!-MP value its source had a much stronger Araucariacean input than that of the
Poolowanna oil. Hence it is likely that the Birkhead crude was generated from a local
Birkhead source rock whereas the Poolowanna oil originated from the upper Poolowanna
source facies.
Sturt-6 This well produced oils from reservoirs in the Birkhead and Patchawarra Formations
and the Mooracoochie Volcanics (Fig. S.gb). The Birkhead oil is well separated from the
Permian and Cambrian oils in Fig. 8.8 and has a molecular signature similar to that of an
upper Poolowanna Formation source rock (Fig. 8.3c). At the same time the Patchawarra oil
appears to differ from the Cambrian oit by having a much lower 1,2,5-TMN/1,3,6-TMN
value, indicating their origin two different Permian source facies.
274
0.75
^
Jurassic
0.50
II III
fL
I
E o.25
fL
A
I
o a
o
o)
o o
o
tO-
T t
0
xI >O¡
I
IV
Permian x
-o.25
-1.5 -1.0 -0.5 0.0 0.5 1.0 1.5
Log (1,2,s-TMN/1,3,6-TMN)
1.0
II a Jurassic III
0.5
A
fL 0.0
I
o)
o
C
o -0.5 +- o f I
o x I
É.
I
o)
o -1.0 I
O
I
Permian I
I
-1.5
I I
I
IV
I
-2.0
-1.0 -0.5 o.o 0.5 1.0 1.5
Log(1,7-DMP/X)
Figure 8.7 Oil source affinity based on retene and isomeric dimethylphenanthrenes (Key
as in Fig. 8.6)
0.7
I
I
Tantanna-1 I
Sturt-3 I
¡ Sturt-6 .o
U)
0.5 (t)
Sturt-7 (ú
L
f
Taloola-2 I
fL
I
O)
È 0 .25
I
(',
o
J
't
Permian I
-0.25
-1.5 -1.0 -0.5 0.0 0.5 1 1
Log(1,2,s-TMN/1,3,6-TMN)
Reservoir
a Namur x Patchawama (Permian)
Birkhead Mooracoochie
I Hutton
Jurassic x Volcanics (Cambrian)
o Poolowanna
Figure 8.8 Source affinity of oils in the stacked reservoirs of five wells in the Sturt,
Tantanna and Taloola fields
Chapter I
TaIooIa-2 This well flowed oils from the Namur, Hutton and Poolowanna reservoirs. Here,
the Poolowanna oil plots between the Permian and Jurassic quadrants of Figure 8.8 and,
therefore is tikely to be of mixed Permian/Jurassic origin, or to have been sourced from the
lower Poolowanna Formation. The Hutton and Namur oils are indicated to be of Jurassic
source affinity.
8.4 Source rock and oil maturity based on aromatic molecular parameters
Aromatic molecular parameters used in the determination of source rock and crude oil
'Welte,
maturity include MPI (Radke and 1983), DNR-I (Radke et al., I982b), TNR-I
(Alexander et a1.,1985) and TNR-2 (Radke et a1.,1986). Their behaviour with respect to
thermal evolution is discussed in Section2.7.2.
In the Poolowanna Trough, there is good agreement between the Rs-l values of the
Poolowanna Formation (0.13-0.797o) and corresponding measured reflectance values (Ro =
O.75-0.79 Vo). However, in the Patchawarra Trough, it is the Rs-2 values (in the range
0.47-O.l|Vo) that more closely match the measured vitrinite reflectance (Ro = 0.57-O.67Vo).
This suggests that no single calibration of MPI is appropriate for both the Poolowanna
Trough and the Patchawarra Trough. Possibly there are significant differences in the burial
and thermal histories of the Poolowanna Formation between these two depocentres.
The DNR-1 values range from 1.16 at 1859 m in Sturt-4 to 3.85 at2505 m in Poolowanna-3
(Table 8.1a). These values correspond to measured vitrinite reflectances of O.64Vo and
O.79Vo Rs, thus showing a good correlation between these two parameters. However,
samples, from 1807 m in Tantanna-3, 1823 m in Tantanna-8, and 1881 m in Sturt-4, with
measured vitrinite reflectance values of 0.627o, O.58Vo and O.57Vo, have DNR-I values of
3.93,3.67 and3.77, respectively. In this case the dimethylnaphthalene ratios indicate a
higher maturity than that derived from vitrinite reflectance. Coincidentally, these samples
happen to be of the silty coal and coal facies (Table 1.2), and therefore this ratio may be
influenced by the type of organic matter or its depositional environment. Even so, DNR-I
in Sturt-4 (Fig. 8.
appears to be a reliable maturity parameter for the Poolowanna sequence
l0). Generally the Poolowanna Trough source rocks seem to be characterised by DNR-I
values >2, whereas in the Patchawarra Trough the values extend over a wider range (1.3-
3.93) and are likely not to be related only to maturity.
217
TMN
DMN
Tantanna-1
P DST 3
Birkhead/Hutton
MPI:0.37
MP
Jurassic source
DMP
t_
DST 2
Hutton
MPI: 0.41
Jurassic source
DST 1
Poolowanna
MPI: 0.81
Jurassic+?Permian
source
Sturt-6
DMN DST 1
Birkhead (Jurassic)
MPI: 0.43
MP
P
DMP
DST 3
Patchawarra (Permian)
MPI: 0.94
DST 6
Mooracoochie (Cambrian)
MPI: 0.88
Sturt-6
DMN DST 1
Birkhead (Jurassic)
MPI: 0.43
MP
P
DMP
DST 3
Patchawara (Permian)
MPI: 0.94
DST 6
Mooracoochie (Pre-Permian)
MPI: 0.88
This is verified by early mature coal samples at 1865 m in Sturt-6, 1881 m in Sturt- , L823
min Tantanna-8, 1807 m in Tantanna-3 and 1862 m in Sturt East-4 which have TNR-I
values between 0.85 and 1.00 while their corresponding Rs is between 0.57 and 0.66Vo. On
the other hand the Poolowanna Trough carbonaceous shales show a rather good correlation
between TNR-I and R6 values. Both DNR-I and TNR-I appeff to correlate well as
illustrated by Figure 8.9. Some of the coal samples plot in the mature zone although they are
of relatively low maturiry according to their measured vitrinite reflectance. Either the
measured vitrinite reflectance was suppressed by bituminite dissemination in vitrinite, in
which case DNR-I and TNR-I are the more reliable maturity indicators, or the elevated
values of the aromatic parameters are due to differences in organic input and"/or depositional
conditions.
Recognising that the fu values of oils correspond to the maturity of their source beds at the
time of expulsion, the observed Rc values may reflect the type of kerogen responsible for oil
generation. This is because different kerogens expel oils at different maturation levels. As
explained by Radke (1987), the oils with Rs values ranging from 0.5 to 0.7Vo are paraffinic,
commonly waxy, and have low sulphur contents. They are derived from liptinite-rich
lacustrine Type I kerogen and thus are known as Type I oils. However, Radke (1987) also
points out that these oils are strikingly similar to the bitumen in source rocks containing
immature Type III kerogen. Non-marine oils with fu values in the range 0.75-0.9Vo were
assigned to either the TypeIIIIII or Type trI/II category. In the present study most of the oils
in Jurassic reservoirs of the Eromanga Basin were generated by Jurassic source rocks
containing early mature to mature Type IIIIII kerogen. The Tantanna-l (Poolowanna) oil,
and those oils in Permian and Cambrian reservoirs, are indicated to be generated from more
mature Type IIIIII or Type III/II kerogens (Table 8.3).
280
6
DNR-1 =5.22(TNR-1 )-1 .595; r=0.958
4
o
I
É.a
z"
o
2
o ^,
tr
o
1
Figure 8.10 Relationship between the dimethylnaphthalene ratio (DNR-l) and the
trimethylnaphthalene ratio (TNR-l) in source rocks of the Poolowanna Formation
Chøpter 8
The Poolowanna-l (Poolowanna) oil, while almost certainly derived from a local lower
Poolowanna source facies, is the most mature of indigenous Jurassic crudes (Rc-l -
O.82Vo). Rs-1 is the preferred measure of this oil's maturity because Rs-l gives a better
correlation with measured vitrinite reflectance (Ro) in the source rocks of the Poolowanna
Trough (see Section 8.4.1).
Table 8.3 Oil families based on aromatic maturity and source affinity parameters
Rc
Reservoirl
Well DST Kerogen type Possible
Vo
source(s)2
1. Oils listed in order of increasing maturity (Rc-2;except for Poolowanna-1, DST 2 where
Rs-l value used).
2. Identified by age, as follows: 1 = Middle Jurassic (Birkhead); 2 = Middle Jurassic (upper
Poolowanna); 3 = Early Jurassic (lower Poolowanna);4 = Permian; 5 = Cambrian
Some oils particularly those in the Namur and Birkhead reservoirs, have very low fu-2
values (<O.55Vo), which are possibly not a reliable indication of their true maturity. Thus the
observed Rs-2 values in this low range are probably due to enrichment of l-methyl-
phenanthrene in the expelled oil. Anomalously high 1-MP/9-MP ratios have been found in
other Birkhead-derived oils and source rocks (D. M. McKirdy, pers. comm.,1997).
282
6
DNR-1=8.863 (TNR-1 ) -3.01 8; r=0.83
4 o
t
E
zo
2
O
0
0.2 0.4 0.6 0.8 1 1.2
TNR-1
Reservoir
o Poolowanna
A Birkhead
Figure 8.11 Relationship between the dimethylnaphthalene ratio (DNR-l) and the
trimethylnaphthalene ratio (TNR-1) in oils from the Poolowanna and Patchawarra
Troughs
Chapter 8
The only available DNR-I values for the oils range between 1.21 and 4.10 (Table 8.16). In
most samples the 1,5-dimethylnaphthalene peak (Fig. 8.1a) was not resolved. The TNR-I
values range fromO.22 in the Sturt-3 (Birkhead) oil to 1.05 in the Tantanna-1 (Poolowanna)
oil. Most of the Permian and Cambrian oils have values ranging from 0.74 to 0.89. The
Sturt-7 (Mooracoochie) oil, with aDNR-I value of 0.51, is an obvious exception, possibly
because of the different organic matter (marine algal) and depositional environment (anoxic
to suboxic) of its source rock. The high TNR-I value of 1.05 in the Tantanna-l
(Poolowanna) oil is consistent with its Permian origin, whilst the values of the other Jurassic
oils show them to be less mature than the Permian and Cambrian oils. The DNR-I versus
TNR-I cross-plot (Fig. 8.11) reveals a good correlation between these two parameters. In
this diagram the Sturt-3 (Birkhead) oil plots well away from the regression line, although the
reason for this is not obvious,
The relationship between the MPI and TNR-2 parameters for the Poolowanna source rock
extractsisillustratedinFigure 8.12. Many samples including the coals, plot well to the left
of the trend line (taken from Radke, 1987 , fig. 25). Most of these off-trend samples are from
the Patchawarra Trough. The Poolowanna Trough samples seem to def,rne a trend of
increasing maturity, whereas the Patchawarra Trough samples a.re more scattered. This sort
of distribution is analogous to that observed for the maceral group distribution (Fig. 4.1).
Therefore the relationship between these two ratios may well reflect variations in organic
facies. So far no other geochemical evidence has indicated signif,rcant secondary alteration in
these samples. Thus, their distribution in this plot is most likely related to maturation and
primary organic matter type.
284
1.0
0.9 /
Coal az' o
0.8 o
A ó
o.7 o o
o
C\
É. 0.6
zF ô
0.5 o
0.4
0.3
0.2
0.2 0.3 0.4 0.5 0.6 o.7 0.8 0.9 1.0
MPI
Field
o Poolowanna O Sturt
Figure 8.12 Relationship between the trimethylnaphthalene ratio (TNR-2) and the
metþlphenanthrene index, (MPI) in source rock facies of the Poolowanna Formation
o o"a
o¡É
kn
öo TNR-2
äoo
o o J
;i. u) o 9
q)
I('l P I@ ¿o
Õ i\) o, \ b
IÞ
ÊÞ F+)
or>a Itu
E(DÞ't
Ë
Ø
A. CD x' F..7
N õ ã +Ë
o
p
Iq)
eô) g
Þì+ o ? H.E
N9 o ir-
l.*) B a>
3t
oo
(D:
€ o
'Þ
(D
t-,ì-
Ø,
v r_¡t r-t
É
Ft
Èä {
Þe Þ
(t
o o\
2". Ø Ø i¡
FÔ o () F
l.)9. À o
rt)
ÞØ o o
Þ- i' C'
¡ßx ã !;; i\
ZE Þ
Hl ¿?r¡r Â:ú
(D (D ^-
ã a
r-.t
r_t
F Fg s I\¡ ¡o
Þ Þ
p È
,r
ãË
ØX
i
Þ
(tI o È
;- ^ô
ÔF ¡¡ I@
È-t \o x I
o
(D
æ Ëõ'Q
a.
\¡ =CD
J.H ¡ß
o Þ=.
5ÊD
I(o
Ø \/5 x
o
r_t J
()
(D b
È
I
Þ
(t
(D
È
o
p
Chapter 8
The relationship between MPI and TNR-2 in the oils is shown in Figure 8.13. As far as
alteration is concerned, only four samples (one from the Hutton two from Birkhead and one
from the Namur reservoirs) plot off-trend, possibly in the zone of water-washed oils.
However, all four of these oils have l-MP/9-MP ratios >1, suggesting araucarian-related
input of l-methylphenanthrene to their source rocks which thereby depresses their MPI
values. The Tantanna-l (Poolowanna) oil plots together with the Permian-sourced oils, thus
confîrming that it originated from a Permian source rock.
Acomparison of the cross-plots for the source rocks (Fig. 8.12) and the oils (Fig. 8.13)
suggests that many of the oils from Jurassic reservoirs could have originated from source
rocks of the same organic facies and maturity as exist within the Poolowanna Formation on
the southwest margin of the Patchawarra Trough.
The oils from the Mooracoochie Volcanics in Sturt-7, the Patchawarra Formation in Sturt-6
and Sturt-7, and the Poolowanna Formation in Tantanna-l have all been assigned a Permian
origin on the basis of isotopic and biomarker evidence. It is therefore signif,rcant that they
plot quite separately from the Jurassic and Cambrian-sourced oils in Figure 8.13. Neither
does the position of these Permian and Canrbrian oils on the plot coincide with any of the
Poolowanna Formation source extracts.
281
Chapter 9
Chapter 9
The Poolowanna Formation in both the Poolowanna Trough and the southwestern
Patchawana Trough is characterised by good source rock facies. Total organic carbon (TOC)
values range from l.5Vo in silty shales to 10Vo in coals. The dispersed organic matter is
commonly enriched in liptinite, and the amount of extractable hydrocarbons (1018-15384
ppm) is evidence of very good to excellent source richness.
Maturity data based on both non-biomarker and biomarker parameters show that these source
rocks at the early mature to mafure stages of hydrocarbon generation from terrestrial
a.re
organic matter. Vitrinite reflectance values of 0.54.9Vo Rs and calculated reflectance values
of 0.47- 0.7lVo Rs are consistent with homohopane (C32: 22S|22S+22R) and sterane (C2e:
Oil and gas-prone Type IVIII kerogen is the major variety of organic matter preserved in the
Poolowanna Formation, as indicated by the relative abundance of the vitrinite, liptinite and
inertinite maceral groups, and verified on the HI versus T*o* cross-plot. The actual range of
kerogen compositions is from Type I/II to Type IVIII in the Poolowanna Trough; and from
Type II/I[ to Type III in the Patchawarra Trough, The presence of significant amounts of
resinite, which may be assigned to kerogen Type I, implies the possibility of generating oil at
early maturity in the Poolowanna Trough. The best example of this resinite-rich facies is the
2496-2524 m interval in Poolowanna-l and2 which has highest recorded HI values (315-
348 mg HC/g TOC). Both resinite and Botryococcus-like telalginite impart a Type I character
to the kerogen but their influence is obscured by the inertinite and thus not easily recognised
by Rock-Eval pyrolysis (i.e. in the HI vs OI and HI vs T,o* cross-plots).
289
Chapter 9
This Type IIIIII kerogen is attributed to the mixed inputs of organic matter from higher
plants, non-marine algae and bacteria found in tenestrial to lacustrine depositional
environments. The land plant input is indicated by very high pristane to phytane ratios (>3)
in conjunction with C2e-dominant sterane and diasterane distributions in some cases, and the
carbon isotopic signatures of the saturated and aromatic hydrocarbon fractions. The algal
input is represented by the presence of Botryococcus-like telalginite, which in turn is a
reliable indicator of lacustrine depositional conditions. The non-waxy affinity shown by the
carbon isotopic signatures may be related to aquatic plants and non-marine bacteria. The
bacterial input or influence is demonstrated by high hopane to sterane ratios (>2), and the
presence of C1a-C16 drimanes and C2e-Cy 2a- and 3B-methylhopanes. The n-alkane
distributions and the isotopic signature of the saturated hydrocarbon fractions give the
impression that algae, bacteria and aquatic plants contributed more organic matter to the
Poolowanna Formation than did land plants.
Oxic to sub-oxic depositional conditions appear to have prevailed during the accumulation of
the organic matter. The oxic conditions are strongly indicated by the aforementioned
pristane/phytane values and a notable predominance of inertinite over vitrinite (i.e. I/V >1) in
most samples. The anomalously high W ratios, in the Patchawarra Trough a¡e related to
deposition of recycled vitrinite and inertinite rather than primary oxidation effects. Sub-oxic
conditions are indicated where pyrite is present and where vitrinite is signihcantly more
abundant than inertinite (W <1). Acyclic isoprenoid to n-alkane ratios (i.e. prln-C17 versüs
pNn-Cß) confirm that oxidising conditions were more prevalent than reducing ones.
The first appe¿ìrance of a new Jurassic flora, the Araucariaceae, midway through the
deposition of the Poolowanna Formation led to a major change in the biomarker signature of
its coals and carbonaceous shales. The resin acids of these conifers contributed a novel suite
of aromatic hydrocarbons to the upper part of the formation, viz. I,2,5-
trimethylnaphthalene, 1-methylphenanthrene, 1,7-dimethylphenanthrene and retene. Elevated
concentrations of these biomarkers (particularly 1,2,5-TMN and l-MP) are displayed
towards the top of the Poolowanna Formation, more so in the Patchawarra Trough where
limnic peat swamps were better developed.
There are important differences between the two depocentres of the Poolowanna Formation
as far as its source rock potential is concerned. The primary differences relate to its
depositional environment, which in the Poolowanna Trough was predominantly telmatic
with arborescent wet forest swamps while in the Patchawarra Trough it was mainly fluvio-
lacustrine with herbaceous limnic swamps. In both the waxy and non-waxy source facies of
the Poolowanna Formation the carbon isotopic signature of the aromatic hydrocarbon
fraction is heavier in the Poolowanna Trough than in the Patchawarra Trough. This isotopic
difference is largely due to the greater maturity of the Poolowanna Trough source rock
290
Chapter 9
facies.
Most of the oils from Jurassic, Permian and Cambrian reservoirs in the study area seem to be
generated from source rocks that were deposited in oxic to sub-oxic terrestrial swamp and
lacustrine environments. They include oils of both waxy and non-waxy source affinity as
illustrated by their n-alkane prof,rles and the ca¡bon isotopic signatures of their aromatic and
saturated hydrocarbons. Algae, bacteria and aquatic plants appear to be the main sources of
these hydrocarbons while the land plant input is significant onlyin the waxy oils. The major
contribution of prokaryotic organisms (bacteria) to the parent source rocks is emphasised by
the high hopane to sterane ratios (>2.5) of the oils. In the Patchawarra Trough, the two
Permian oils together with the out-of-place Tantanna-1 (Poolowanna) and Sturt-6
(Mooracoochie) crudes have MPl-derived maturities (Rc = 0.8-0.97o) which indicate that
they were derived from mature source rocks containing Type III/II kerogen, presumably in
the Early Permian Patchawarra Formation. Less mature Type IVI and Type IIIIII kerogen of
Early to Middle Jurassic age ale the likely sources of the Jurassic oils (Rc = 0.5-O.l5Vo).
The Sturt-7 (Mooracoochie) oil is of suboxic, algal source affinity and probably originated
from an unidentified Cambrian marine mudstone containing Type IVIII kerogen in the
underlying Warburton Basin. In the Poolowanna Trough, the waxy Poolowanna-l
(Poolowanna) crude has an unusually light carbon isotopic signature which in part reflects its
alteration by water washing. It is the most mature of the Jurassic oils analysed in this study
(Rc = 0.82%) and was probably generated locally from Type IVIII kerogen in the lower
Poolowanna Formation.
The oils in the Permian and Cambrian reservoirs of the Patchawarra Trough appear to be
generatedfrom source rocks of higher maturity than those in the Jurassic reservoirs. The
Namur Sandstone and Birkhead Formation oils are clearly related to early mature source
rocks. The present maturity range of the local Poolowanna Formation source rocks,
corresponds rather well to that of the oils in adjacent reservoirs of the Poolowanna
Formation, Hutton Sandstone, Birkhead Formation and Namur Sandstone. Thus, on the
grounds of both kerogen type and maturity, the Poolowanna Formation is a possible source
of these oils. The evidence which supports this argument includes the cross-plot of measured
and calculated vitrinite reflectance versus carbon isotopic signature of the aromatic
hydrocarbons; and that of trimethylnaphthalene ratio (TNR-2) versus methylphenanthrene
index (MPI). In both plots, most of the Jurassic oils are separated from those in the Permian
and Cambrian reservoirs and closely match the Poolowanna Formation source rocks. The
one Jurassic-reservoired oil which plots together with the Permian ones has a Permian
aromatic biomarker signature. Finally on the sterane isomerisation diagram [i.e.
ac¿a(20S)/crcra(20R) versus oÞÞ(20R)/ctctct(20R)l the Jurassic oils plot along the
maturation pathway delineated by the Poolowanna Formation source rock facies.
29r
Chapter9
A clear distinction between the Jurassic and Permian/Cambrian oils in the Patchawarra
Trough can be made using the previously described Araucariacean markers particularly
I,2,5-tnmethylnaphthalene and l-methylphenanthrene. Again, all but one of the oils
recovered from Jurassic reservoirs have unambiguous Middle Jurassic aromatic biomarker
signatures, and therefore must have originated from source rocks in either the upper
Poolowanna Formation or the Birkhead Formation. In the Poolowanna Trough, the mass
fragmentograms of triterpanes (mlz l9l), metþlhopanes (m/z 205) and steranes (mlz 217)
of source rock facies in the lower Poolowanna Formation correlate well with those of the
Oils of mixed Permian and Jurassic origins are difficult to distinguish from those which
originated within the Early Jurassic source rock facies of lower Poolowanna Formation. One
such example may be the Taloola-2 (Poolowanna) oil which has a transitional a¡omatic
biomarker signature and a maturity (Rs = 0.70Vo) at the high end of the range for bona fide
Jurassic oils. On the other hand the Tantanna-l (Poolowanna) and Sturt-6 (Mooracoochie)
oils are clearly of Permian origin by virtue of their lack of Araucariacean resin biomarkers
and their relatively high maturity (Rc =0.784.84Vo).
Generally, most of the oils in Jurassic reservoirs of the Patchawarra Trough appear to have
been generated from within the Eromanga Basin, and probably from the Poolowanna
Formation. The presence of significant amounts of resinite and resinous vitrinite imparts the
potential for early generation of the immature oils (Rc - O.SVo) found in the Hutton,
Birkhead and Namur reservoirs of the Tantanna and Taloola fields, although these oils could
also have originated from a local Birkhead source rock. Those oils which lack Jurassic
biomarkers may have been generated solely from source rock facies in the lower half of the
Poolowanna Formation which was deposited prior to the emergence of the Araucariacean
conifers.
Finally, it needs to be pointed out that the only source rock samples analysed in this study
came from the crest of anticlinal structures on the southwest margin of the Patchawarra
Trough. The effective Poolowanna oil kitchen may be located further to the northeast in a
deeper part of the basin, e.g. near Lake Hope-1 (Fig. I.2b).
292
Chapter 9
Bibtiography
van Aarssen, B. K. G. and de Leeuw, J. V/. (1991) Structural elucidation of polymeric
diterpenoids in fossil gymnosperm resins by means of pyrolysis combined with
GC-MS. Am. Chem. Soc. Div. FueI Chem. Prepr.36,774-780.
Abelson, P. H. and Hoering, T. C. (1961) Carbon isotope fractionation in formation of
amino acids by photosynthetic organisms. Proc. Natl. Acad. Scl. US 47, 623-632.
Aizenshtat, Z. (1973) Perylene and its geochemical significance. Geochim. Cosmochim.
Acta 37,559-568.
Albaiges, J. (1981) Identification and geochemical significance of long chain acyclic
isoprenoid hydrocarbons in crude oils. In Advances in Organic Geochemistry 1979
(Edited by Douglas, A. G. and Maxwell, J. R.), pp. 19-28. Pergamon, Oxford.
Albaiges, J. and Torradas, J. M. (1974) Significance of even-carbon n-paraffin preference
of Spanish crude oil. Nature 250,561-568.
Albrecht, P., Vandenbroucke, M. and Mandengue, M. (1976) Geochemical studies on the
organic matter from the Douala Basin (Cameroon). I. Evolution of the extractable
organic matter and the formation of petroleum. Geochim. Cosmochim. Acta 40,
19t-199.
Alekseyev, F. 4., Chakmachev, V. 4., Krylova, T. A. and Vinogradova, T. L. (1975)
Variation of hydrocarbon composition of low-boiling cuts and isotopic composition
of their carbon during migration. Int. GeoL Rev.17,895-902.
Alexander, G.,Hazai, I., Grimalt, J. and Albaiges, J. (1987) Occurrence and
transformation of phyllocladanes in brown coals from Nograd Basin, Hungary.
Geochim. Cosmochim. Acta 5I, 2065-207 3.
Alexander, R., Gray, M. D., Kagi, R. I. and Woodhouse, G. W. (1980) Proton magnetic
resonance spectroscopy as a technique for measuring the maturity of petroleum.
Chem. Geol.30, 1-14.
Alexander, R., Kagi, R. and Noble, R. (1983) Identification of bicyclic sesquitelpenes,
drimane, and eudesmane in petroleum. J. Chem. Soc., Chem. Comnt.,226-228.
Alexander, R., Kagi, R. and Sheppard, P. (1984) I,8-Dimethylnaphthalene as an indicator
of petroleum maturity . Nature 308,442-443.
Alexander, R., Kagi, R. I., Rowland, S. J., Sheppard, P. N., and Chirila, T. V. (1985)
The effects of thermal maturity on distributions of dimethylnaphthalenes and
trimethylnaphthalenes in some ancient sediments and petroleum. Geochim.
Cosmochim. Acta 49, 385-395.
Alexander, R., Noble, R. 4., and Kagi, R. I. (1987) Fossil resin biomarkers and their
application in oil to source-rock correlation, Gippsland Basin, Australia. Aust. Pet.
Explor. Assoc. J.27(l), 63-12.
Alexander, R., Larcher, A. V., Kagi, R. L, and Price P. L. (1988) The use of plant-derived
293
Chapter 9
biomarkers for correlation of oils with source rocks in the Cooper / Eromanga Basin
system, Australia. Aust. Pet. Explor. Assoc. J.28(I),310-324.
Allan, J. and Larter, S. R. (1983) Aromatic structures in coal macerals and kerogens. In
Advances in Organic Geochemistry 1981. (Edited by Bjor@y, M. et al.), pp. 534-
545. Pergamon Press, Oxford.
Alpern, B. (1980) Pétrographie du kérogenè. In Kerogen. (Edited by Durand, B.), pp.
339-311. Éditions Technip, Paris.
Anders, D. E., Doolittle, F. G. and Robinson, W. E. (1973) Analysis of some aromatic
hydrocarbons in benzene-soluble bitumen from Green River Shale. Geochím.
C osmochim. Acta. 37, l2I3-I228.
Anderson, K. 8., Winans, R. E., Botto, R. E. (1992) The nature and fate of natural resins
in the geosphere II. Identification, classification and nomenclature of resinite. Org.
Geochem.18, 829-841.
Anonymous (1988) Tantanna-1 well proposal and drilling program. Report for Santos Ltd.
(unpublished).
Aplin, R. T., Cambie, R. C., and Rutledge, P. S. (1963) The taxonomic distribution of
some diterpene hydrocarbons. Phytochemistry 2, 205 -214.
Aquino Neto, F.R., Trendel, J. M., Restle, 4., Connan, J., Albrecht, P. and Ourisson, G.
(1982) Novel tricyclic terpanes (C1e and C2e) in sediments and petroleums.
Tetrahedron Lett. 23, 2021 -2030.
Aquino Neto, F. R., Trendel, J. M., Restle, 4., Connan, J. and Albrecht, P. A. (1983)
Occurrence and formationof tricyclic terpanes in sediments and petroleums. In
Advances in Organic Geochemistry 1981 (Edited by Bjor@y, M. et aI.), pp. 659-
676. V/iley, Chichester.
Armstrong, J. D. and Barr, T. M. (1986) The Eromanga Basin. An overview of exploration
and potential.ln Contributions to the Geology an^d Hydrocarbon Potential of the
Eromanga Basin (Edited by Gravestock, D. I., Moore, P. S. and Pitt, G. M.), pp.
25-38. Geol. Soc.Aust. Spec. Publ. No. 12.
Balme, B. E. (1952) On some spore specimens from British Upper Carboniferous coals.
Geol. Mag.89, 175-184.
Barghoorn, E. S. (1949) Degradation of plant remains in organic sediments. Harvard Univ.
Bot. Mus. Leafl.l4(I), l-20.
Barnes, M. A. and Barnes, W. C. (1983) Oxic and anoxic diagenesis of diterpenes in
lacustrine sediments. In Advances in Organic Geochemistry 1981. (Edited by
Bjor@y,M. et al.), pp.289-298. Wiley, Chichester.
Barr, T.M. and Youngs, B.C. (1981) Cuttapirrie-1, an oil discovery in the Early Jurassic of
the Eromanga Basin. Aust. Pet. Explor. Assoc. J.21(l),60-70.
Barrick, R. C. and Hedges, J. I. (1981) Hydrocarbon geochemistry of the Puget Sound
region - IL Sedimentary diterpenoid, steroid and triterpenoid hydrocarbons.
Geochim. Cosmochim. Acta 45, 381-392.
294
Chapter 9
Bartram, K. M., Jeram, A. J. and Selden, P. A. (1987) Arthropod cuticles in coal. J. Geol.
(Lond.) 144, 513-511.
,Soc.
Battersby, D. G. (1916) Cooper Basin gas and oil fields. In Economic Geology of
Australasia and Papua New Guinea 3. Petroleum. (Edited by Leslie, R. B., Evans,
H. J. and Knight, C. L.), pp. 321-368. Australian Institute of Mining and
Metallurgy Monograph 7, Melbourne.
Beech, A. (1989) Sturt East-4 well completion report. Report for Santos Ltd.
(unpublished).
(lg7l) ttclt'Cratios of plants in relation to the pathway of
Bender, M. M. Variations in the
photosynthetic carbon dioxide fixation. Phytochemistry I0, 1239 -1244.
Bendoraitis, J. G. (1974) Hydrocarbons of biogenic origin in petroleum - aromatic
triterpenes and bicylic sesquiterpenes. In Advances in Organic Geochemistry 1973
(Edited by Tissot, B. and Bienner, F.), pp. 209-224. Éditions Technip, Paris.
l3C
Benner, R., Fogel, M. L., Sprague, E. K., and Hodson, R. E. (1987) Depletion of in
lignin and its implications for stable carbon isotope studies. Nature 329,708-110.
Berkaloff, C., Casadevall, E., Largeau, C., Metzger, P., Perracca, S. and Virlet, J. (1983)
The resistant polymer of the walls of the hydrocarbon-rich alga Botryococcus
braunii. Phytochemistry 22, 389 -397 .
Bisseret, P., Zunde| M. and Rohmer, M. (1985) Prokaryotic triterpenoids 2. 2B-Methyl
hopanoids from Methylobacterium organophilum and Nostoc muscorum, a new
series of prokaryotic triterpenoids. European J. Biochem. 150,29-34.
Blumer, M. (1973) Chemical fossils: trends in organic geochemistry. Pure AppI. Chem.34,
591-609.
Blumer, M. (1975) Curtisite, idrialite and pendletonite, polycyclic aromatic hydrocarbon
minerals: their composition and origin. Chem. Geol.16,245-256.
Blumer, M. and Thomas, D. V/, (1965) Phytadienes in zooplankton. Science 149, 1148-
1149.
Blumer, M. and Youngblood, W.'W. (1975) Polycyclic aromatic hydrocarbons in soils and
Recent sediments. Science 1.88, 53-55.
Bocks, S. M., Cambie, R. C. and Takahashi, T. (1963) Podocarpaceae. YIII. Macrophyllic
acid, a bisditerpenoid from Podocarpus macrophyllus. Tetrahedron 19, 1 109-1 1 16.
Bodard, J. M., Douthitt, C. 8., Denton, P., Edwards, N., Gates, 4., Gregory, R. T.,
Menhenitt, C., Pianalto,E., Wall, V. J. and Walsh, R. (1985) Hydrocarbon
reservoir quality in the CooperÆromanga Basins. Progress Report to the National
Energy Research, Development and Demonstration Council. NERDDC Project No.
808 (unpublished).
Boon, J. J., HineS, H., Burlingame, A. L. I., Klok, J., Rijpstra, W. I. C., De Leeuw, J.
W., Edmunds, K. E. G. (1983) Organic geochemical studies of
and Eglinton,
Solar Lake laminated cyanobacterial mats. In Advances in Organic Geochemistry
1981 (Edited by Bjor/y, M. et aL),pp.201-221. Wiley, Chichester.
295
Chapter9
Boreham, C. J., Crick, L H., and Powell, T. G, (1988) Altemative calibration of the
Methylphenanthrene Index against vitrinite reflectance: Application to maturity
measurements on oils and sediments. Org. Geochem. t2,289-294.
Brassell, S. C. (1985) Molecular changes in sediment lipids as indicators of systematic early
diagenesis. Phil. Trans. R. Soc. Lond. Ser. A315,57-75.
Bray, E. E., and Evans, E. D. (1961) Distribution of n-paraffins as a clue to recognition of
source beds. Geochim. Cosmochim. Acta 22, 2-I5.
Bray, E. E. and Evans, E. D, (1965) Hydrocarbons in non-reservoir-rock source beds. Am.
Assoc. Pet. GeoI. 8u1L 49,248-251.
Brooks, J. D. (1970) The use of coals as indicators of the occurrence of oil and gas. Aust.
Pet. Explor. Assoc. "f. 10(1), 35-40.
Brooks, J. D. and Steven, J. R. (1967) The constitution of an Australian brown coal resin.
FueI 46, 13-18.
Cambie, R. C. and Weston, R. J. (1968) Chemotaxonomy of the New Tnaland
Podocarpaceae. J. New Zealand Inst. Chem.32,I05-I21.
Cambie, R. C., Madden, R. J., and Parnell, J. C, (I97I) Chemistry of the Podocarpaceae
XXVil. Constituents of some Podocarpus and other species. Aust. J.Chem. 24,
2n -221.
Campbell, W. P. and Todd, D. (1942) The structure and configuration of resin acids.
Podocarpic acid and ferruginol. J. Am. Chem. Soc.64,928-935.
Carman, R. M., and Sutherland, M. D. (1979) Cupressene and other diterpanes of
Cupressus species. Aust. J. Chem.32, 1132-11'42
Carruthers, W. (1956) The constituents of high-boiling petroleum distillates. Part III.
Anthracene homologues in a Kuwait Oil. J. Chem. Soc.,603-607.
Carruthers, Vy'. and Douglas, A. G. (1961) I, 2-Benzanthracene derivatives in a Kuwait
mineral oil. Nature 192,256-257.
Carter, F. D., Copp, F.L., Sanjiva Rao, 8., Simonsen, J. L. and Subramaniam, K. S.
(1939) Constituents of some Indian essential oils. Part XXVI. The structures of 1-
cx- and B-curcumenes.J. Chem. Soc.,l5O4-1509
296
Chapter 9
297
Chapter 9
Cosgrove, J. L. (1987) South-west Queensland gas - a resource for the future. Aust. Pet.
Explor. J. 27 (I), 245-263.
Craig, H. (1953) The geochemistry of the stable carbon isotopes. Geochim. Cosmochim.
Acta.3, 53-92.
Craig, H. (1957) Isotopic standards for carbon and oxygen and correction factors for mass
spectrometric analysis of carbon dioxide. Geochim. Cosmochim. Acta t2, 133-
t49.
Cranwell P. 4., Eglinton G. and Robinson, N. (1987) Lipids of aquatic organisms as
298
Chapter 9
299
Chapter9
300
Chapter 9
Galimov, E. M. (1973) carbon Isotopes in oil-Gas Geology, Nedra, Moscow, 3g4 p.;
National Aeronautics and Space Administration, 'Washington D. C.
[Translation
from Russianl, 395 p..
Garrels, R. M. and Lerman, A. (1984) Coupling of sedimentary sulfur and carbon cycles -
an improved model. Am. J. Sci.2B4,9g9-1007.
Garrels, R. M', McKenzie, F. T., and Hunt, C. (1975) Chemical Cycles and the Gtobal
Environment. Katfmann, Los Altos.
Gatehouse, C. G. (1986) The geology of the V/arburton Basin in South Australia. Aust.
J.
Earth Sci.33, 161-180.
Gehman, H. M. (1962) Organic matter in limestones. Geochim. Cosmochim. Acta26, gg5-
897.
Geissman, T. A', Sim, K. Y. and Murdoch, J. (1967) Organic minerals. picene and
chrysene as constituents of the mineral curtisite (idrialite). Experientia
23,lg3-7g4.
Gelpi, E., Schneider, H., Mann, J., and oro, J. (1970) Hydrocarbons of geochemical
significance in microscopic algae. p hytochemistry 9, 603 -612.
Geological Survey Of Queensland, (1985) Eromanga Basin lithostratigraphic nomenclature.
Queensland Govt Min. J.86,446.
Gilby, A. R. and Foster, C. B. (1988) Early Permian palynology of Arckaringa Basin,
South Australia. P alaeonto graphica. Abt B : P alae ophytolo gie 209, 161 -Ig l.
Gilby, A. R., and Mortimore, I. R. (19S9) The prospects for Eromanga oil accumulations in
the northern Cooper Basin region, Australia. InThe Cooper and Eromanga Basins,
Australia (Edited by O'Neil, B. J.), pp. 391-403. Proceedings of the petroleum
Exploration Society of of Petroleum Engineers, Australian
Australia, Society
society of Exploration Geophysicists (sA Branches), Adelaide.
Goldhaber' M. B. and Kaplan, I. R. (1974) The marine sulphur cycle. In The
Sea @dited
by Goldberg, E. D.) vol. 5, pp. 569-655. John wiley and sons, New york.
Goodarzi, F. (1984) chitinous fragments in coal. Fue|63,1504-1507.
Goosens, H., De Leeuw, J. w., schenck, p.4., and Brassell, s. c. (19g4) Tocopherols as
likely precursors of pristane in ancient sediments crude oils. Nature 312,440-442.
Gough, L. L (1964) conifer resin constituents. chem. Ind.,2o59-2060.
Gould, R. E. (1975) The succcession of Australian pre-Tertiary megafossil floras.
Botanical
Rev.4l, 453-483.
Gould, R. and Shibaoka, M. (1980) Some aspects of the formation and petrographic
features of coal members in Australia with special reference to the Tasman
orogenic
zone. Aust. CoaI Geol. Z, I-29.
van Graas, G., de Leeuw, J. w., Schenck, p. A. and Haverkamptz,I.
J. (19g1) Kerogen
of Toarcian shales of Paris Basin. Geochim. Cosmochim. Acta 45,2465-247I.
van Graas, G., Baas, J. M. 4., de Graaf, v. and de Leeuw, J. w. (rggz)
Theoretical
organic geochemistry. I. of several cholestane isomers
Thermodynamic stability
calculated by molecular mechanics. Geochim. Cosmochim. Acta 46,1025-1032.
301
Chapter 9
Guennel, G. K. (1952) Fossil spores of the Alleghenian coals in Indiana. Indiana Geol.
Surv., Rept. Prog. 4, I-40.
Guy, R. D., Reid, D. M. and Krouse, H. R. (1986) Factors affecting
t3clt2c ratios of
inland halophytes. I. Controlled studies on growth and isotopic composition of
P uc c ine IIa hut alliana. C an. J. B ot. 64, 2693 -2699 .
302
Chapter 9
40, ll4r-r143.
ten Haven, H. L., Littke, R. And Rullkötter, J. (1992) Hydrocarbon biological markers in
Carboniferous coals of different maturities. ln Biological Markers in Sediments and
Petroleum (Edited by Moldowan, J. M., Albrecht, P. and Philp, R. P.), pp. 142-
Cliffs, N. J.
155. Prentice Hall, Englewood
Hayatsu, R., Winnans, R. E., Scott, R. G., Moore, L. P. and Studier, M. H. (1978)
Trapped organic compounds and aromatic units in coals. Fuel 57, 54I-548.
Hedberg, H. D. (1968) Significance of high wax oils with respect to genesis of petroleum.
Am. Assoc. Pet. GeoI. BuII. 52,136-750.
Heissler, D., Ocampo, R., Albrecht, P., Riehl, J. and Ourisson, G. (1984) Identihcation of
long-chain tricyclic terpene hydrocarbons (C21-C30) in geological samples. ,/.
Chem. Soc. Chem. Comm.,496-498.
Héroux, Y., Chagnon, A. and Bertrand, R. (1919) Compilation and correlation of major
thermal maturation indicators. Am. Assoc. Pet. GeoI. BulI. 63,2128-2144
Hites, R. 4., LaFlamme, R. E. and Fanington, J. W. (1977) Sedimentary polycyclic
aromatic hydrocarbons: the historical record. Science 198, 829-831.
Hodgson, G. W., Hitchon, 8., Taguchi, Baker, B. L. and Peake, E. (1968)
K.,
Geochemistry of porphyrins, chlorins and polycyclic aromatics in soils sediments
and sedimentary rocks . Geochim. Co smochim. Acta 32, 7 37 -77 2.
Hollingworth, R. J. S. (1989) The exploration history and status of the Cooper and
Eromanga Basins. In The Cooper and Eromnnga Basins, Australia (Edited by
O'Neil, B. J.), pp. 1-3. Proceedings of the Petroleum Exploration Society of
Australia, Society of Petroleum Engineers, Australian Society of Exploration
Geophysicists (SA Branches), Adelaide.
Holzer, G., Oro, J. and Tornabene, T. G. (1919) Gas chromatographic analysis of natural
lipids from methanogenic and thermoacidophilic bacteria. ,I. Chromatogr. 186,
795-809.
Hong, 2.H., Li, H. X., Rullköter, J. and Mackenzie, A. S. (1986) Geochemical
application of steranes and triterpane biological marker compounds in the Linyi
Basin. Org. Geochem. 10, 433-439.
Horstman, E. L. (1994) The effect of inertinite on kerogen appraisal by programmed
pyrolysis, North V/est Shelf, Australia. Aust. Pet. Explor. Assoc. J. 34(l), 291-
306.
Huang, V/. Y. and Meinschein, W. G. (1919) Sterols as ecological indicators. Geochim.
Cosmochim. Acta. 43, 739-7 45.
Hudson, J. D. (1982) Pyrite in ammonite-bearing shales from the Jurassic of England and
Germany. Sedimentolo gy 29, 639-667 .
Hughes, W. 8., Holba, A. G., Miller, D. E. and Richardson, J. S. (1985) Geochemistry of
In Geochemistry in Exploration of the Norwegian Shelf
greater Ekofisk crude oils.
(Edited by Thomas, B. M.), pp.75-92. Graham and Trotman, London.
303
Chapter 9
Hunt, J. W. (1988) Sedimentation rates and coal formations in the Permian basins of eastern
Australia. Aust. J. Earth ^Sci.
35, 259-214.
Hunt, J. M. and Jamieson, G. W. (1956) Oil and organic matter in source rocks of
petroleum. Am. Assoc. Pet. Geol. Bull.40,47l-488
Hussler, G., Chappe, B.,'Wehrung, P., Albrecht, P. (1981) Czt-Czs ring aromatic steroids
in Cretaceous black shales. Nature 294,556-558.
Hutton, A. C. and Cook, A. C. (1980) Influence of alginite on the reflectance of vitrinite
from Joadja, NS'W, and some other coals and oil shales containing alginite. Fuel
59, 7 rr-7 16.
Ikan, R., Aizenshtat, 2., Baedecker, M. J. and Kaplan, I. R. (1975) Thermal alteration
experiments on organic matter in recent marine sedimets. I. Pigments. Geochim.
Cosmochim. Acta 39, 173-185.
Ingram, L. L., Ellis, J., Crisp, P. T. and Cook, A. C. (1983) Comparative study of oil
shales and shale oils from the Mahogany zone Green River Formation (USA) and
Kerosene Creek seam, Rundle Formation (Australia) . Chem. Geol.38, 185-212.
International Committee for Coal Petrology (1975) International Handbook of CoaI
Petrography, second supplement., Centre National de la Recherché Scientific,
Paris.
Iyengar, M. S., and Lahir, A. (1959) Einige Beitrage zur Aufklarung der struktur und
konstitution der Kohle. Brennst-chem 40, 8-13.
Jackson, K. S., McKirdy, D. M. and Deckelman, J. A. (1984) Hydrocarbon generation in
the Amadeus Basin, central Australia. Aust. Pet. Explor. Assoc J.24(l),42-65.
Jacobson, S. R., Hatch, J. R., Teerman, S. C. and Askin, R. A. Middle Ordovician organic
matter assemblages and their effect on Ordovician-derived oils. Am. Assoc. Pet.
Geol. BuIl. 72, 1090-1100.
Jenkins, C. C. (1987) The organic geochemical correlation of crude oils from Triassic to
Late Cretaceous reservoirs of the Cooper and Eromanga Basins. M.Sc. thesis,
University of Adelaide.
Jenkins, C. C. (1989) Geochemical correlation of source rocks and crude oils from the
Cooper and Eromanga Basins. In The Cooper ønd Eromanga Basins, Australia
(Edited by O'Neil, B. J.), pp. 525-540. Proceedings of the Petroleum Explorarion
Society of Australia, Society of Petroleum Engineers, Australian Society of
Exploration Geophysicists (SA Branches), Adelaide.
Jones, R. W. and Edison, T. A. (1979) Microscopic of kerogen related to
observations
geochemical parameters with emphasis on thermal maturation.In Low Temperature
Metamorphisim of Kerogen and CIay Minerals (Edited by Oltz, D. F.), pp. L-I2.
Symposium in Geochemistry, Pacif,rc Section, Soc. Econ. Palaeont. Miner., Los
Angeles.
Kagya, M. N. (1993) The source rock and petroleum geochemistry of Poolowanna
Formation (Basal Jurassic), Eromanga Basin. Report to South Australia Department
304
Chapter 9
305
Chapter 9
306
Chapter 9
241-258.
Marchioni, D. L. (1916) Investigation into the palaeoenvironment of the Foybrook
Fotmation in the Hebden-Ravensworth area, Hunter Valley, New South Wales.
Ph.D. Thesis, University of Newcastle, NSW.
Martin, R.L.,'Winters, J. C. and Williams, J. A. (1963) Distribution of n-paraffins in crude
oils and their implications to origin of petroleum. Nature 199, 1190-1 193.
Marzi, R. and Rullkötter, J. (1992) Qualitative and quantitative evolution and kinetics of
biological marker transformations - Laboratory experiments and application to the
Michigan Basin. In Biological Markers in Sediments and Petroleum (Edited by
Moldowan, J. M., Albrecht, P. and Philp, R. P.), pp. 18-41. Prentice Hall,
Englewood Cliffs, N. J.
307
Chapter 9
McKirdy, D. M. (1981, 1983) Petroleum geochemistry and source rock potential of the
Arrowie, Pedirka, Cooper and Eromanga Basins, central Australia. Report for Delhi
Petroleum Pty Ltd (unpublished).
McKirdy, D. M. (1982) Aspects of the source rock and petroleum geochemistry of the
Eromanga Basin. In Eromanga Basin Symposium, Summary Papers (Compiled by
Moore, P. J. and Mount, T. J.), pp. 258-259. Geological Society of Australia and
Petroleum Exploration Society of Australia, Adelaide..
McKirdy, D. M. (1984a) Notes on the source affinity and thermal maturity of Eromanga
Basin crude oils. AMDEL Report W243184 for Delhi Petroleum Pty Ltd.
(unpublished).
McKirdy, D. M. (1984b) Source rock and petroleum geochemistry, Naccowlah Block,
southwest Queensland. AMDEL Report F6753184 for Delhi Petroleum Pty Ltd.
(unpublished).
McKirdy, D. M. (1985) Rapid evaluation of non-marine petroleum source rocks. National
Energy Research Development and Demonstration Program: Project 261. End-of-
grant technical report.
McKirdy, D. M., and Chivas, A. R. (1992) Nonbiodegraded aromatic condensate
associated with volcanic supercritical ca¡bon dioxide, Otway Basin: implications for
primary migration from terestrial organic matter. Org. Geochem.1.S,6II- 627.
McKirdy, D. M., Aldridge, A. K. and Ypma, P. J. M. (1983) A geochemical comparison of
some crude oils from pre-Ordovician carbonate rocks. In Advances in Organic
Geochemistry 1981 (Edited by Bjorgy, M. et aL),pp.99-107. v/iley, chichesrer.
McKirdy, D. M., Kantsler, A. J., Emmett, J. K. and Aldridge, A. K. (19s4) Hydrocarbon
genesis and organic facies in Cambrian carbonates of the Eastern Officer Basin,
South Australia. In
Petroleum Geochemistry and Source Rock Potential of
carbonnte Rocks (Edited by Palacas, J. G.), pp. 13-31. Studies in Geology 1g.
Am. Assoc. Pet. Geol., Tulsa.
McKirdy, D. M., cox, R. E., volkman, J. K., and Howell, v. J. (1986a) Botryococcane
in a new class of Australian non-marine crude oils. Nature 320, 57 -s9.
McKirdy, D.M., Emmett, J. K., Mooney, B. A., cox, R. E. and'watson, B. L. (19g6b)
Organic geochemical facies of the Cretaceous Bulldog Shale, western Eromanga
Basin, South Australia. Contributions to the Geology and Hydrocarbon Potential of
theEromangaBasin (Editedby Gravestock, D. I., Moore, p. S. and pitt, G. M.),
pp.287-304. Geol. Soc.Aust. Spec. Publ. 12.
Meinschein, W. G. (1961) Significance of hydrocarbons in sediments and petroleum.
Geochim. Cosmochim. Acta 22, 58-64.
Meischein, w. G. (1959) origin of petroleum. Am. Assoc. pet. Geol. BulI. 43, gz5-943.
Michaelis, W. and Albrecht, P. (1979) Molecular fossils of Archaebacteria in kerogen.
N at urw iss ens c haft en 66, 420 -421 .
Michaelsen, B. H. and McKirdy, D. M. (1989) Organic facies and petroleum geochemistry
308
Chapter 9
309
Chapter 9
311
Chapter 9
3t2
Chapter 9
3t3
Chapter 9
314
Chapter9
315
Chapter 9
316
Chapter 9
Spackman, W., Davis, A. and Mitchell, G. D. (1976) The fluorescence of liptinite macerals.
Brigham Young Univ. Geol. Stud.22(3),59-75. provo, Utah.
Stach, E. (1964) Zur Untersuchung de Sporinits in Kohlen-Anschliffen. Fortschr. GeoL
Rheinld. u. Westf. 12,403-420.
Stach, E. and Alpern, B. (1966) Inertodetrinit, Makrinit und Mikrinit. Fortschr Geol.
Rheinld. Westf. 13, 969-980.
Stach, E., Mackowsky, M.-Th., Teichmüller, M., Taylor, G. H., Chandra, D. and
Teichmüller, R. (1982) Stach's Textbook of CoaI Petrology (3rd edition). Gebrüder
Borntraeger, Berlin.
Stahl, W. J. (1977) Carbon and nitrogen isotopes in hydrocarbon research and exploration.
Chem. GeoL 20, l2I-149.
Stahl, V/. J. (1979) Carbon isotopes in petroleum geochemistry. In Lectures in Isotope
Geology (Edited by Jager, E. and Hunziker, J. c.), pp. 274-282. springer-verlag,
Berlin.
Stahl, V/. J. (1978) Source rock crude oil correlation by isotopic type-curves. Geochim.
Cosmochim. Acta 42, 1573-1577.
Standards Association of Australia (1986) Coal maceral analysis. Australian Standard AS
2856. The Standards House, North Sydney, N.S.W.
Stewart, W. N. (1983) Paleobotany and the Evolution of Plants.. Cambridge University
Press, Cambridge.
Stockey, R. A. (1982) The araucariaceae: An evolutionary perspective. Rev. Paleobot.
Palynol. 37, 133-154.
stopes, M. c. (1935) on the petrology of banded bituminous coal.
FueI14,4-13.
Strachan, M., Alexander, R. and Kagi, R. I. (1938) Trimethylnaphthalenes in crude oil and
sediments:Effects of source and maturity. Geochim. Cosmochim. Acta 52, 1255-
1264.
Streibl, M. and Herout, V. (1969) Terpenoids-especially oxygenated mono-, sesqui-, di-
and triterpenes. In Organic Geochemistry: Methods and Results. (Edited by
Eglinton, G. and Murphy, M. T. J.), pp. 4or-424. Springer-verlag, Berlin.
Stuiver, M. (1978) Atmospheric carbon dioxide and carbon reservoir changes. Science 199,
253-258.
Summons, R. E. and lValter, M. R. (1990)
Molecular fossils and microfossils of
prokaryotes and protists from Proterozoic sediments. Am. J. Sci. 290-A,,2lZ-244.
Summons, R. E. and Jahnke, L. L. (1992) Hopenes and hopanes methylated in ring A
Correlation of the hopanoids from extant methylotrophic bacteria with their fossil
analogues. In Biological Markers in Sediments and Petroleum (Edited, by
Moldowan, J. M., Albrecht, P. and philp, R. p.), pp. rg2-r94. prentice Hall,
Englewood Cliffs, N. J.
Summons, R. E. (1993) Biogeochemical cycles: A review of fundamental aspects of organic
matter formation, preservation and composition. In Organic Geochemistry:
317
Chapter 9
Principles and Applications (Edited by Engel, M. H. and Macko, S. A.), pp. 3-17 .
Plenum, New York.
Swan, E. P. (1965) Identity of a hydrocarbon found in a forest soil. Forest Product J.
t5(7),272-213.
Sweeney, R. E. and Kaplan, I. R. (1973) Pyrite framboid formation: Laboratory synthesis
and marine sediments. Econ. GeoI.68,618-634.
Tan, Y. L. and Hait, M. (1981) Biogenic and abiogenic polynuclear aromatic hydrocarbons
in sediments from two remote Adirondack lakes. Geochim. Cosmochim. Acta 45,
2267-2279.
Taylor, G. H. and Liu, S. Y. (1987) Biodegradation in coals and other organic-rich rocks.
Fuel 66, 1269-1273.
Taylor, G. H., Liu, S.Y. and Smyth, M. (1988) New light on the origin of Cooper Basin
o1l. Aust. Pet. Explor. Assoc. J. 28 (I),303-309.
Taylor, G. H. and Lui, S. Y. (1989) The maturation of liptinite. In Macerals'89 (Edited by
Thomas, c. G. and Strachan, M. G.), pp. 11/1-s. cslRo Division of coal
Technology, Sydney.
Taylor, J., Wardroper, A. M. K. and Maxwell, J. R. (1980) Extended hopane derivatives in
sediments: identification by lH NMR. Tet.rahedron Lett. 655-656.
Teichmüller, M. (1971) Anwendung Kohlenpetrographischer Methoden bei der Erdöl-und
Erdgasprospektion. Erdöl u. Kohle 24,69-76.
Teichmüller, M. (I974a) Uber neue Macerale der Liptinitgruppe und die Entstehung de
Mikrinits. Fortschr. GeoI. Rheinld. Westf. 24,37-64.
Teichmüller, M. (1974b) Generation of petroleum-like substances in coal seams as seen
under microscope.In Advances in Organic Geochemistry 1973 (Edited by Tissot,
B. and Bienner, F.), pp. 379-408. Éditions Technip, paris.
Teichmüller, M. (1982) Origin of the petrographic constituents of coal. In Stach's Textbook
of Coal Petrology,3rd edition , pp.219-294. Gebrüder Borntraeger.
Teichmüller, M. (1985) Organic petrology of source rocks, history and state of the art. Org.
Geochem.10, 581-599.
Thiessen, R., Sprunk, G, C. and o'Donnell, H, J. (1931) Microscopic study of Elkhorn
coal bed at Jenkins, Letcher County, Kentucky. U.S. Bur. Mines Tech. pap. 506,
1-30.
Thomas, B. M. (1982) Land plant source rocks for oil and their significance in Australian
basins. Aust. Pet. Explor. Assoc. J.22 (I), 164-178.
Thomas, B. R. (1969) Kauri resins - modern and fossil. In Organic Geochemistry: Methods
an"d Results (Edited by Eglinton, G. and Murphy, M. T. J.), pp. 599_61g.
Springer-Verlag, Berlin.
Thornton, R. C. N. (1919) Regional stratigraphic analysis of the Gidgealpa Group,
southern Cooper Basin, Australia. Geol. Surv. South Australia BuII. 49.
Tissier, M. J. and Saliot, A. (1933) Pyrolytic and naturally occurring polycyclic aromatic
318
Appendix
Reports and abstracts
Michaelsen, B. H., Kagya, M. L. N., McKirdy, D. M., Paull, R., and Yu, X. (1996)
Geochemical comparison of source rock and oils from the western Cooper Basin and
supedacent Eromanga Basin. Report for the Departrnent of Mines and Energy South
Australia.
Conference abstracts
Kagya, M. L. N., B. H. and McKirdy, D. M. (1995) Source rock and
Michaelsen,
petroleum geochemistry of the Early Jurassic Poolowanna Formation, western
Eromanga Basin. In Proceedings ol the 1995 Australian Organic Geochemistry
Conference, Abstracts (Compiled by Michaelsen, B. H. and McKirdy, D. M.), p'
40. University of Adelaide, South Australia
322
4th Conference on Petroleum Geochemistry and Exploration in the Afro-Asian
Region, Programme and Abstracts, pp. 63-64. Afro-Asian Association of Petroleum
Geoehemists, Arusha, T attzania.
323