You are on page 1of 21

Organic

Geochemistry
Organic Geochemistry 35 (2004) 937–957
www.elsevier.com/locate/orggeochem

Geochemistry of oil in fluid inclusions in a middle


Proterozoic igneous intrusion: implications for the source
of hydrocarbons in crystalline rocks
a,* b,1 c,2 b,1
A. Dutkiewicz , H. Volk , J. Ridley , S.C. George
a
School of Geosciences and Institute of Marine Science, University of Sydney, Sydney, NSW 2006, Australia
b
CSIRO Petroleum, P.O. Box 136, North Ryde, NSW 1670, Australia
c
Department of Earth and Planetary Sciences, Macquarie University, Sydney, NSW 2109, Australia
Received 18 December 2003; accepted 23 March 2004
(returned to author for revision 19 February 2004)
Available online 30 April 2004

Abstract

The ca. 1280 Ma dolerite sill within the Mesoproterozoic Roper Group in the Roper Superbasin, Australia,
contains evidence for at least two episodes of hydrocarbon migration represented, respectively, by solid bitumen
with a ketone-rich extract, and a mixture of a high maturity gas-condensate and a lower maturity oil within oil-
bearing fluid inclusions. The ketone isomers are formed by flash pyrolysis of kerogen during the intrusion of the
dolerite sill [Org. Geochem. 21 (1994) 829] and represent the first and oldest phase of hydrocarbon migration. The
gas condensate and oil were subsequently trapped as a mixture within fluid inclusions at diagenetic temperatures
and pressures of around 110 °C and 250 bars, significantly after cooling of the sill and likely during the Neo-
proterozoic reactivation of the Roper Superbasin. Either (1) these fluids migrated together and mixed in the res-
ervoir or (2) an earlier oil charge was flushed by a later condensate charge and the oil-condensate mixture was
trapped within single fluid inclusions. Oil inclusions occur chiefly within albitised zones of labradorite laths within
the dolerite matrix, and within transcrystalline microfractures cutting vein calcite and rarely vein quartz. Oil in-
clusions trapped in the vein calcite are accompanied by hypersaline Ca/Mg brines. Gas chromatography–mass
spectrometry of oil extracted from inclusions within the dolerite matrix shows that the oil is non-biodegraded and
was therefore trapped relatively quickly within the host minerals. Trace amounts of biomarkers indicate that the
inclusion oil is of a biogenic origin and excludes any abiotic processes that are apparent sources of hydrocarbons in
many crystalline rocks. Monomethylalkanes, pentacyclic terpanes chiefly comprising hopanes and diahopanes, and
very low concentrations of steranes and diasteranes indicate input from cyanobacterial organic matter with a minor
contribution from eukaryotes. The hydrocarbons are likely derived from Proterozoic source rocks such as the
directly overlying Velkerri Formation and/or the underlying Barney Creek Formation from the McArthur Group.
The study has implications for the source of hydrocarbons in non-sedimentary rocks and suggests that fluid in-
clusions can be used not only to distinguish between biogenic and abiogenic sources of hydrocarbons but also to

*
Corresponding author. Fax: +61-2-9351-0184.
E-mail addresses: adriana@geosci.usyd.edu.au (A. Dutkiewicz), Herbert.Volk@csiro.au (H. Volk), John.Ridley@geo.unibe.ch (J.
Ridley), Simon.George@csiro.au (S.C. George).
1
Fax: +61-2-94908197.
2
Present address: Institute of Geological Sciences, University of Bern, Baltzerstrasse 1-3, CH-3012, Bern, Switzerland. Fax: +41-31-
631-4843.

0146-6380/$ - see front matter Ó 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.orggeochem.2004.03.007
938 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

provide critical information about the composition of the biomass, fluid migration events, and conditions under
which hydrocarbons in different geological settings are trapped.
Ó 2004 Elsevier Ltd. All rights reserved.

Keywords: Precambrian oil; Oil inclusions; Biomarkers; Roper Group; Roper Superbasin; Dolerite

1. Introduction allows distinction between biogenic and abiogenic


sources, but can give important information about the
Where igneous and metamorphic rocks host methane ancient biomass and about oil migration. It has re-
and higher-order hydrocarbons, their exact origin in cently been documented that some ancient oils have
many instances remains controversial. The possibility of been preserved for billions of years within fluid in-
an abiogenic origin has sometimes been evoked, for clusions in Archaean and Proterozoic sandstones and,
example, through post-magmatic Fischer–Tropsch re- although low in abundance, are sufficiently pristine to
actions (e.g., Szatmari, 1989; Ettner et al., 1996; Salvi yield biomarkers and information about the timing
and Williams-Jones, 1997; Potter et al., 1998), outgas- and conditions of entrapment (Dutkiewicz et al., 1998,
sing of the primordial mantle (e.g., Gold, 2001) or res- 2003a). Furthermore, survival of complex hydrocarbons
peciation of a primary carbonic fluid (e.g., Kogarko is evident within fluid inclusions at temperatures up to
et al., 1987). Notable occurrences which have been ex- 300 °C (Dutkiewicz et al., 2003b). Therefore, preserva-
plained through abiogenic origins include: (1) hydro- tion of hydrocarbons in fluid inclusions should be pos-
carbon gas inclusions in Ilimaussaq igneous complex in sible in crystalline rocks, a large proportion of which are
southern Greenland (Konnerup-Madsen et al., 1979), (2) of Proterozoic age.
methane and higher hydrocarbons in granite and dol- Here, we report on the discovery and composition of
erite intrusions at Siljan, Sweden (Jeffrey and Kaplan, abundant oil inclusions in a Middle Proterozoic dolerite
1988), (3) hydrocarbon gas seeps in the Zambales sill that was intruded into a sedimentary sequence
Ophiolite, Philippines (Abrajano et al., 1988), (4) hy- known to contain oil shows such as bitumen and oil
drocarbon inclusions containing saturated and unsatu- seeps (e.g., George et al., 1994). We use a combination
rated aliphatics in pegmatite quartz in granite at Strange of inclusion petrography, microthermometry and mo-
Lake, Canada (Salvi and Williams-Jones, 1997), (5) lecular characterisation, including biomarker analysis of
methane-bearing fluid inclusions in alkaline igneous in- the inclusion oils to demonstrate that this magmatic
trusions of the Kola igneous province (Potter et al., rock hosted oil is biogenic and was trapped on migration
1998) and (6) methane and higher hydrocarbons in through the sill after cooling. We further use biomarkers
fractures in crystalline rocks of the Canadian shield to constrain the source rocks of the oil and the nature of
(Sherwood Lollar et al., 2002). In contrast, a widespread the source biomass and suggest that a similar approach
occurrence of bitumen in basement rocks in Scandinavia can be used to distinguish between biogenic and abio-
and gas-condensate inclusions with alkane homologues genic sources of hydrocarbons, particularly in rocks
up to C15 in the crystalline basement in South Norway which no longer contain oil or gas in the pore spaces or
(Munz et al., 2002) and numerous complex hydrocar- have no obvious association with conventional source
bons in granite plutons in the British Isles (Parnell, 1988) rocks or a fracture network. Our fluid inclusion ap-
have been attributed to overlying sedimentary source proach can also provide a better understanding of pe-
rocks and have been derived from the thermal decom- troleum systems associated with crystalline rocks,
position of organic matter. Solid bitumens within a particularly in terms of the timing, entrapment condi-
dolerite dyke at Oklo in Gabon (Mossman et al., 2001) tions and migration pathways, which are still poorly
have similarly been interpreted to be the products of oil understood.
migration from surrounding sedimentary rocks. A
comprehensive list of hydrocarbon occurrences associ-
ated with igneous rocks can be found in Schutter 2. Geological setting
(2003a).
Extractable organic matter and oil are relatively The marine sedimentary succession of the ca. 1430
uncommon in early Precambrian rocks due to low Ma Roper Group (Jackson et al., 1986; Powell et al.,
concentrations of total organic carbon in ancient 1987) from the intracratonic Roper Superbasin, Aus-
source rocks, breakdown of hydrocarbons under tralia (Fig. 1) is one of the oldest and least deformed and
metamorphic conditions and loss of hydrocarbons least metamorphosed hydrocarbon-bearing Proterozoic
over time (e.g., Hunt, 1996). However, where present sequences in the world. This study investigates a dolerite
their molecular and isotopic composition not only sill within the Roper Group sequence, which is likely
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 939

Fig. 1. (a) Geological map of the Roper Superbasin showing the location of the Friendship-1 and Borrowdale-2 wells (after Pietsch
et al., 1991). (b) Stratigraphy of the Roper Basin and the Friendship-1 core (after Summons et al., 1988). Absolute ages are from
McDougall et al. (1965), Kralik (1982), Jackson et al. (1999), Jackson et al. (2000) and Page and Sweet (1998).
940 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

part of a widespread intrusive suite dated at 1280 Ma Burial history analysis of the Roper Group (Jackson
(McDougall et al., 1965) emplaced into pre-existing et al., 1988) suggests that the Velkerri Formation
fractures during an extension event. The sill is directly reached its maximum burial at around 1400 Ma, with
overlain by Velkerri Formation shales with total organic estimated burial temperatures of about 75 °C. Soon after
carbon (TOC) levels of up to 8% (Jackson and Raiswell, burial the sequence experienced steady uplift until about
1991; Warren et al., 1998) and is underlain by Bessie 500 Ma, followed by a period of relative stability
Creek Sandstone (Fig. 1) where a distinct contact punctuated by four minor uplift and burial events in the
metamorphic fabric (‘‘baking’’ texture) induced by the Phanerozoic (Jackson et al., 1988).
intrusion is evident (Dutkiewicz et al., 2003a). Oil shows
and bitumen have been observed in the Bessie Creek
Sandstone and in the dolerite (Ledlie and Torkington, 3. Materials and methods
1988; Ledlie et al., 1988; George et al., 1994), and live oil
has been observed to seep out of the marine Velkerri The dolerite sill was sampled from the drill core
Formation shale (Powell et al., 1987), which has been Friendship-1, collected in 1988 by Pacific Oil & Gas Pty
the subject of several geochemical studies (Jackson et al., Ltd. to test the petroleum prospectivity of the Upper
1986; Summons et al., 1988; Warren et al., 1998; George Roper Group (Fig. 1). Friendship-1 is located about 30
and Ahmed, 2002). Solid bitumens in the sill form thin km NNE of Borrowdale-2, which was the subject of an
linings to fracture surfaces and amorphous coverings to, earlier fluid inclusion study (Dutkiewicz et al., 2003a).
and intimate mixtures with, host minerals of the fracture Both wells penetrated complete sections of the Bessie
walls. They are non-fluorescing under UV excitation Creek Sandstone, the dolerite sill and the Velkerri For-
(George et al., 1994). Hydrocarbon extracts of a fracture mation. The dolerite sill at Friendship-1 was penetrated
hosting this solid bitumen show the presence of an un- between 262.5 and 362.2 mKB (metres below kelly
usual suite of saturated ketone isomers formed most bushing), and between 434.5 and 526.47 mKB at Bor-
likely by flash pyrolysis during intrusion of the dolerite rowdale-2. Correlations between Friendship-1 and Bor-
sill (George and Jardine, 1994). The extract was rich in rowdale-2 are based on consistent stratigraphic
polar compounds with n-alkanes only 3–4 times more relationships, lithology and petrography.
abundant than the ketones (George and Jardine, 1994). Eight doubly polished, 100-lm thick sections of the
Different extract fractions, including asphaltenes, polar dolerite sill from depths of 272.2–272.4 mKB and two
compounds, aromatic hydrocarbons and aliphatic hy- ordinary thin sections were examined using an optical
drocarbons, all have d13 C values between )31‰ and microscope equipped with transmitted, reflected and
)33‰ (George et al., 1994) and fall within the range epifluorescent light, by scanning electron microscopy
expected for organic matter derived from kerogens (Philips SEM 505 and Philips SEM XL30) including
(Lewan, 1986) and crude oils generated from organic backscattered electron-dispersive spectroscopy (EDS)
matter (Yeh and Epstein, 1981). No results of biomarker and X-ray mapping. Microthermometry of 220 oil and
analyses were published in George et al. (1994). Further aqueous fluid inclusions was carried out on a Linkham
hydrocarbon shows include oil inclusions in post-meta- heating-freezing stage calibrated using synthetic H2 O
morphic quartz-filled micro-fractures in the Bessie Creek and CO2 inclusions. In order to avoid stretching and
Sandstone (Dutkiewicz et al., 2003a; Volk et al., 2003). leakage of fluid inclusions caused by expansion of ice,
These were trapped at relatively low temperatures of heating runs were conducted before the cooling runs as
60 °C and contain a range of biomarkers derived recommended by Goldstein and Reynolds (1994).
mainly from prokaryotic organic matter. In order to specifically target fluid inclusion oil and
Deformation of the Roper Group is mild on a re- its molecular markers, a fresh sample of the dolerite with
gional scale (Rawlings, 1999). Structural inversion post- no calcite veins or solid bitumen was selected from ap-
dates an east–west extension that was accompanied by proximately the same depth as the one used in the
the intrusion of the dolerite sills and pre-dates devel- George and Jardine (1994) study. Determination of the
opment of the Neoproterozoic to Palaeozoic Arafura molecular composition of oil inclusions was carried out
Basin (Powell et al., 1987; Haines et al., 1999; Rawlings, using the methods similar to those described in George
1999). The inversion event may be related to the as- et al. (1998). A sub-sample was lightly crushed into
sembly of Rodinia ca. 1300–1000 Ma (Dalziel, 1992). granule and very coarse sand-sized fractions and cleaned
The deformation resulted in structural reactivation, in hydrogen peroxide, hot Aqua Regia and hot chromic
wrench-faulting, steepening of earlier faults and mild acid. This treatment removed any calcite that was
folding. The region has been essentially stable, experi- present initially, but most of the dolerite matrix minerals
encing slow erosion since deformation (Haines et al., were preserved during this procedure, including feldspar
1999) and much of the present structural grain in the and quartz that host oil inclusions. The cleaned fraction
region is due to the Mesoproterozoic inversion event (17.5 g) was then crushed under dichloromethane (off-
(Sweet et al., 1999). line crushing) and the extracted fluid inclusion oil, which
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 941

was comprised of the co-mingled contents of fluid in- 1 cm thick, wavy, varicated and variegated, and over-
clusions from the dolerite matrix minerals, was analysed prints patches of spilitic altered dolerite. It is rimmed by
by gas chromatography–mass spectrometry (GC–MS) a 2–3 mm thick, pink-coloured zone of bleached wall-
on a Hewlett–Packard 5890 gas chromatograph inter- rock composed of clay-dusted alkali feldspar, quartz
faced to a VG AutospecQ Ultima mass spectrometer. and fine hematite plates. The vein walls are largely
Single ion monitoring (SIM), metastable reaction mon- coated with zoned drusy euhedral quartz crystals that
itoring (MRM) and magnet scanning modes of detection grew into open- or fluid-filled space from nuclei on the
were used. Before the inclusion fluids were extracted by walls. Inclusions outline up to four growth zones al-
off-line crushing, system blank experiments using the though crystals with a single growth zone are most
same experimental conditions and glassware (including common (Fig. 3). The youngest growth zone of quartz is
the same amount of solvent) were carried out in order to in all cases more irregular and is either intergrown with
determine the extent of any hydrocarbon contribution the coarse calcite vein infill or is embayed by calcite
from the crushing cylinder and any carry-over from (Fig. 3). Vein calcite infill is coarse-grained and is in-
previously analysed samples or general laboratory tergrown with fine chlorite. Calcite deformation twins,
background. System blanks that were deemed to be planar microfaults and chlorite-lined stylolitic seams
sufficiently clean were acquired one day prior to crush- indicate one or more phase of deformation after vein
ing of the dolerite fraction, and the sample was subse- infill (Fig. 3).
quently crushed using exactly the same glassware. Low Three types of fluid inclusions are recognised in the
molecular weight compounds are not retained during dolerite as determined by petrography and microther-
sample workup using the off-line crushing method. mometry. Most of the fluorescing oil-bearing fluid in-
Therefore, the molecular composition of low molecular clusions are hosted within just two minerals: vein calcite
compounds was assessed using a direct on-line crushing and plagioclase crystals (Fig. 3).
method in a Quantum MSSV-1 Thermal Analysis Sys-
tem inlet mounted onto the same GC–MS system that 4.1. Type 1
was used to analyse the liquid off-line extracts. Liberated
compounds were thermally extracted in a helium flow at Fluid inclusions that form clusters within the core of
300 °C and focussed in a cryogenic trap prior to GC and along growth faces of drusy vein quartz (Fig. 3(d)).
separation (George et al., 1998). Whereas the off-line These inclusions are aqueous and no primary fluorescent
crushing method is advantageous for the analysis of oil-bearing inclusions occur within the growth zones. All
traces of C12þ hydrocarbons such as biomarkers, the on- inclusions are extremely small, usually 5 lm in longest
line method yields complementary information on the dimension and rarely exceed 10 lm. Most contain water
gasoline range hydrocarbon distribution of fluid inclu- and gas and occasionally an unidentified and likely ac-
sion oils (e.g., Ruble et al., 1998; Volk et al., 2002). On- cidental solid phase. The gaseous phase usually com-
line crushing coupled with gas chromatography has been prises 10% of the total inclusion volume; however, the
successfully used to examine the gasoline range hydro- gas content varies from 0 vol% (single phase inclusions)
carbons in fluid inclusions in Phanerozoic (Jensenius to  90 vol%. Some of the gas bubbles are mobile at
and Burruss, 1990; Wiggins et al., 1993; Tseng et al., room temperature.
1999; Volk et al., 2000, 2003) and Middle Proterozoic
rocks (Newell et al., 1993; Mauk and Burruss, 2002). 4.2. Type 2

Fluid inclusions within extensive, sub-parallel trans-


4. Oil Inclusions in the dolerite sill crystalline microfractures cutting the calcite crystals,
calcite twins and rarely euhedral quartz crystals bor-
The dolerite sill is generally unaltered. Fresh dolerite dering the vein (Fig. 3(c)–(h)). These trails cut the
is composed of ophitic augite, labradorite laths, ilmen- growth zones and hence post-date entrapment of Type 1
ite, quartz, minor apatite, andesine and scattered pat- inclusions. They include both oil inclusions (Type 2a)
ches of granophyric quartz–K-feldspar intergrowth. and coeval aqueous inclusions (Type 2b). Type 2a fluid
Pyroxene is rimmed by hornblende in some sections of inclusions contain oil and gas and occasionally water
rock. At least two phases of hydrothermal alteration are (up to 15 vol%) and/or bitumen. They are irregular,
recognised. Incipient to pervasive spilitic alteration is oval, spherical, sub-spherical, rectangular or square in
marked by chlorite, leucoxene and recrystallisation of shape and are notably larger (around 10–20 lm in long
primary plagioclase and granophyric K-feldspar to al- dimension) than oil-bearing fluid inclusions in the pla-
bite. Incipient alteration is often marked by patchy re- gioclase. Virtually all of these oil-bearing fluid inclusions
placement of labradorite plagioclase by albite (Fig. 2). A have blue fluorescence (yellow-fluorescing individuals
near-vertical quartz–calcite–chlorite vein is intersected are rare) and occasionally have a yellow tinge in trans-
approximately 10 m below the sill top. The vein is up to mitted light. Coeval Type 2b aqueous fluid inclusions
942 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

Fig. 2. Photomicrograph and elemental maps of altered labradorite in Friendship-1 dolerite. Bright areas indicate enrichment in a
particular element, dark areas indicate depletion. (a) Abundant blue-fluorescing oil-bearing fluid inclusions within alteration zones in a
labradorite lath. White arrow indicates the edge of a triangle-shaped alteration zone and serves as a reference point for backscattered
SEM and elemental map images. UV excitation. (b) Backscattered electron image showing alteration zones within labradorite lath.
White arrow indicates the same location as in (a). (c) Elemental map showing distribution of Ca. Note that altered zones are poor in
Ca. (d) Elemental map showing distribution of Al. Altered zone contains slightly less Al than the rest of the crystal. (e) Elemental map
showing distribution of K. The labradorite crystal and its alteration zones are poor in K. Surrounding minerals show a granophyric
texture with K-rich areas most likely due to K-feldspar intergrowth. (f) Elemental map showing distribution of Mg. A fracture cutting
the length of the labradorite lath appears to be rich in Mg. The labradorite alteration zones are poor in Mg. (g) Elemental map showing
distribution of Na. Note that the alteration zones in the labradorite crystal are relatively rich in Na. (h) Elemental map showing Si
distribution. Labradorite alteration zones show a slight enrichment compared to the rest of the crystal. The brightest, Si-rich area is
probably quartz most of which forms a granophyric texture with intergrown K-feldspar (lower right).
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 943

Fig. 3. Photomicrographs of Friendship-1 dolerite fluid inclusions. (a,b,c,e,g) UV-epifluorescence; (d,f,h) transmitted light. (a,b)
Clusters of small oil-bearing fluid inclusions within alteration zones in labradorite laths of the dolerite matrix. The majority of oil
inclusions have blue fluorescence. (c,d) Vein calcite (Cal) post-dating euhedral quartz (Q) and its overgrowth. Note the presence of
chlorite (Chl) within the calcite crystals. Calcite and quartz crystals are cut by a microfracture containing blue-fluorescing oil-bearing
fluid inclusions. In addition, the quartz crystal contains abundant tiny primary non-fluorescing inclusions which are dominantly
aqueous in composition. (e,f) Abundant blue-fluorescing oil inclusions within transcrystalline microfractures cutting calcite crystals
(Cal), calcite deformation twins and part of the quartz crystal (Q). (g,h) Euhedral quartz crystals (Q) lining the wall of the vein and
containing abundant aqueous fluid inclusions but no oil-bearing fluid inclusions. Quartz is post-dated by vein calcite (Cal) which is in
turn post-dated by stylolitic vein chlorite. Microfractures containing abundant blue-fluorescing oil-bearing fluid inclusions cut the vein
calcite and chlorite.
944 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

are usually sub-rectangular to rectangular in shape and 5.1. Type 1 inclusions


about 10 lm in longest dimension and contain water and
gas (water vapour?). The vein is also known to contain Because of their extremely small size, only 40 fluid
fracture-hosted solid bitumen (George et al., 1994) but it inclusions were suitable for microthermometric mea-
was not apparent in our samples and may therefore be surements. All of these inclusions were located in the
localised in occurrence. main growth zone of vein quartz and are interpreted to
have been trapped during the growth of the quartz
4.3. Type 3 crystals. TH values for aqueous fluid inclusions in the
vein quartz fall in a wide range from 95 to >160 °C
Fluid inclusions within clusters and discontinuous (Fig. 4), the latter being the maximum for our heating
trails of up to 30 individuals within matrix altered zones experiments due to leakage of some inclusions at that
of labradorite laths (Figs. 2(a) and 3(a) and (b)) and temperature. Quantitative statistics on this relatively
rarely andesine, augite and quartz. Hydrocarbon inclu- small data set are of limited value, as the maximum
sions in the plagioclase laths are very small (10 lm or less) homogenisation temperature has not been measured and
and contain oil and generally between 10–40 vol% gas. it is possible that the distribution contains more than
Gas bubbles are frequently mobile at room temperature. one mode. Most of the aqueous fluid inclusions showed
Individual clusters often comprise oil inclusions with persistent metastability on heating from a supercooled
blue, yellow and occasionally orange fluorescence col- state, particularly for inclusions which homogenised at
ours, although clusters with a single fluorescence colour relatively low temperatures, with final ice melting taking
(e.g., blue or yellow) also occur. Poor sealing properties place above 0 °C. For stable inclusions, however, ice
of the host feldspar within some alteration zones may melting temperatures are quite high and define a uni-
have resulted in a fluorescence colour change (e.g., Munz, modal distribution ranging from )1.3 to )0.3 °C
2001). The characteristics of the yellow and orange (Fig. 4). These temperatures correspond to salinities of
fluorescing inclusions, including the more variable in- 0.5–2.2 wt% NaClequiv: (Oakes et al., 1990) and indicate a
clusion sizes, ragged shapes similar to those described in slightly brackish waters much different from the hy-
Munz (2001) and variable gas-oil ratios, suggest that persaline brine associated with subsequent vein calcite
these are more likely than the blue fluorescent individuals fracturing and oil migration.
to have undergone leakage, stretching or modification
due to imperfect sealing within the mineral host. The near 5.2. Type 2 inclusions
uniform blue fluorescence colour of inclusions in micro-
fractures in more robust minerals such as vein calcite and A total of 127 coeval oil and aqueous fluid inclusions
quartz supports this interpretation. X-ray mapping of was measured in transcrystalline microfractures in the
several labradorite laths reveals that the oil inclusions are vein calcite. Type 2a oil inclusions homogenised between
generally restricted to zones which are relatively poor in 33 and 124 °C, while the Type 2b aqueous inclusions
calcium and aluminium and rich in sodium and silica homogenised at the expected higher temperatures (e.g.,
(Fig. 2), indicating albite composition. Rare, small (10 Munz, 2001) between 74 and 135 °C (Fig. 4). After su-
lm or less) coeval aqueous inclusions are associated with percooling to about )90 °C, the first melt in the aqueous
the oil-bearing inclusions; however, most of these aque- fluid inclusions generally occurred around )62 °C, sug-
ous inclusions appear to define different fluid inclusion gesting a calcium and/or magnesium-rich brine. Ice
trails and are hence interpreted to be unrelated to the oil melting temperatures (TM;ice ) show a distinct bimodal
migration event. Although feldspars tend to be water- distribution (Fig. 4) with the first mean of )30  1 °C
wet, roughened and dissolved surfaces are often oil-wet corresponding to a salinity of 26 wt% NaClequiv: and the
(Fassifihri et al., 1992) and will preferentially trap hy- second mean of )18  0.5 °C corresponding to a salinity
drocarbon phases. of 20 wt% and NaClequiv: (Oakes et al., 1990).

5.3. Type 3 inclusions


5. Fluid inclusion microthermometry
In view of the likelihood of post-entrapment modi-
All of the measured oil and aqueous fluid inclusions fication, most microthermometric measurements on
in the dolerite homogenise to a liquid and were thus Type 3 inclusions in alteration zones in the feldspar were
trapped in the liquid phase. There is a broad range of made on blue-fluorescing oil inclusions of similar size
overlapping homogenisation temperatures (TH ) for the and phase ratios which form small clusters of individuals
three inclusion populations (Fig. 4), suggesting that en- and appear to have been adequately sealed within the
trapment of these fluids in the different host minerals feldspar crystals. TH values of these inclusions range
occurred at similar temperatures, well within the limits from 56 to 127 °C and show a large scatter (Fig. 4). The
of diagenesis. range may be the result of imperfect sealing, stretching,
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 945

10 >160 35

9 Type 1 Aqueous Inclusions


Type 1 Aqueous Inclusions 30 in Vein Quartz
8 in Vein Quartz
7 25

Frequency
6
Frequency

20
5
15
4

3 10
2
5
1

0 0
0 20 40 60 80 100 120 140 160 180 200 -35 -30 -25 -20 -15 -10 -5 0
(a) TH (oC) (b) TM,ice (oC)

25 6

Type 2a Oil Inclusions Type 2b Aqueous Inclusions


5
20 in Vein Calcite in Vein Calcite

4
Frequency
Frequency

15

10
2

5
1

0 0
0 20 40 60 80 100 120 140 160 0 20 40 60 80 100 120 140 160
(c) TH (oC) (d) TH (oC)

7 9

Type 2b Aqueous Inclusions 8 Type 3 Oil Inclusions


6
in Vein Calcite in Feldspar
7
5
6
Frequency

Frequency

4 5

3 4

3
2
2
1
1

0 0
-35 -30 -25 -20 -15 -10 -5 0 0 20 40 60 80 100 120 140 160
(e) TM,ice (oC) (f) TH (oC)

Fig. 4. Histograms of microthermometric analyses on fluid inclusions in Friendship-1 dolerite. TH , homogenisation temperature; TM;ice ,
melting temperature of ice. All inclusions homogenised to a liquid. Inclusion types are discussed in the text.

heterogeneous trapping, necking down or simply trap- and averaging about 25 °C are determined. The repro-
ping at different temperatures. The last explanation ducibility of TH measurements indicates that the range of
seems ruled out as the modes are not apparent in the TH values is not caused by leakage or stretching of the
measurements of individual clusters, for which ranges of inclusions during the heating and cooling experiments.
homogenisation temperature of between 10 and 40 °C If stretching or leakage did occur in the geological
946 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

environment, then the relatively restricted range of precaution any potential calcite-hosted inclusions were
temperatures indicates that the TH values are relatively removed from the sample by treatment with acids, so the
small overestimates of the actual trapping conditions, inclusion oil analysed in this study was derived only
and that the trapping temperatures would have fallen from mainly feldspar and rare andesine, augite and
within the diagenetic range of temperatures. quartz oil-bearing fluid inclusions in the dolerite matrix.
The off-line fluid inclusion oil extract is extremely rich,
yielding more than 464 times the amount of n-alkanes (587
6. Molecular composition of inclusion oil ng C12 –C32 n-alkanes/g minerals crushed) compared to the
system blank and the final outside rinse (Fig. 5(A)). Low
The fluid inclusion petrography shows that the frac- molecular weight n-alkanes (C8 –C13 ) are much more
ture hydrocarbon extract analysed by George and Jar- abundant than C15–32 n-alkanes, indicating a dominant
dine (1994) would have included solid bitumen and condensate composition in the inclusions. This is con-
inclusion oil hosted in fracture calcite. In order to sim- firmed by the results from the on-line crushing (Fig. 6),
plify interpretation of the geochemical data, a sample which show high amounts of C5 –C9 hydrocarbons, espe-
lacking calcite-quartz veining was chosen. However, as a cially n-alkanes, with a maximum at n-C6 . Other abun-

Fig. 5. Partial mass chromatograms (A–H) for the Friendship-1 fluid inclusion oil and the corresponding system blanks, which were
drawn to the same scale as the fluid inclusion oil chromatograms using the internal standard. n-Alkanes and the internal standard
(squalane) are shown in the m=z 85 mass chromatogram of the inclusion oil (A) and the corresponding system blank (B). Total ion
chromatograms (not shown) are very similar to the m=z 85 partial mass chromatograms. Monomethylalkanes are expanded in the
partial m=z 85 mass chromatogram (C); MN, methylnonanes; MD, methyldecanes; MUD, methylundecanes; MDD, methyldodecanes
and MTD, methyltridecanes. (D) Pentacyclic triterpane distribution in the inclusion oil (m=z ¼ 191:2); Ts, C27 18a(H),22,29,30-tris-
norneohopane; Tm, C27 17a(H),22,29,30-trisnorhopane; C29 ab, 17a(H),21b(H)-30-norhopane; C29 Ts, 18a(H)-30-norneohopane; C30 *,
C30 17a(H)-diahopane; C29 ba, 17b(H),21a(H)-30-norhopane; C30 ab, 17a(H),21b (H)-hopane; C30 ba, 17b(H),21a(H)-hopane; C31–33 ab
22S and R, 17a(H),21b (H)-homo-, bishomo-, and trishomohopane (22S and R). (E) System blank corresponding to the pentacyclic
triterpane distribution. (F) Alkylnaphthalenes in the inclusion oil (added m=z ¼ 142:1 þ 156:1 þ 170:1); MN, methylnaphthalenes; EN
and DMN, ethyl- and dimethylnaphthalenes; and TMN, trimethylnaphthalenes. (G) System blank corresponding to alkylnaphthalene
distribution. (H) Alkylphenanthrenes in the inclusion oil (m=z ¼ 178:1, 192.1 and 206.1) and corresponding system blanks (below); P,
Phenanthrene; MP, methylphenanthrenes; EP and DMP, ethyl- and dimethylphenanthrenes. Chromatograms under respective m=z
traces show system blank scaled by using the internal standard.
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 947

Fig. 6. Total ion chromatogram (TIC), (A) and partial mass chromatograms (B–D) from the on-line crushing of the Friendship-1 fluid
inclusion oil sample (C5 –C9 range). (B) m=z ¼ 57 partial mass chromatogram, showing mainly normal alkanes (nCx) and some
branched alkanes. (C) m=z ¼ 55 partial mass chromatogram, emphasising the cyclic alkanes. (D) Added m=z ¼ 78 þ 91 partial mass
chromatogram, showing benzene and toluene. Other peak abbreviations: iC5, iso-pentane; MP, methylpentane; MCP, methylcyclo-
pentane; MH, methylhexane; DMP, dimethylpentane; DMCP, dimethylcyclopentane; MCH, methylcyclohexane; MHe, methylhep-
tane; DMCH, dimethylcyclohexane.

dant hydrocarbons in the m=z 85 partial mass chromato- the corresponding system blanks (Fig. 5). The C2 –C4
gram are monomethylalkanes, particularly methylnon- alkylbenzenes (87%) are significantly more abundant
anes, methyldecanes, methylundecanes (MUD) and than C1 –C5 alkylnaphthalenes (9.3%), C1 –C3 alkylphe-
methyldodecanes (MDD; Fig. 5(A)). Expansion of the nanthrenes (3.1%), C1 –C2 alkylbiphenyls (0.5%) and C1 –
m=z 85 mass chromatogram in the range of MUD and C2 alkyldibenzothiophenes (0.1%). This very marked low
MDD (Fig. 5(B)) shows that mid-chain methylated iso- molecular weight predominance in the distribution of
mers (e.g., 4-, 5- and 6-MUD) are of similar abundance as aromatic compound classes corroborates low-molecular
the terminal isomers such as 2- and 3-MUD. weight dominated n-alkane and n-alkylcyclohexane dis-
Alkylcyclohexanes and methylalkylcyclohexanes are tributions and is consistent with a gas condensate com-
abundant, and their abundance maximises at C12 and position of at least some of the analysed oil inclusions.
C11 , respectively, and attenuates rapidly with increasing The low content of dibenzothiophene and alkyldibenzo-
carbon number. This is consistent with observations of thiophenes relative to alkylphenanthrenes (Table 2) is
Hoffmann et al. (1987) who showed that alkylcyclo- consistent with a provenance from a low sulphur source
hexane profiles typically follow n-alkane profiles. Care- rock. In the C7 hydrocarbons of the Friendship-1 inclu-
ful examination of magnet scan data failed to detect any sion oil, there is a strong preference for isoalkanes (3-RP),
ketones in the inclusion oil. The lack of a hump of an with low alkylcyclopentanes (5-RP) and toluene + meth-
unresolved complex mixture in the m=z 85 mass chro- ylcyclohexane (6-RP). Based on the interpretative scheme
matogram and the presence of n-alkanes is good evi- of ten Haven (1996), this is strong evidence against any
dence that unlike the overlying Velkerri Formation terrestrial component and may be indicative of a lacus-
(Warren et al., 1998) and most solid bitumens with the trine depositional environment.
Roper Group (George et al., 1994), the inclusion oil has
never been affected by biodegradation and was thus 6.1. Biomarkers
trapped relatively quickly within the host minerals.
Aromatic compounds including alkylbenzenes, alkyl- Trace amounts of biomarkers were detected in the
naphthalenes, alkylphenanthrenes, alkylbiphenyls and Friendship-1 inclusion oil, but not in the system blank,
alkyldibenzothiophenes are present in the inclusion oil where they are either entirely absent or in much lower
and were detected in much higher abundances than in abundances relative to the internal standard (Fig. 5(D)
948 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

and (E)). Bicyclic sesquiterpanes including two C14 iso- Molecular distributions of the aromatic compounds
mers, drimane, three rearranged C15 isomers and ho- appear to be primarily controlled by thermal maturity.
modrimane were detected. The lower molecular weight In order to assess the level of thermal maturity of
C14 isomers are much more abundant than drimane (a the inclusion oil, aromatic ratios of the inclusion oil
C15 isomer) and homodrimane (a C16 isomer; Table 1). (Table 2) can be compared to published values from
C23 tricyclic and C24 tetracyclic terpanes are much less three different wells intersecting the middle Velkerri
abundant than pentacyclic terpanes (Table 1), and other Formation at different maturity stages (George and
tricyclic terpanes are below detection limit. Pentacyclic Ahmed, 2002). In the inclusion oil, there is a marked
terpanes present include Ts, Tm, C29 and C30 hopanes shift from high maturities based on lower molecular
and moretanes, C31 –C34 homohopanes, C30 diahopane weight parameters to lower maturities based on higher
and gammacerane (Fig. 5(D)). C30 ab hopane is the molecular weight parameters. For example, the meth-
most abundant hopane detected. The presence of ylnaphthalene ratio (2.1; Fig. 5(F)) is similar to the
gammacerane was confirmed by MRM. Furthermore, deeper sections of the middle Velkerri Formation in the
C13 –C16 acyclic isoprenoids are of very low abundance, McManus-1 well, which reached the late oil window
and norpristane, pristane and phytane (C18 –C20 acyclic (George and Ahmed, 2002). The dimethylnaphthalene
isoprenoids) are virtually absent. The m=z 205 mass ratio-1 (5.0) is similar to the shallower sections of the
chromatogram indicates the presence of C31 and C32 2a- middle Velkerri Formation in McManus-1, and trim-
methylhopanes, giving a signal about half that of the ethylnaphthalene ratios are similar to the deeper sections
corresponding hopanes in this chromatogram. Trace of the middle Velkerri Formation in the Walton-2 well
amounts of C27 –C29 steranes and diasteranes were de- (peak oil window). The methylphenanthrene index 1
tected by SIM and MRM analysis. (0.76; Fig. 5(E)) is similar to the middle Velkerri For-
mation in the Shea-1 well, which is in the peak oil
6.2. Thermal maturity window (George and Ahmed, 2002), and a calibration
(Radke and Welte, 1983) of the methylphenanthrene
Hopane/moretane ratios and homohopane 22S= index 1 with vitrinite reflectance (0.86% Rc ) supports this
ð22S þ 22RÞ ratios are at or near equilibrium values maturity level. Other maturity parameters such as those
(Table 1) and the Ts/Tm ratio (1.4) is consistent with based on alkylbiphenyls and alkyldibenzothiophenes
an oil window maturity for the inclusion oil. C29 ster- also mostly support a peak oil window maturity, except
anes are more abundant than C27 and C28 steranes, for the dimethylphenanthrene ratio which is anoma-
steranes are more abundant the diasteranes, and C29 lously low (Table 2). These aromatic maturity data re-
sterane maturity parameters also support a mid oil flect two different maturity fluids in this inclusion
window maturity (Table 1). sample, a low molecular weight dominated fluid (C11 )

Table 1
Biomarker parameters for the Friendship-1 inclusion oil
Parameter Value
Drimane/homodrimane 3.0
C14 bicyclic sesquiterpanes/(drimane + homodrimane) 4.5
C23 tricyclic terpane/C30 ab hopane 0.21
C24 tetracyclic terpane/C30 ab hopane 0.12
C24 tetracyclic terpane/(C24 tetracyclic terpane + C23 tricyclic terpane) 0.36
Ts/Tm{ 1.4
C29 Ts/C29 ab hopane{ 0.30
C30 */C30 ab hopane 0.07
C30 hopane ab=ðab þ baÞ 0.90
C31 ab 22S=ð22S þ 22RÞ{ 0.57
Gammacerane/C30 ab hopane{ 0.08
C29 ab hopane/C30 ab hopane 0.60
C31 ab hopane/C30 ab hopane 0.69
C27 :C28 :C29 abb steranes 25:33:42
C29 ba diasteranes/(aaa þ abb steranes){ 0.31
C29 aaa 20S=ð20S þ 20R) steranes{ 0.43
C29 abb=ðabb þ aaa) steranes{ 0.51
For hopane abbreviations see Fig. 5. All ratios were calculated from SIM data (m=z 123, 191, 218), except for those marked { which
were calculated from MRM data.
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 949

Table 2
Aromatic compound parameters for the Friendship-1 inclusion oil
Parameter Value
Methylnaphthalene ratio (MNR: 2-MN/1-MN) 2.1
Ethylnaphthalene ratio (ENR: 2-EN/1-EN) 2.8
Dimethylnaphthalene ratio (DNR-1) 5.0
Trimethylnaphthalene ratio (TNR-1) 0.69
Trimethylnaphthalene ratio (TNR-2) 0.72
Tetramethylnaphthalene ratio-1 (TeMNr-1) 3.3
Trimethylnaphthalene ratio (TMNr) 0.78
Tetramethylnaphthalene ratio (TeMNr) 0.87
Pentamethylnaphthalene ratio (PMNr) 0.78
Methylphenanthrene index (MPI-1) 0.76
Calculated vitrinite reflectance (Rc ) 0.86
Methylphenanthrene distribution fraction (MPDF) 0.42
Methylphenanthrene ratio (MPR; 2-MP/1-MP) 0.97
Dimethylphenanthrene ratio (DPR) 0.10
Methylbiphenyl ratio (MBpR: [3-MBp/2-MBp]) 7.1
Dimethylbiphenyl ratio (DMBpR-x) 6.7
Dimethylbiphenyl ratio (DMBpR-y) 9.7
Phenanthrene/dibenzothiophene 13.1
Methyldibenzothiophene ratio (MDR: 4-MDBT/1-MDBT) 6.0
Dimethyldibenzothiophene ratio (DMDR) 1.16
DNR-1 ¼ [2,6- + 2,7-DMN]/1,5-DMN; TNR-1 ¼ 2,3,6-TMN/[1,4,6- + 1,3,5-TMN]; TNR-2 ¼ [2,3,6- + 1,3,7-TMN]/[1,4,6- + 1,3,5 +
1,3,6-TMN]; TeMNr-1 ¼ (2,3,6,7-TeMN/1,2,3,6-TeMN); TMNr ¼ 1,3,7-TMN/[1,3,7 + 1,2,5-TMN]; TeMNr ¼ 1,3,6,7-TeMN/[1,3,6,7 +
1,2,5,6-TeMN]; PMNr ¼ 1,2,4,6,7-PMN/[1,2,4,6,7 + 1,2,3,5,6-PMN]; MPI-1: 1.5  [3-MP + 2-MP]/[P + 9-MP + 1-MP]; Rc ¼ [0.6  MPI-
1 + 0.4] (Radke and Welte, 1983); MPDF ¼ [3-MP + 2-MP]/R MPs; DPR ¼ [3,5 + 2,6-DMP + 2,7-DMP]/[1,3 + 3,9 + 2,10 + 3,10-DMP +
1,6 + 2,9 + 2,5-DMP]; DMBpR-x ¼ 3,5-DMBp/2,5-DMBp; DMBpR-y ¼ 3,30 -DMBp/2,30 -DMBp; DMDR ¼ 4,6-DMDBT/[3,6 + 2,6-
DMDBT].

with a high maturity and a higher molecular weight value is higher than 95% of oils in the original calibra-
dominated fluid (C13þ ) with a lower maturity. tion set (Bement et al., 1995) and is indicative of gen-
Low molecular weight hydrocarbons (C1 –C9 ) are eration in the late oil window, hence consistent with
abundant in the Friendship-1 inclusion oil (Fig. 6). The the maturity parameters of low weight aromatic
C5 –C9 range of these is dominated by n-alkanes, other compounds.
isomers that could be identified include methylalkanes,
dimethylalkanes, alkylcyclopentanes, alkylcyclohexanes
and aromatic hydrocarbons. Ratios based on some of 7. Discussion
these compounds are given in Table 3. Benzene and
toluene are of very low abundance relative to n-alkanes 7.1. History of oil migration
(Fig. 6(D), Table 3) and only traces of furan and
methylfurans could be identified. These data suggest no The presence of an unusual suite of saturated ketone
significant contribution of water soluble compounds isomers in the co-extract of solid bitumen and calcite-
dissolved in aqueous fluid inclusions to the low molec- hosted oil-bearing fluid inclusions (George and Jardine,
ular weight hydrocarbons of the inclusion oil, as is 1994), but their absence in the inclusion oil studied here
sometimes found in samples dominated by aqueous in- suggests that the ketones were present either in the solid
clusions (Ruble et al., 1998). Based on parameters H and bitumen or in oil-bearing fluid inclusions within the vein
I (Table 3), the Friendship-1 inclusion oil has a very high minerals, but not in oil-bearing fluid inclusions trapped
maturity, ‘‘supermature’’ according to Thompson in the feldspar, quartz, andesine and augite of the dol-
(1983). This interpretation is based not only on the erite matrix. Although it is possible based on the geo-
dominance of n-C7 over other C7 isomers (n-C7 repre- chemical evidence that some of the ketone-containing oil
sents nearly 50% of all C7 hydrocarbons; Table 3), but was trapped in calcite-hosted oil-bearing fluid inclu-
also on the dominance of monomethylalkanes over sions, petrographic and geochemical data suggest that
dimethylcyclopentanes (parameter I). The K1 parameter this is unlikely. The differences between the earlier geo-
is near unity (0.96), and the 2,4-/2,3-dimethylpentane chemical data and those of this study indicate at least
ratio is 0.89, which corresponds to a calculated oil two phases of hydrocarbon migration during the ther-
generation temperature (TC , Table 3) of 138 °C. This mal history of the dolerite sill. A history with multiple
950 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

Table 3
Low molecular weight hydrocarbon parameters for the Friendship-1 inclusion oil
Parameter Value
i-C5 /n-C5 1.2
Benzene/n-C6 0.03
Toluene/n-C7 0.03
(n-C6 +n-C7 )/(cyclohexane + MCH) 8.1
Heptane value (H) 49.5
Isoheptane value (I) 5.4
n-C7 /MCH 4.6
Cyclohexane/methylcyclopentane 0.73
n-C7 /2-MH 4.5
n-C7 /methylcyclopentane 6.2
3-Methylpentane/benzene 6.3
MCH/toluene 8.1
3-Methylpentane/n-C6 0.21
Benzene/toluene 2.1
Methylcyclopentane/MCH 0.73
K1 (2-MH + 2,3-DMP)/(3-MH + 2,4-DMP) 0.96
2-MH/3-MH 0.92
2,4-DMP/2,3-DMP 0.89
TC (°C) 138
K2 (3,3-DMP + 2,3-DMP + 2,4-DMP + 2,2-DMP)/(2-MH + 3-MH + 1,1-DMCP + c-1,3-DMCP + t-1,3-DMCP) 0.26
N2 /P3 (1,1-DMCP + c-1,3-DMCP + t-1,3-DMCP)/(3,3-DMP + 2,3-DMP + 2,4-DMP + 2,2-DMP) 0.77
% n-C7 of total C7 49.6
% 3RP (3,3-DMP + 2,3-DMP + 2,4-DMP + 2,2-DMP + 2-MH + 3-MH) 61.1
% 5RP (1,1-DMCP + c-1,3-DMCP + t-1,3-DMCP + t-1,2-DMCP + ECP) 14.6
% 6RP (MCH + toluene) 24.2
FI oil ratios derived from FID-equivalent data. MCH, methylcyclohexane; MH,P methylhexane; DMP, dimethylpentane; DMCP,
dimethylcyclopentane; ECP, ethylcyclopentane. Heptane value (H) ¼ 100  n-C7 / cyclohexane through to methylcyclohexane.
Isoheptane value (I) ¼ (2-MH + 3-MH)/(c-1,3) + t-1,3) + t-1,2-DMCP). TC ¼ 140 + [15  (ln {2,4-DMP/2,3-DMP})] (Bement et al.,
1995; Mango, 1997).

episodes of oil migration is supported by the petro- netic temperatures. The highest TH values are found in
graphic observations and microthermometric data. primary aqueous fluid inclusions in growth bands of
The first and oldest phase of migration involved ke- the vein quartz and imply minimum temperatures of
tone-rich pyrolytically-derived oil and most likely oc- trapping of about 130 °C. These temperatures are
curred during or soon after flash maturation of the significantly higher than temperatures at maximum
overlying organic-rich Velkerri Formation ca. 1280 Ma burial of the sequence and indicate that quartz pre-
(McDougall et al., 1965) as the sequence was intruded cipitated as the dolerite sill was cooling. The absence
by the dolerite sill (George and Jardine, 1994). The oil of primary oil-bearing fluid inclusions in the vein
would have migrated along fractures and vesicles in the quartz suggests that the initial phase of oil migration
cooling sill, where it would have experienced further ceased between the formation of the ketone-rich solid
thermal degradation to form solid bitumen in fractures bitumen and crystallisation of vein quartz, and was
and veins of the dolerite (George and Jardine, 1994). relatively short-lived and short-ranged. Evidently fluid
High heating rates may have also increased expulsion conduits between the source rock and the dolerite sill
efficiency by producing higher pressures and subsequent were closed during the post-intrusive cooling of the
stress fracturing in the source rock (Y€
ukler and Wallace, host sequence and may have only been opened again
1990). Schutter (2003b) notes that oil may also migrate during the Neoproterozoic reactivation of the Roper
into igneous rocks which attain lower vapour pressures Superbasin.
as they cool and that its movement may be further en- Type 3 oil inclusions in albitised zones of matrix
hanced by igneous volatiles and hydrothermal fluids. feldspar were trapped during or after spilitic alteration
Textural relationships and TH measurements indi- of the dolerite and albitisation of the feldspar. Albiti-
cate that the majority of oil inclusions were trapped sation is a fairly common diagenetic and low-tempera-
during events post-dating the fracture bitumen after ture hydrothermal process in sedimentary and igneous
the dolerite had cooled substantially towards diage- rocks (Petersson and Eliasson, 1997; Ehrenberg and
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 951

Jakobsen, 2001), where it usually leads to enhanced Evidently, this phase of oil migration took place once
porosity and increased reservoir capacity (James, 1995). all the porosity was destroyed by vein-filling minerals
Dissolution porosity, however, is controlled by the de- and the dolerite was fractured during the Neoprote-
gree of albitisation (Boles, 1982) such that only the rozoic reactivation event which also caused extensive
partially albitised plagioclases have intracrystalline po- deformation in the underlying Bessie Creek Sandstone
rosity (as is the case at Friendship-1) compared to (Dutkiewicz et al., 2003a; Volk et al., 2003). The
completely albitised plagioclases which lack porosity temperatures of entrapment inferred for these fracture
altogether. The altered zones in the Friendship-1 dolerite hosted oil inclusions and the inclusions in feldspar are
are not associated with obvious microfractures and can generally lower than those of aqueous fluid inclusions
be explained by the style of replacement similar to that within the vein quartz, suggesting that entrapment
recorded by Hirt et al. (1993) along sub-microscopic en temperatures decreased over time perhaps as a result
echelon (0 0 1) and (1 1 0) cleavages. Empirical observa- of the sill cooling and structural reactivation of the
tions and modelling suggest that albitisation occurs at Roper Superbasin in the Neoproterozoic (e.g., Powell
temperatures between 100 and 150 °C (i.e., during dia- et al., 1987).
genesis) (Boles, 1982; Baccar et al., 1993) and even as
low as 60 °C under low pCO2 conditions (Baccar et al., 7.2. Source of the oil
1993). The oil inclusions display a mean TH of 104  13
°C, so trapping temperatures were likely within the The geochemical, petrographic and microthermo-
range typically associated with albitisation. As the var- metric data from fluid inclusions in Friendship-1, in
iability of fluorescence colours is within inclusions of combination with the earlier isotope data (George et al.,
one assemblage of inclusions of this type, it is attributed 1994), strongly support a biogenic origin for the oil and
to compositional changes due to partially poor sealing exclude any abiotic processes that are apparent sources
of feldspars (Munz, 2001) rather than compositional of hydrocarbons in many non-sedimentary rocks (e.g.,
changes due to different source inputs or hydrocarbon Potter and Konnerup-Madsen, 2003). The textural re-
maturities (George et al., 2001). lationships and fluid inclusion microthermometry indi-
Type 2 fluid inclusions trapped in transcrystalline cate that oil inclusion entrapment occurred at diagenetic
micro-fractures cutting the vein calcite, along with temperatures and significantly after cooling of the sill.
some oil inclusions within feldspar alteration zones, By far the strongest evidence for biogenic origins comes
record the lowest TH values and texturally represent from the presence of biomarkers (molecular fossils),
the most recent fluid flow event in the dolerite sill traces of which have been preserved in the fluid inclu-
involving hydrocarbon fluids. Although these fractures sions for over one billion years.
cut calcite crystal boundaries, they only rarely extend The high abundance of n-alkanes, alkylcyclohexanes,
from the vein calcite into dolerite matrix minerals, due methylalkylcyclohexanes and monomethylalkanes, and
to high contrasts in physical properties between these the low abundance of acyclic isoprenoids including
different mineral phases (Groshong, 1988). The TH pristane and phytane in the Friendship-1 fluid inclusion
values of oil and coeval aqueous fluid inclusions show oil are typical of Proterozoic oils and sediments in gen-
similar distributions with means of 79  19 and 110  eral (e.g., Fowler and Douglas, 1987; Summons et al.,
16 °C, respectively. The TH of the aqueous inclusions 1988), and are similar to the Borrowdale-2 fluid inclu-
is taken as a good estimate of the entrapment tem- sion oil from the underlying Bessie Creek Sandstone
peratures (Nedkvitne et al., 1993). Entrapment pres- (Dutkiewicz et al., 2003a; Volk et al., 2003), as well as
sures were most likely around 250 bars (Dutkiewicz hydrocarbon extracts from associated solid bitumens
et al., 2003a). Entrapment of the oil was in the liquid (George et al., 1994) and source rocks, including the
phase. The location, petrography and microthermo- overlying Velkerri Formation (e.g., Summons et al.,
metric behaviour of these inclusions is strikingly sim- 1988). The overall C13þ geochemistry of the oil is similar
ilar to those described for the underlying Bessie Creek to hydrocarbon extracts from the overlying Velkerri
Sandstone (Dutkiewicz et al., 2003a; Volk et al., 2003) Formation shale. Biomarkers in the Friendship-1 fluid
and they are thus interpreted to be recording the same inclusion oil are dominated by hopanes, derived mainly
migration event. In both cases the inclusions occur in from the membrane lipids of prokaryotes (Ourisson
extensive sub-parallel transgranular or transcrystalline et al., 1987). The high content of monomethylalkanes
microfractures, the oil inclusions display a similar size likely indicates a cyanobacterial lipid input (Summons
and a uniform blue fluorescence colour and coeval et al., 1988), but these are somewhat lower in abundance
aqueous inclusions trap at least two populations of relative to n-alkanes than in either Borrowdale-2 fluid
hypersaline magnesium and/or calcium-rich brines, inclusions (Dutkiewicz et al., 2003a; Volk et al., 2003)
which may have been accessed at similar times from or in bitumen and source rock extracts from the
different pockets of fluids during formation of an ex- Roper Superbasin (Summons et al., 1988). This likely
tensive fracture network (e.g., McNutt et al., 1990). reflects the high thermal maturity of the dolerite-hosted
952 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

inclusion oil. Another marker for cyanobacterial input is high maturity light oil was trapped in some inclusions,
the presence of 2a-methylhopanes, which have been re- and a different and probably earlier charge of oil gener-
lated to cyanobacterial oxygenic photosynthesis (Sum- ated in the peak oil window was trapped in other inclu-
mons et al., 1999). The small amount of gammacerane sions. Second, condensate or high maturity light oil
that is present is indicative of a source rock deposited picked up a tail of lower maturity oil during migration,
under highly reducing conditions within a stratified water and this mixture was trapped in inclusions. Third, an
column (Sinninghe Damste et al., 1995). Unlike the earlier oil charge was flushed by a later condensate or
Borrowdale-2 fluid inclusion oil from the underlying high maturity light oil with the oil-condensate mixture
Bessie Creek Sandstone, very small amounts of steranes trapped in inclusions (Aplin et al., 2000). Based on geo-
and diasteranes are present in the Friendship-1 fluid in- chemistry alone, it is not possible to distinguish these
clusion oil. These are dominated by C29 steranes, with three possibilities. Microthermometric experiments,
low relative amounts of C27 steranes and diasteranes however, on 220 fluid inclusions failed to identify phase
most likely derived from eukaryotes. Although high C29 / changes consistent with the presence of a discrete gas-
C27 sterane ratios are generally associated with a terres- condensate phase within individual fluid inclusions
trial contribution of organic matter to Upper Palaeozoic (Munz et al., 2002; Teinturier et al., 2002; Grimmer et al.,
and younger source rocks (Huang and Meinschein, 2003). Therefore, fluid inclusion oil in the dolerite at
1979), high ratios have also been found in early Palaeo- Friendship-1 is interpreted to contain a mixture of gas-
zoic and Precambrian rocks deposited prior to the evo- condensate and peak-oil window mature oil.
lution of land plants (Grantham, 1986; Wang and Philp, The composition of these phases, including the bio-
1997; H€ old et al., 1999; Arouri et al., 2000). In addition, marker content of the inclusion oils, indicates that the
C27 –C29 steranes have been found in a wide range of fluids have likely been expelled from Precambrian source
Proterozoic and Archaean rocks (Summons and Walter, rocks. There is no evidence for contamination of this
1990; Brocks et al., 2003, 1999), including the Barney fluid inclusion oil by Phanerozoic-derived hydrocar-
Creek, Velkerri and McMinn formations from the bons. The exceptionally organic rich marine Velkerri
McArthur Basin (Summons et al., 1988; Summons and Formation (Warren et al., 1998) reached oil window
Walter, 1990), which show variations in the relative maturities in more deeply buried sections of the Roper
abundance of C29 isomers. In Precambrian rocks, ster- Superbasin in the Middle Proterozoic (George and Ah-
anes have been interpreted to be derived from sterols of med, 2002) and also supplied oil to the underlying Bessie
eukaryote cell membranes (Summons and Walter, 1990) Creek Sandstone (Dutkiewicz et al., 2003a). Less mature
although the exact phylogenetic position of these eu- Velkerri Formation sediments contain steranes and di-
karyotes remains unknown (e.g., Brocks et al., 2003), as asteranes with a similar distribution as the inclusion oil
does the cause of relative abundances of steranes in (Summons et al., 1988), although these could not be
Proterozoic and Archaean rocks. Significantly, the Vel- detected in more mature Velkerri Formation sediments.
kerri Formation is known to contain abundant acrit- Alternative source rocks, including the McMinn and
archs, including the distinctive protist Tappania (Javaux Barney Creek Formation sediments, contain more C27
et al., 2001). Overall composition and maturity param- steranes than C29 steranes, and have a high relative
eters derived from the molecular composition of the abundance of diasteranes (Summons et al., 1988). These
Friendship-1 inclusion oil are indicative of analysis of a specific biomarker comparisons, together with organic
mixture. Low molecular weight (C11 ) parameters such richness and proximity to the dolerite sill of the Velkerri
as the gasoline range hydrocarbons (H, I, TC ) and the Formation, suggest that this is the most likely source of
methylnaphthalene ratio indicate a high maturity, the Friendship-1 fluid inclusion oil. However, it is diffi-
probably in the late oil generation window (about 1.3% cult to assess if and to what extent, alternative and more
vitrinite reflectance equivalent [VRE]). In contrast, mature organic-rich formations in the Roper Superbasin
higher molecular weight parameters (C13þ ) including with a similar source character to the Velkerri Forma-
most biomarker, trimethylnaphthalene, alkylphenanth- tion, such as the Palaeoproterozoic Barney Creek
rene, alkylbiphenyl and alkyldibenzothiophene ratios Formation (Summons et al., 1988) contributed to the
indicate a lower maturity, in the peak oil generation gas-condensate range proportion of the fluid inclusion
window (0.8% VRE). The very strong dominance of oils. One line of evidence is the C7 hydrocarbon ring
C5 –C13 n-alkanes in the Friendship-1 inclusion oil is preference, dominated by isoalkanes (3-RP), which in-
typical of a condensate composition, and this is also in- dicates a lacustrine depositional environment, based on
dicated by the much higher abundance of alkylbenzenes the interpretative scheme of ten Haven (1996). As the
(C8 –C10 ) compared to other aromatic hydrocarbons. Barney Creek Formation was deposited in a lacustrine
Higher molecular weight hydrocarbons are certainly or lagoonal setting 1690 Ma and has total organic car-
present in the inclusion oil, but in significantly lesser bon (TOC) contents of up to 10% (Summons et al.,
abundance. There are three plausible explanations for 1988), this could well be the source of the gas-conden-
this interpreted mixture. First, separate condensate or sate. Based on Rock-Eval pyrolysis, the maturity of the
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 953

Barney Creek Formation ranges from marginally ma- occur in commercially viable accumulations (Koning,
ture to overmature (Summons et al., 1988), and thus 2003; Schutter, 2003a). Reservoir rocks may be basalts,
may have expelled gas or gas-condensates. Interestingly, andesites, rhyolites, syenites, granites, gneisses and mica-
there is no evidence of lacustrine sourcing in the higher schists and even kimberlites (Schutter, 2003a), with pe-
molecular weight portion of the Friendship-1 inclusion troleum products ranging from oil (light and heavy) and
oil, which is consistent with only expulsion of C13 gas to bitumen. However, as Schutter (2003b) points out,
hydrocarbons. igneous rocks and source rocks are rarely considered
An alternative explanation is that the gas-conden- together in a petroleum system context. As these oc-
sate, together with the ketone-rich fraction and the high currences form one of the last frontiers for petroleum
maturity oil was derived from the Velkerri Formation. If exploration, it is crucial to understand the source of
this were the case, then the C7 hydrocarbon source data hydrocarbons, their migration paths and their emplace-
would have to be assumed to be misleading. The low- ment conditions. In order to address some of these issues
ermost section of the Velkerri Formation closest to the we recommend, a fluid inclusion approach especially in
dolerite intrusion would have experienced the greatest rocks where the fluids are no longer present in the pore
thermal effect resulting in formation of ketones, which spaces but residues of solid bitumen or graphite remain.
were preserved as solid bitumen within the dolerite sill. The molecular composition of fluid inclusions should
However, the effects of thermal stress decrease rapidly also be used, especially where carbon isotope data are
with distance from the intrusion, and the zone of alter- equivocal or where the evidence for abiogenic hydro-
ation on either side of an intrusion is usually twice the carbons is largely circumstantial. Although compound-
thickness of the intrusive body (Saxby and Stephenson, specific isotope studies may distinguish between biogenic
1987; Raymond and Murchison, 1991). Consequently, and abiogenic sources of hydrocarbons (Sherwood Lol-
the composition and maturity of hydrocarbons gener- lar et al., 2002), molecular geochemistry of petroliferous
ated by pyrolysis induced by igneous intrusions will vary fluids can provide additional information not only on
with distance from the heat source and thus mimic the detailed composition of the trapped hydrocarbons
metagenesis and catagenesis stages of burial-driven but also on their maturity, their origin and the compo-
petroleum generation. It is possible, therefore, that the sition of the biomass from which the petroleum was
gas-condensate was generated in parts of the Velkerri derived. Petrography and microthermometry of fluid
Formation that were not hot enough to form pyrolysis inclusions can provide information on the timing of
products but were hotter than sections that experienced entrapment of fluids, the temperature and pressure
a temperature regime driven by burial-induced heating conditions of trapping and information on any coeval
alone. Migration of both oil generated by burial heating, gases and liquids. A better understanding of petroleum
and from sill-influenced sections of the Velkerri For- occurrences in non-sedimentary rocks may lead to more
mation would have occurred during or following the accurate predictions where they occur, insights into the
Neoproterozoic reactivation of the Roper Superbasin, diversity of the primordial biosphere and the nature of
whereas true pyrolysis products were limited to an early life in hot environments.
oil charge transformed to solid bitumen by thermal py-
rolytic alteration of a Velkerri-derived oil. This ther-
mally solidified oil would be restricted to the hot dolerite 8. Summary and conclusions
sill, although pyrolysed oils have the potential to be
preserved in fluid inclusions (Volk et al., 2002). Absence 1. A Mesoproterozoic dolerite sill had a complex fluid
of gas-condensate in the Bessie Creek Sandstone fluid migration history, with trapping of several phases
inclusions may be due to migration being fairly short- of hydrocarbon migration, both during the actual in-
range and controlled by permeability differences be- trusive episode (ca. 1.3 Ma) and also during the sub-
tween the dolerite and the underlying sandstone and the sequent cooling of the sill.
total volume of migrating hydrocarbon. Future work 2. Solid bitumens represent the oldest hydrocarbon mi-
analysing hydrocarbon extract from different levels gration event and preserve an unusual suite of ketone
within the Velkerri Formation should help to discern the isomers generated during flash maturation of the
exact source of the hydrocarbons. overlying Velkerri Formation source rock (George
and Jardine, 1994).
7.3. Implications for the source of hydrocarbons in 3. Hydrocarbons within oil inclusions trapped in doler-
crystalline rocks ite matrix minerals and vein minerals represent a later
episode of hydrocarbon migration, probably at dia-
A list of hydrocarbon occurrences in and associated genetic conditions.
with igneous rocks recently compiled by Schutter 4. Oil inclusions occur in clusters in albitised zones of
(2003a) clearly illustrates that these are quite common, labradorite crystals and in transcrystalline fractures
yet overlooked settings for petroleum which occasionally cutting vein calcite and rarely vein quartz.
954 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

5. The oil in the fluid inclusions was trapped in the liquid Aplin, A.C., Larter, S.R., Bigge, M.A., Macleod, G., Swar-
phase at around 110 °C and 250 bars, most likely dur- brick, R.E., Grunberger, D., 2000. Confocal microscopy of
ing the Neoproterozoic reactivation and uplift stage of fluid inclusions reveals fluid-pressure histories of sediments
the Roper Superbasin and involved a mixture of gas- and an unexpected origin of gas condensate. Geology 28,
1047–1050.
condensate and a mature oil and was, at least partly,
Arouri, K., Conaghan, P.J., Walter, M.R., Bischoff, G.C.O.,
accompanied by hypersaline Ca/Mg brines. Grey, K., 2000. Reconnaissance sedimentology and hydro-
6. GC–MS analysis shows that the inclusion oil is non- carbon biomarkers of Ediacarian microbial mats and
biodegraded and has a composition similar to other acritarchs, lower Ungoolya Group, Officer Basin. Precam-
Proterozoic oils and sedimentary organic matter. brian Research 100, 235–280.
There is geochemical evidence for the presence of a Baccar, M.B., Fritz, B., Made, B., 1993. Diagenetic albitization
very mature, light condensate fluid and a peak oil- of K-feldspar and plagioclase in sandstone reservoirs:
window maturity oil. Petrographic evidence does thermodynamic and kinetic modeling. Journal of Sedimen-
not suggest a trapping of these fluids in discrete fluid tary Petrology 63, 1100–1109.
inclusion assemblages and mixing of the hydrocarbon Bement, W.O., Levey, R.A., Mango, F.D., 1995. The temper-
ature of oil generation as defined with C7 chemistry
phases predated trapping.
maturity parameter (2,4-DMP/2,3-DMP ratio). In: Grimalt,
7. Biomarkers (mainly hopanes, monomethylalkanes, J.O., Dorronsoro, C. (Eds.), Organic Geochemistry: Devel-
alkylcyclohexanes and traces of steranes) point to a opments and Applications to Energy, Climate, Environment
source rock dominated by prokaryotic cyanobacterial and Human History. A.I.G.O.A. Donostia-San Sebastian,
organic matter with a small contribution from eu- pp. 505–507.
karyotes. This rules out an abiotic origin for the in- Boles, J.R., 1982. Active albitization of plagioclase, Gulf Coast
clusion oil. Tertiary. American Journal of Science 282, 165–180.
8. The most likely source rocks include the directly Brocks, J.J., Buick, R., Summons, R.E., Logan, G.A., 2003. A
overlying Velkerri Formation and the deeper ca. 1.6 reconstruction of Archean diversity based on molecular
Ga Barney Creek Formation from the underlying fossils from the 2.78 to 2.45 billion-year-old Mount Bruce
Supergroup, Hamersley Basin, Western Australia. Geochi-
McArthur Group, which may have contributed the
mica et Cosmochimica Acta 67, 4321–4335.
higher maturity gas-condensate. Brocks, J.J., Logan, G.A., Buick, R., Summons, R.E., 1999.
9. Oil inclusions provide pristine data on the composi- Archean molecular fossils and the early rise of eukaryotes.
tion of the Earth’s early biosphere and can lead to Science 285, 1033–1036.
a better understanding of poorly-understood petro- Dalziel, W.D., 1992. On the organization of American plates in
leum systems associated with crystalline rocks. the Neoproterozoic and the breakout of Laurentia. Geology
Today 2, 237–241.
Dutkiewicz, A., Rasmussen, B., Buick, R., 1998. Oil preserved
in fluid inclusions in Archaean sandstones. Nature 395, 885–
Acknowledgements 888.
Dutkiewicz, A., Volk, H., Ridley, J., George, S., 2003a.
We thank the Northern Territory Geological Survey Biomarkers, brines, and oil in the Mesoproterozoic, Roper
Superbasin, Australia. Geology 31, 981–984.
for access to samples and reports; Robinson Quezada
Dutkiewicz, A., Ridley, J., Buick, R., 2003b. Oil-bearing CO2 –
for help with gas chromatography-mass spectrometry CH4 –H2 O fluid inclusions: oil survival since the Palaeopro-
analysis; Tom Bradley for thin section preparation; the terozoic after high temperature entrapment. Chemical
Electron Microscopy Unit at Sydney University for as- Geology 194, 51–79.
sistance with SEM analysis. We are grateful to Robert Ehrenberg, S.N., Jakobsen, K.G., 2001. Plagioclase dissolution
C. Burruss and Colin Barker for their careful and con- related to biodegradation of oil in Brent Group sandstones
structive reviews of this paper. This research was sup- (Middle Jurassic) of Gullfaks Field, northern North Sea.
ported by an Australian Research Council fellowship Sedimentology 48, 703–721.
and grant to A. Dutkiewicz. Ettner, D.C., Lindblom, S., Karlsen, D., 1996. Identification
and implications of light hydrocarbon fluid inclusions from
the Proterozoic Bidjovagge gold-copper deposit, Finnmark,
Associate Editor – Rolando di Primio
Norway. Applied Geochemistry 11, 745–755.
Fassifihri, O., Robin, M., Rosenberg, E., 1992. Wettability of
reservoir rock at the pore scale-contribution of cryoscan-
ning electron-microscopy. Revue de l’Institut Francais du
References Petrole 47, 685–701.
Fowler, M.G., Douglas, A.G., 1987. Saturated hydrocarbon
Abrajano, T.A., Sturchio, N.C., Bohlke, J.K., Lyon, G.L., biomarkers in oils of late Precambrian age from Eastern
Poreda, R.J., Stevens, C.M., 1988. Methane–hydrogen gas Siberia. Organic Geochemistry 11, 201–213.
seeps, Zambales Ophiolite, Philippines: deep or shallow George, S.C., Ahmed, M., 2002. Use of aromatic compound
origin? Chemical Geology 71, 211–222. distributions to evaluate organic maturity of the Proterozoic
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 955

middle Velkerri Formation, McArthur Basin, Australia. In: Jackson, M.J., Powell, T.G., Summons, R.E., Sweet, I.P., 1986.
Keep, M., Moss, S.J. (Eds.), Proceedings of the Petroleum Hydrocarbon shows and petroleum source rocks in sedi-
Exploration Society of Australia Symposium, The Sedimen- ments as old as 1.7  109 years. Nature 322, 727–729.
tary Basins of Western Australia 3, pp. 253–270. Jackson, M.J., Raiswell, R., 1991. Sedimentology and carbon-
George, S.C., Jardine, D.R., 1994. Ketones in a Proterozoic sill. sulphur geochemistry of the Velkerri Formation, a mid-
Organic Geochemistry 21, 829–839. Proterozoic potential oil source in northern Australia.
George, S.C., Lisk, M., Eadington, P.J., Quezada, R.A., 1998. Precambrian Research 54, 81–108.
Geochemistry of a palaeo-oil column: Octavius 2, Vulcan Jackson, M.J., Southgate, P.N., Page, R.W., 2000. Gamma-ray
Sub-basin. In: Purcell, P.G., Purcell, R.R. (Eds.), Proceed- logs and U-Pb zircon geochronology – essential tools to
ings of Petroleum Exploration Society of Australia Sympo- constrain lithofacies interpretation of Palaeoproterozoic
sium, The Sedimentary Basins of Western Australia 2, pp. depositional systems. In: Grotzinger, J.P., James, N.P.
195–210. (Eds.), Carbonate Sedimentation and Diagenesis in the
George, S.C., Llorca, S.M., Hamilton, P.J., 1994. An integrated Evolving Precambrian World. SEMP Special Publication
analytical approach for determining the origin of solid 67, pp. 23–41.
bitumens in the McArthur Basin, northern Australia. Jackson, M.J., Sweet, I.P., Page, R.W., Bradshaw, B.E., 1999.
Organic Geochemistry 21, 235–248. The South Nicholson and Roper Groups: evidence for the
George, S.C., Ruble, T.E., Dutkiewicz, A., Eadington, P.J., early Mesoproterozoic Roper Superbasin. In: Bradshaw,
2001. Assessing the maturity of oil trapped in fluid B.E., Scott, D.L. (Eds.), Integrated Basin Analysis of the Isa
inclusions using molecular geochemistry data and visually- Superbasin using Seismic, Well-log and Geopotential Data:
determined fluorescence colours. Applied Geochemistry 16, an Evaluation of the Economic Potential of the Northern
451–473. Lawn Hill Platform. AGSO Record 1999/19, Canberra.
Gold, T., 2001. The Deep Hot Biosphere: the Myth of Fossils Jackson, M.J., Sweet, I.P., Powell, T.G., 1988. Studies on
Fuels. Copernicus Books, Heidelberg, Germany. 243 pp. petroleum geology and geochemistry, Middle Proterozoic,
Goldstein, R.H., Reynolds, T.J., 1994. Systematics of Fluid McArthur Basin Northern Australia I: petroleum potential.
Inclusions in Diagenetic Minerals. SEMP Short Course 31. Australian Petroleum Exploration Association Journal 28,
Society for Sedimentary Geology, Tulsa, Oklahoma, USA, 283–302.
pp. 199. James, R.A., 1995. Application of petrographic image analysis
Grantham, P.J., 1986. The occurrence of unusual C27 and C29 to the characterization of fluid-flow pathways in a highly-
sterane predominance in the two types of Oman crude oil. cemented reservoir: Kane Field, Pennsylvania, USA. Jour-
Organic Geochemistry 9, 1–10. nal of Petroleum Science Engineering 13, 141–154.
Grimmer, J.O.W., Pironon, J., Teinturier, S., Mutterer, J., Javaux, E.J., Knoll, A.H., Walter, M.R., 2001. Morphological
2003. Recognition and differentiation of gas condensates and ecological complexity in early eukaryotic ecosystems.
and other oil types using microthermometry of petroleum Nature 412, 66–69.
inclusions. Journal of Geochemical Exploration 78–79, 367– Jeffrey, A.W.A., Kaplan, I.R., 1988. Hydrocarbons and inor-
371. ganic gases in the Gravberg-1 well, Siljan Ring, Sweden.
Groshong, R.H., 1988. Low-temperature deformation mecha- Chemical Geology 71, 237–255.
nisms and their interpretation. Geological Society of Jensenius, J., Burruss, R.C., 1990. Hydrocarbon–water inter-
America Bulletin 100, 1329–1360. actions during brine migration: evidence from hydrocarbon
Haines, P.W., Rawlings, D.J., Sweet, I.P., Pietsch, B.A., Plumb, inclusions in calcite cements from Danish North Sea oil
K., Madigan, T.L., Krassay, A.A., 1999. Blue Mud Bay- fields. Geochimica et Cosmochimica Acta 54, 705–713.
1:250 000 Geological Map Series. Explanatory Notes, SD Kogarko, L.N., Kosztolanyi, C., Ryabchikov, I.D., 1987.
53-7. Australian Geological Survey Organisation and Geochemistry of the reduced fluid in alkali magmas.
Northern Territory Geological Survey, Darwin. Geochemistry International 24, 20–27.
Hirt, W.G., Wenk, H.-R., Boles, J.R., 1993. Albitization of Koning, T., 2003. Oil and gas production from basement
plagioclase crystals in the Stevens sandstone (Miocene), San reservoirs: examples from Indonesia, USA and Venezuela.
Joaquin Basin, California, and the Frio Formation (Oligo- In: Petford, N., McCaffrey, K.J.W. (Eds.), Hydrocarbons in
cene), Gulf Coast, Texas: a TEM/AEM study. Geological Crystalline Rocks. Geological Society Special Publication
Society of America Bulletin 105, 708–714. 214, pp. 83–107.
Hoffmann, C.F., Foster, C.B., Powell, T.G., Summons, R.E., Konnerup-Madsen, J., Larsen, E., Rose-Hansen, J., 1979.
1987. Hydrocarbon biomarkers from Ordovician sediments Hydrocarbon-rich fluid inclusions in minerals from the
and fossil cells of the alga Gloeocapsomorpha prisca Zalessky alkaline Ilimaussaq intrusion, South Greenland. Bulletin de
1917. Geochimica et Cosmochimica Acta 51, 2681–2697. Mineralogie 102, 642–653.
H€
old, I.M., Schouten, S., Jellema, J., Sinninghe Damste, J.S., Kralik, M., 1982. Rb-Sr age determinations on Precambrian
1999. Origin of free and bound mid-chain methyl alkanes in carbonate rocks of the Carpentarian McArthur Basin,
oils, bitumens and kerogens of the marine, Infracambrian Northern Territory, Australia. Precambrian Research 18,
Huqf Formation (Oman). Organic Geochemistry 30, 1411– 157–170.
1428. Ledlie, I.M., Torkington, J., 1988. Friendship-1, EP 5, McAr-
Huang, W.Y., Meinschein, W.G., 1979. Sterols as ecological thur Basin. Northern Territory Well Completion Report.
indicators. Geochimica Cosmochimica Acta 43, 739–745. CRAE Report No. 303624, Pacific Oil & Gas Pty. Ltd.
Hunt, J.M., 1996. Petroleum Geochemistry and Geology. W.H. Ledlie, I.M., Weste, G., Torkington, J., 1988. Borrowdale-1
Freeman, New York. 743 pp. and 2, EP 5, McArthur Basin, Northern Territory. Well
956 A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957

Completion Report. CRAE Report No. 303625, Pacific Oil Bauhinia Downs SE53-3. Northern Territory Geological
& Gas Pty. Ltd. Survey, Darwin.
Lewan, M.D., 1986. Stable carbon isotopes of amorphous Potter, J., Konnerup-Madsen, J., 2003. A review of the
kerogens from phanerozoic sedimentary rocks. Geochimica occurrence and origin of abiogenic hydrocarbons in igneous
et Cosmochimica Acta 50, 1583–1591. rocks. In: Petford, N., McCaffrey, K.J.W. (Eds.), Hydro-
Mango, F.D., 1997. The light hydrocarbons in petroleum – a carbons in Crystalline Rocks. Geological Society Special
critical review. Organic Geochemistry 26, 417–440. Publication 214, pp. 151–173.
Mauk, J.L., Burruss, R.C., 2002. Water washing of proterozoic Potter, J., Rankin, A.H., Treloar, P.J., Nivin, V.A., Ting, W.,
oil in the midcontinent rift system. American Association of Ni, P., 1998. The preliminary study of methane inclusions in
Petroleum Geologists Bulletin 86, 1113–1127. alkaline igneous rocks of the Kola igneous province, Russia:
McDougall, I., Dunn, P.R., Compston, W., Webb, A.W., implications for the origin of methane in igneous rocks.
Richards, J.R., Bofinger, V.M., 1965. Isotopic age determi- European Journal of Mineralogy 10, 1167–1180.
nations on precambrian rocks of the carpentaria region, Powell, T.G., Jackson, M.J., Sweet, I.P., Crick, I.H., Boreham,
northern territory, Australia. Geological Society of Austra- C.J., Summons, R.E., 1987. Petroleum Geology and Geo-
lia Journal 12, 67–90. chemistry, Middle Proterozoic McArthur Basin. Bureau of
McNutt, R.H., Frape, S.K., Fritz, P., Jones, M.G., MacDon- Mineral Resources. Report Number 1987/48.
ald, I.M., 1990. The 87 Sr/86 Sr values of Canadian shield Radke, M., Welte, D.H., 1983. The methylphenanthrene index
brines and fracture minerals with applications to ground- (MPI); a maturity parameter based on aromatic hydrocar-
water mixing, fracture history and geochronology. Geochi- bons. In: Bjorøy, M. et al. (Eds.), Advances in Organic
mica et Cosmochimica Acta 54, 205–215. Geochemistry 1981. Wiley, Chichester, pp. 504–512.
Mossman, D.J., Gauthier-Lafaye, F., Jackson, S., 2001. Car- Rawlings, D.J., 1999. Stratigraphic resolution of a multiphase
bonaceous substances associated with the paleoproterozoic intracratonic basin system: the McArthur Basin, northern
natural fission reactors of Oklo, Gabon: paragenesis, Australia. Australian Journal of Earth Sciences 46, 703–723.
thermal maturation and carbon isotopic and trace element Raymond, A.C., Murchison, D.G., 1991. The relationship
compositions. Precambrian Research 106, 135–148. between organic maturation, the widths of thermal aureoles
Munz, I.A., 2001. Petroleum inclusions in sedimentary basins: and the thickness of sills in the Midland Valley of Scotland
systematics, analytical methods and applications. Lithos 55, and Northern England. Journal of the Geological Society of
195–212. London 148, 215–218.
Munz, I.A., Yardley, B.W.D., Gleeson, S.A., 2002. Petroleum Ruble, T.E., George, S.C., Lisk, M., Quezada, R.A., 1998.
infiltration of high-grade basement, South Norway: pres- Organic compounds trapped in aqueous fluid inclusions.
sure–temperature–time–composition (P–T–t–X) constraints. Organic Geochemistry 29, 195–205.
Geofluids 2, 41–53. Salvi, S., Williams-Jones, A.E., 1997. Fischer–Tropsch synthe-
Nedkvitne, T., Karlsen, D.A., Bjørlykke, K., Larter, S., 1993. sis of hydrocarbons during sub-solidus alteration of the
Relationship between reservoir diagenetic evolution and Strange Lake peralkaline granite, Quebec/Labrador, Can-
petroleum emplacement in the Ula field, North Sea. Marine ada. Geochimica et Cosmochimica Acta 61, 83–99.
and Petroleum Geology 10, 255–270. Saxby, J.D., Stephenson, L.C., 1987. Effect of an igneous
Newell, K.D., Burruss, R.C., Palacas, J.G., 1993. Thermal intrusion on oil shale at Rundle (Australia). Chemical
maturation and organic richness of potential petroleum Geology 63, 1–16.
source rocks in Proterozoic Rice Formation, North Amer- Schutter, S.R., 2003a. Occurrences of hydrocarbons in and
ican Mid-continent rift system, Northeastern Kansas. around igneous rocks. In: Petford, N., McCaffrey, K.J.W.
American Association of Petroleum Geologists Bulletin (Eds.), Hydrocarbons in Crystalline Rocks. Geological
77, 1922–1941. Society Special Publication 214, pp. 35–68.
Oakes, C.S., Bodnar, R.J., Simonson, J.M., 1990. The system Schutter, S.R., 2003b. Hydrocarbon occurrence and explora-
NaCl-CaCl2 -H2 O: I. The ice liquidus at 1 atm total pressure. tion in and around igneous rocks. In: Petford, N., McCaf-
Geochimica et Cosmochimica Acta 54, 603–610. frey, K.J.W. (Eds.), Hydrocarbons in Crystalline Rocks.
Ourisson, G., Rohmer, M., Poralla, K., 1987. Prokaryotic Geological Society Special Publication 214, pp. 7–33.
hopanoids and other polyterpenoid sterol surrogates. An- Sherwood Lollar, B., Westgate, T.D., Ward, J.A., Slater, G.F.,
nual Review of Microbiology 41, 301–333. Lacrampe-Couloume, G., 2002. Abiogenic formation of
Page, R.W., Sweet, I.P., 1998. Geochronology of basin phases alkanes in the Earth’s crust as a minor source of global
in the western Mount Isa Inlier, and correlation with the hydrocarbon reservoirs. Nature 416, 522–524.
McArthur Basin. Australian Journal of Earth Sciences 45, Sinninghe Damste, J.S., Kenig, F., Koopmans, M.P., K€ oster,
219–232. J., Schouten, S., Hates, J.M., de Leeuw, J.W., 1995.
Parnell, J., 1988. Migration of biogenic hydrocarbons into Evidence for gammacerane as an indicator of water column
granites: a review of hydrocarbons in British plutons. stratification. Geochimica et Cosmochimica Acta 59, 1895–
Marine and Petroleum Geology 5, 385–396. 1900.
Petersson, J., Eliasson, T., 1997. Mineral evolution and element Summons, R.E., Jahnke, L.L., Hope, J.M., Logan, G.A., 1999.
mobility during episyenitization (dequartzification) and 2-Methylhopanoids as biomarkers for cyanobacterial oxy-
albitization in the postkinematic Bohus granite, southwest genic photosynthesis. Nature 400, 554–557.
Sweden. Lithos 42, 123–146. Summons, R.E., Powell, T.G., Boreham, C.J., 1988. Petroleum
Pietsch, B.A., Rawlings, D.J., Creaser, P.M., Kruse, P.D., geology and geochemistry of the Middle Proterozoic
Ahmad, M., Ferenczi, P.A., Findhammer, T.L.R., 1991. McArthur Basin, Northern Australia: III. Composition of
A. Dutkiewicz et al. / Organic Geochemistry 35 (2004) 937–957 957

extractable hydrocarbons. Geochimica et Cosmochimica Australia: evidence from oil inclusions and their geochem-
Acta 52, 1747–1763. istries. Journal of Geochemical Exploration 78–79, 437–441.
Summons, R.E., Walter, M.R., 1990. Molecular fossils and Volk, H., Horsfield, B., Mann, U., Suchy, V., 2002. Variability of
microfossils of prokaryotes and protists from proterozoic petroleum inclusions in vein, fossil and vug cements:
sediments. American Journal of Science 290-A, 212–244. geochemical study in the Barrandian Basin (Lower Palaeo-
Sweet, I.P., Brakel, A.T., Rawlings, D.J., Haines, P.W., Plumb, zoic, Czech Republic). Organic Geochemistry 33, 1319–1341.
K.A., Wygralak, A.S., 1999. Mount Marumba, Northern Volk, H., Mann, U., Burde, O., Horsfield, B., Suchy, V., 2000.
Territory-1:250 000 Geological Map Series. Explanatory Petroleum inclusions and residual oils: constraints for
Notes, SD 53-6. Australian Geological Survey Organisation deciphering petroleum migration. Journal of Geochemical
and Northern Territory Geological Survey, Darwin. Exploration 701, 307–311.
Szatmari, P., 1989. Petroleum formation by Fischer–Tropsch Wang, H.D., Philp, R.P., 1997. Geochemical study of potential
synthesis in plate tectonics. American Association of source rocks and crude oils in the Anardako Basin,
Petroleum Geologists Bulletin 73, 989–998. Oklahoma. American Association of Petroleum Geologists
Teinturier, S., Pironon, J., Walgenwitz, F., 2002. Fluid inclu- Bulletin 81, 249–275.
sions and PVTX modelling: examples from the garn Warren, J.K., George, S.C., Hamilton, P.J., Tingate, P., 1998.
formation in well 6507/2-2, Haltenbanken, Mid-Norway. Proterozoic source rocks: sedimentology and organic char-
Marine and Petroleum Geology 19, 755–765. acteristics of the Velkerri Formation, Northern Territory,
ten Haven, H.L., 1996. Applications and limitations of Australia. American Association of Petroleum Geologists
Mango’s light hydrocarbon parameters in petroleum corre- Bulletin 82, 442–463.
lation studies. Organic Geochemistry 24, 957–976. Wiggins, W.D., Harris, P.M., Burruss, R.C., 1993. Geochem-
Thompson, K.F.M., 1983. Classification and thermal history of istry of post-uplift calcite in the Permian Basin of Texas and
petroleum based on light hydrocarbons. Geochimica et New Mexico. Geological Society of America Bulletin 105,
Cosmochimica Acta 47, 303–316. 779–790.
Tseng, H.-Y., Burruss, R.C., Onstott, T.C., Omar, G., 1999. Yeh, H.W., Epstein, S., 1981. Hydrogen and carbon isotopes of
Palaeofluid-flow circulation within a Triassic rift basin: petroleum and related organic matter. Geochimica et
evidence from oil inclusions and thermal histories. Geolog- Cosmochimica Acta 45, 753–762.
ical Society of America Bulletin 111, 275–290. Y€
ukler, M., Wallace, G., 1990. The effect of heating rate on
Volk, H., Dutkiewicz, A., George, S., Ridley, J., 2003. Oil kerogen maturation and oil generation. American Associ-
migration in the Middle Proterozoic Roper Superbasin, ation of Petroleum Geologists Bulletin 74, 796.

You might also like