Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/260013748

Mechanisms for permeability modification in


the damage zone of a normal fault, northern
Perth Basin, Western Australia

ARTICLE in MARINE AND PETROLEUM GEOLOGY · FEBRUARY 2014


Impact Factor: 2.64 · DOI: 10.1016/j.marpetgeo.2013.10.012

CITATIONS READS

5 21

3 AUTHORS, INCLUDING:

Hugo K.H. Olierook Joseph Hamilton


Curtin University ana-min
9 PUBLICATIONS 21 CITATIONS 64 PUBLICATIONS 5,061 CITATIONS

SEE PROFILE SEE PROFILE

Available from: Hugo K.H. Olierook


Retrieved on: 10 December 2015
Marine and Petroleum Geology 50 (2014) 130e147

Contents lists available at ScienceDirect

Marine and Petroleum Geology


journal homepage: www.elsevier.com/locate/marpetgeo

Mechanisms for permeability modification in the damage zone


of a normal fault, northern Perth Basin, Western Australia
Hugo K.H. Olierook a, *, Nicholas E. Timms a, P. Joseph Hamilton b
a
Department of Applied Geology, Curtin University, GPO Box U1987, Perth, WA 6845, Australia
b
Department of Applied Mathematics, Australian National University, Canberra, ACT 0200, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Faults and their associated damage zones in sedimentary basins can be sealing, impeding fluid flow and
Received 9 February 2013 creating permeability barriers, or open, creating fluid pathways. This impacts the reservoir potential of
Received in revised form rocks in fault damage zones. Stylolitization and fracturing severely impacted permeability through
20 July 2013
compartmentalization and cementation of Apium-1, an exploration hole drilled in the northern Perth
Accepted 24 October 2013
Available online 30 October 2013
Basin, Western Australia. Apium-1 is located 1 km into the hanging wall block damage zone of a major
NNW-trending normal fault. The drill core consists of fine- to medium-grained quartz arenite overlain by
a coarse-grained lag and capped by impermeable shale. It was quantitatively characterized by sedi-
Keywords:
Fault zone
mentary and structural logging, and microstructural and porosity-permeability analysis. Fractures and
Fault damage stylolites in the damage zone of the major fault are shown to have been sealed. Extensional cracks have
Stylolites been sealed by quartz precipitation; shear fractures that locally preserve brecciation are always quartz
Sandstone and siderite cemented; stylolites are common and contain halos of quartz cementation. In each case,
porosity was reduced to approximately 1%, with concomitant reduction of permeability to <<0.01 mD.
These structures are observed to be interconnected in the core and are likely to form a larger-scale 3D
network of steeply-dipping fractures and shallowly-dipping stylolites. The bulk permeability of the
damage zone would reflect the permeability of the fractures and stylolites, compartmentalizing the
Mesozoic rocks in the northern Perth Basin into elongate NW-SE trending blocks if the magnitude of
stress does not exceed the cemented rock strength.
Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction as the fault accumulates displacement by many slip events (Childs


et al., 1996; Kim et al., 2000). Fault damage can occur as extensional
In sedimentary basins, faults and their associated ‘damage microcracks, conjugate shear fractures, polymodal fractures, brec-
zones’ can be sealing or transmissive to fluids, depending on the ciation, cataclasis, solution seams and stylolites (Kim et al., 2003,
deformation mechanisms and styles of damage involved (Fisher 2004). The assessment of strain accommodation in fault damage
and Knipe, 1998; Lloyd and Knipe, 1992; Odonne, 1990; Rutter zones by the characterization of meso- and microstructures, and
and Elliott, 1976; Shipton and Cowie, 2003). The established view diagenesis is useful for the assessment of structural modification of
of faults and fault zone architecture is that they comprise a ‘core fluid flow properties of petroleum reservoirs and aquifers (Bonson
zone’ (usually on the order of decimetres, even for very large et al., 2007; Johansen et al., 2005; Johnson, 2002; Kim et al., 2004;
displacement faults) associated with one or more principal Mory and Iasky, 1996; Tada and Siever, 1989). The relative timing of
displacement surface(s). The core zone is surrounded by broadly diagenetic fabrics and structural deformation is fundamental to the
tabular ‘damage zones’ e essentially where the host rocks contain understanding of permeability of fault zones (Billi et al., 2003;
structures that accommodate some strain but considerably less Faulkner et al., 2010; Fisher and Knipe, 1998, 2001; Kim and
than the principal displacement surface(s) (Billi et al., 2003; Kim Sanderson, 2010; Odling et al., 2004; Peacock et al., 1998; Volk
et al., 2000, 2004). Damage zones are considered to be relict ‘pro- et al., 2009). It is often difficult to predict the permeability of
cess zones’ e where damage occurs ahead of the fault tip, modified fault zones because they commonly form complex, three-
dimensional arrays of deformation structures that evolve
throughout a fault’s history (Fisher and Knipe, 2001; King et al.,
* Corresponding author. Tel.: þ61 432 632 167. 2008). Despite the significant implications of fault zones for hy-
E-mail address: h.olierook@student.curtin.edu.au (H.K.H. Olierook). drocarbon and geothermal exploration, groundwater management

0264-8172/$ e see front matter Ó 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.marpetgeo.2013.10.012
H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147 131

and CO2 storage in the Perth Basin, there are no direct studies of represents the first detailed structural analysis of a fault zone in the
their structure, deformation mechanisms, and effects on perme- Perth Basin and demonstrates the potential of an integrated
ability in this basin. microstructural approach to fault seal analysis, with applications
The aim of this study is to provide an integrated, quantitative including petroleum exploration, hot sedimentary aquifer (HSA)
analysis of sedimentology, petrography, diagenetic fabrics, orien- geothermal exploration, aquifer management, and CO2 capture and
tation and distribution of meso- and microstructures, and porosity storage.
and permeability of potential reservoir rocks in the damage zone of
a large fault in the northern Perth Basin. This will enable an 2. Geological setting
assessment of the controls on reservoir/aquifer quality and so
contribute to risk reduction for petroleum and geothermal explo- The Perth Basin is an extensive northesouth sedimentary basin
ration and groundwater extraction. This has been achieved by a in south-western Australia (Fig. 1A) that has had over 360 petro-
detailed analysis of core from one proximal well (Apium-1) and two leum exploration holes, 200 groundwater production wells, and
distal wells (Centella-1 and Mondarra-4) in the hanging wall more recently exploration for geothermal energy and carbon stor-
damage zone of a normal fault in the northern Perth Basin (Fig. 1). age (Champ, 2010; Crostella and Backhouse, 2000; Ennis-King and
This study integrates electron backscatter diffraction mapping Wu, 2005; Mory and Iasky, 1996; Timms et al., 2012). The majority
(EBSD), scanning electron microscope cathodoluminescence (SEM of hydrocarbon exploration and extraction is concentrated in the
CL) imaging, automated mineral analysis (AMA) and image analysis northern onshore Perth Basin. This section of the Perth Basin
techniques to quantify deformation microstructures associated contains an asymmetric graben known as the Dandaragan Trough,
with faulting, and their impacts on fluid pathways. This study an axis-parallel sub-basin bound to the east by the NeS trending

Figure 1. (A) Basin subdivisions in the northern Perth Basin. Apium-1, Mondarra-4 and Centella-1 are located in the Beharra Springs Terrace, east of an unnamed major fault;
modified from Crostella and Backhouse (2000). (B) Stratigraphy of the northern Perth Basin; Apium-1, Mondarra-4 and Centella-1 cores were taken from within the Early Triassic
Kockatea Shale to the Late Permian Dongara Sandstone (boxed area); modified from Mory and Iasky (1996).
132 H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147
H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147 133

Darling Fault, and compartmentalised to the north and south by of Centella-1 and Mondarra-4 cores were used to assess the lateral
major oblique fault zones (Fig. 1A) (Cockbain, 1990; Harris, 1994; extent of the damage zone in the hanging wall block of the major
Playford et al., 1976; Wilde and Nelson, 2001). The western regional fault.
margin of the Dandaragan Trough is the Cadda Terrace (Mory and Lithologies are described in terms of grain-size (Wentworth
Iasky, 1996). The Dandaragan Trough is the deepest part of the scale), sorting, rounding/sphericity and proportions of modal
Perth Basin filled by up to 15 km of Permian to Recent sedimentary minerals, using Folk’s sandstone classification scheme (Folk, 1965;
rock (Cockbain, 1990; Pryer et al., 2005). Wentworth, 1922). Ultraviolet light images of Apium-1 core were
Apium-1, Mondarra-4 and Centella-1 are located in the Dan- used to identify the presence of petroleum residue and diagenetic
daragan Trough, northern Perth Basin, and are 1.0 km, 4.8 km and cements (such as illite and siderite), (Lemon and Cubitt, 2009;
5.4 km to the east and in the hanging wall block of an unnamed Mozley, 1989). The orientation and depth of fractures and stylo-
NNW-trending fault, respectively (Fig. 1A). The unnamed fault is to lites were quantified using established methods (Blenkinsop and
the east of and sub-parallel with the Mountain Bridge Fault. It is Doyle, 2010). The thickness and amplitude of stylolites and any
oriented favourably for reactivation in the present-day stress field cross-cutting relationships were recorded. The average well devi-
(King et al., 2008). The fault is approximately 50 km long, with a ation of Apium-1 and Centella-1 is 5 and 16 from vertical,
maximum, predominantly dip-slip net displacement of 800 m respectively, with an average azimuth of 023 and 097, respectively.
along the fault plane approximately 10 km to the south of Apium-1, This has been accounted for in orientation data for structures.
identified from seismic imaging. The net displacement along the Laboratory measurements of porosity and permeability were
fault adjacent to Apium-1 is approximately 600 m. Apium-1, Cen- made from 47 core plug samples from Apium-1 by ARC Energy LtdÒ
tella-1 and Mondarra-4 are located on an uplifted fault block with using mercury-injection capillary pressure (Mitchell, 2004). Core
four-way dip closure, bound by the NNW-trending unnamed fault plugs were taken normal to the core axis (i.e., their axes are sub-
to the west, a NNW-trending fault to the east, an E-trending fault to horizontal, parallel to bedding), and have external diameters of
the south and a WNW-trending fault to the north (Fig. 1). 2.54 cm. This meant that core plugs were typically obliquely ori-
The depositional environment in the northern Perth Basin was ented at w30 to the strike of the steeply-dipping fractures, and
marine-dominated from the Permian to the Triassic, resulting in parallel to sub-horizontal stylolites. Twenty-four additional sam-
the deposition of the Dongara Sandstone and Kockatea Shale (Mory ples were taken from Apium-1 between 2758.50 and 2781.90 m,
and Iasky, 1996). These formations were deposited during the main impregnated with blue resin and used to make polished thin sec-
rift phase of basin development as Greater India and Australia tions to examine the petrography and microstructure. High reso-
separated during Gondwana breakup (Harris, 1994). All three wells lution (2 mm pixel dimension) plane- and cross-polarized optical
intersect the Late Permian Dongara Sandstone, which is overlain by photomicrographs of the whole thin sections were acquired using
the Early Triassic Kockatea Shale (Fig. 1B), identified from paly- the Zeiss Axio Imager 2 microscope at CSIRO, Perth, and were used
nology and breaks in induction-electrical and gamma-ray sonic to characterize the mineralogy via point counting. Detrital quartz
logs (Mory and Iasky, 1996; Owad-Jones and Ellis, 2000). The versus quartz overgrowths were able to be differentiated due to the
Kockatea ShaleeDongara Sandstone section comprises three presence of detrital iron oxide grain coatings. Porosity was quan-
distinct facies, subdivided into 7 units: (1) The basal Kockatea Shale tified in 2D from thin sections via an image analysis technique
is part of an offshore marine facies (unit 1); (2) the transition be- using ImageJÒ software (Rasband, 2011). The ‘Threshold Colour’
tween the two formations is part of a transgressive shoreface de- plug-in (Landini, 2010) permitted the area fraction of blue resin,
posit (unit 2), and; (3) the upper Dongara Sandstone consists of a and thus porosity, to be determined. Two values for porosity were
foreshore facies (units 3e7). There is a low order sequence strati- discriminated: ‘blue resin’ and ‘green-blue resin’, which represent
graphic boundary present between facies 1 and 2, marking a clear and partially clay-occluded porosity, respectively. The total 2D
change from a highstand systems tract (Dongara Sandstone) to a porosity values show moderate correlation with core plug porosity
transgressive systems tract (Kockatea Shale) (Mitchell, 2004; Mory values (R2 ¼ 0.558), (Fig. 7C). However, 2D porosity values are
and Iasky, 1996). The stratigraphy in Centella-1 and Mondarra-4 consistently lower and changes in porosity are exaggerated in
were sufficiently similar to that of Apium-1 to allow correlation comparison to core plug values. Nevertheless, 2D porosity values
between wells (Fig. 2). may still be used for relative porosity variations between samples
The Dongara Sandstone has previously been targeted as a pe- where core plugs are absent.
troleum reservoir (Mory and Iasky, 1996). However, the majority of Scanning electron microscope cathodoluminescence (SEM CL)
producing reservoirs in the Dongara Sandstone are in the footwall of imaging was used to resolve quartz overgrowths and/or cements
major faults. The permeability of hanging wall reservoirs tends to be because it is sensitive to variations in trace elements and defects,
significantly less (Mory et al., 2005; Mory and Iasky, 1996). Perme- and unlike most other techniques, it provides strong contrast be-
ability measurements of 47 core plugs taken from the Dongara tween detrital and diagenetic phases and can resolve syntaxial
Sandstone in Apium-1 found permeability variations by up to two veins and overgrowths in quartz (Augustsson and Bahlburg, 2003;
orders of magnitude that are not explained by relatively homoge- Kwon and Boggs, 2002; Lloyd and Knipe, 1992). Cath-
nous sandstone (Mitchell, 2004). The analysis of the contributing odoluminescence images were collected at Curtin University using
factors to this permeability variation are analysed in this study. a Philips XL30 scanning electron microscope fitted with a KE De-
velopments cathodoluminescence detector with working distances
3. Approach and methodology between 14.5 and 16.9 mm, a spot size of 6 and accelerating volt-
ages of 10.0 kVe12.0 kV.
The sedimentology and mesoscopic structures were character- Crystallographic orientation mapping was done to characterize
ized by sedimentary and structural logging of drill core recovered microscopic deformation mechanisms in the vicinity of damage in
from Apium-1, Centella-1 and Mondarra-4 (Fig. 2). Structural logs key samples via electron backscatter diffraction (EBSD) at Curtin

Figure 2. Core log of the Dongara sandstone formation in Apium-1, Mondarra-4 and Centella-1, with the horizontal distance between wells shown. Structures are drawn to scale;
basal fractures attributed to core handling are not drawn. Mineralogy is shown as pie charts, assessed from mineralogy point counts of transmitted, cross polarized light thin
sections of samples at arrowed locations. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
134 H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147

University, using the Zeiss EVO SEM with Oxford Instruments’ EBSD
system (Channel 5.10). This is a quantitative technique that can
resolve: (a) mineral phases, such as quartz and siderite; (b) grain
geometry statistics, outlining the orientation of particular grains;
(c) crystallographic preferred orientation data, and; (d) intra-grain
deformation, including lattice strain and quartz dauphiné twins.
Dauphiné twins in quartz are a 60 rotation around the c-axis
(<0001>), which primarily occurs at grainegrain impingement
boundaries (Comer, 1972; Mørk and Moen, 2007; Wenk et al., 2011).
For EBSD mapping, a working distance of 15.0 mm, a spot size of 5.5,
a stage tilt of 70.0 and an accelerating voltage of 20.0 kV was
utilized. The EBSD data were collected and processed with the
Oxford Instruments’ Channel 5.10 software using established pro-
cedures (Prior et al., 1999; Reddy et al., 2007). Visual comparison of
the processed data with optical photomicrographs indicates that no
significant artefacts were introduced by the noise reduction
routine. Dauphiné twins were redrawn from EBSD maps due to a
systematic misindexing issue that affected some grains, which
resulted in false isolated data points with 60 / Figure 3. Graph showing the stylolite and fracture populations recorded in the Don-
<0001 > misorientation with the host grain due to close rotational gara Sandstone per metre, versus the distance into the hanging wall block of the major
crystal symmetry of quartz. NNW-trending regional fault.
High-resolution (2 mm step-size) quantification of major and
trace mineralogy in two stylolite regions was done to assess the Dongara Sandstone decrease with distance from the major regional
degree of enrichment of immobile detrital phases (e.g., zircon and fault, with the highest density in the most proximal well (Apium-
TiO2), and thus quartz dissolution associated with stylolitization. 1), when compared to the distal wells (Centella-1 and Mondarra-4)
This was done via automated mineral analysis (AMA) (Dilks and (Fig. 3). Several other wells proximal to Apium-1 could not be used
Graham, 1985; Goldie Divko et al., 2010; Messent and Farmer, because the same facies observed in Apium-1 were not cored, and
2008; Reid, 1989) using the FEI QEMSCANÒ at Ammtec Ltd.Ò, therefore equivalent structural measurements could not be made.
Perth, Western Australia. This instrument comprises a scanning The presence of deformation structures in the three wells dem-
electron microscope with four light-element, energy dispersive X- onstrates the existence of a broad fault damage zone. However, due
ray spectrometers (EDS). QEMSCANÒ utilizes a software suite to the paucity of wells in the region, it is uncertain if the number of
(iMeasure) controlling automated data acquisition. Modal miner- structures per metre reduces logarithmically with distance from
alogy was determined from the X-ray spectral data using iExplorer. the fault, or has another relationship. Unfortunately, the cm- to m-
The working distance was 23 mm, with an accelerating voltage of scale fractures measured in Apium-1 are sub-seismic resolution,
15 kV and a probe current of 3 nA. Samples were coated using a and therefore the damage intensity versus distance relationship
10 nm-thick evaporative carbon coat to disperse electron charge cannot be quantified using seismic data.
induced in the SEM. Note that the method, as used, does not allow At the mesoscopic scale, fractures occur as steeply-dipping
distinction between minerals of same/similar chemistry, such as shear fractures with a normal sense of displacement and open
anatase and rutile, or illite and muscovite. sub-horizontal extensional fractures (Fig. 4). Shear fractures are
limited to fine- to medium-grained quartz arenite and are
4. Results cemented by red-brown siderite (units 3 and 4). Shear fractures
have variable dips and generally strike NW, parallel to minor faults
4.1. Sedimentary logging and petrography in the region (Fig. 5A). Jogs are occasionally present between shear
fractures that exhibit NeS trending, horizontal cylindrical holes up
Seven distinct units in Apium-1 can be grouped into three facies to 2 cm in diameter, which indicate a reverse sense of displacement.
(Mitchell, 2004). (1) Unit 1 is part of an offshore marine facies, Shear fractures become locally orthogonal to bedding-parallel
consisting of shale with lensoidal intercalations of medium- to stylolites. Open extensional fractures show no evidence of prior
coarse-grained sandstone. This ranges from an offshore marine shelf cementation. These are typically sub-horizontal, and are far more
(wholly shale), to lower offshore (<5 cm lenses of sand) to rare abundant in shale (mean w22 open fractures per metre) than in
upper offshore (10 cm beds of sandstone). (2) Unit 2 is part of a sandstone (mean 1.2 open fractures per metre). These fractures, less
transgressive shoreface deposit, consisting of coarse- to very coarse- than 1 mm wide, are parallel to the present-day maximum prin-
grained, friable, chlorite-cemented sandstone. (3) Units 3e7 are part cipal stress, i.e., sub-horizontal (King et al., 2008), and can therefore
of a foreshore facies, consisting of massive, fine- to medium-grained not be drilling-induced as they would need to be perpendicular to
quartz arenite. Point counting of thin sections shows that the Don- the present-day maximum principal stress, i.e., vertical (Fig. 5B).
gara Sandstone in Apium-1 is dominated by single crystal quartz Open fractures are not consistently in the same locations in
(67e84%), with minor detrital feldspar, authigenic quartz and clay opposite half-cores, are not observed in image logs, nor do they
minerals, and rare lithic fragments, rutile, zircon, pyrite, siderite and show any evidence of disking. These observations imply that open
Fe oxides/oxyhydroxides/siderite (Fig. 2). The sandstone units show fractures formed during core extraction, handling and slabbing and
very little mineralogical variation. were consequently not plotted on Figure 5A.
Stylolites are most common in fine- to medium-grained ho-
4.2. Meso-scale structural analysis mogenous quartz arenite and have a range of morphologies in
Apium-1. Stylolites show a strong linear relationship between
Fractures and stylolites were observed in fine- to medium- thickness and amplitude, which implies an evolution to thicker,
grained sandstones in Apium-1. Measurements of fractures and higher amplitude stylolites over time. They are variably filled with
stylolites from the same facies and intervals in the top of the predominantly organic carbon and clay minerals, but minor
H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147 135

Figure 4. Core (A) and transmitted, plane-polarized light (B) images of a typical cemented fracture, and; core (C) and transmitted, plane-polarized light (D) images of a typical open
fracture, where mineralogy changes are not apparent across fractures. Arrow indicates way-up direction. Samples have been impregnated with blue resin to highlight porosity. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

amounts of framework grains are present. Thick, clay- and organic 4.3. Microstructural analysis of fractures in Apium-1
carbon-rich stylolites occur predominantly in units 5 and 6 that
contain minor clay horizons (0.1e5 mm). Some stylolites have a Microstructural analysis of meso-scale shear fractures from
ribbon form defined by one or two strands over one millimetre in Apium-1 samples shows that they are accommodated by micro-
thickness, and up to ten thinner strands (Fig. 6E,F). Stylolites less brecciation, which is localised in discrete 0.1 mme2 mm bands
than 0.5 mm thick have low amplitudes (Fig. 6). along fracture surfaces, and are always cemented by siderite. In
Stylolites in Apium-1 and Centella-4 are most commonly some cases, micro-breccia zones are soft-linked by relay zones of
bedding-parallel (sub-horizontal), but are oriented at higher angles stylolites splaying out of fracture tip zones. Cathodoluminescence
proximal to cemented shear fractures (Fig. 5C,D). Seventy-five imaging reveals that fragmented clasts containing intragranular
percent of stylolites are sub-horizontal in Apium-1 and are inde- extensional fractures filled with quartz cement (i.e., syntaxial
pendent of other deformation structures, and the remainder have quartz veins) are present within the breccia zone, indicating that
formed at the tip zone of fractures and dip up to 60 . Steeply- these fractures preceded brecciation (Fig. 9C). Broader damage
dipping stylolites tend to be coincident with open fractures. How- zones that envelop micro-brecciated cores of shear fractures also
ever, these fractures are attributed to core handling. Sub-horizontal comprise quartz veins, oriented between 25 and 40 to shear
stylolites vary systematically in dip with a mean dip of 04 and fractures (Fig. 9C). The density of quartz veins increases toward
consistent strike (sub-parallel to the main regional fault), (Fig. 5C). meso-scale fracture surfaces (Fig. 9G). Total extension accommo-
Stylolites of similar orientations are spatially clustered, but there is dated by quartz veins in the vicinity of the shear fractures shown in
no systematic distribution with depth within the logged intervals Figure 8C is 5.2% (measured in the vein-normal direction from CL
(Fig. 7B). Stylolites in Centella-1 are less systematically oriented, images). Automated EBSD mapping shows that rare crystal-plastic
with a mean dip of 16 (Fig. 5D). Mondarra-4 shows only a few strain in quartz is localized to grain boundaries and occasionally
bedding-parallel stylolites. as intra-grain deformation, shown by low-angle (<2 ) mis-
Ultraviolet (UV) images of Apium-1 core show that hydrocarbon orientations in quartz grains (Fig. 9D). Dauphiné twins are localized
residue is clustered in two sections (Fig. 8). Both are approximately at grain boundaries and triple junctions (Fig. 9D). Open, unsealed
1.5 m thick, at depths of 2770.8 me2772.1 m and 2772.5 me intragranular fractures are extremely rare in all samples except in
2774.1 m. Specific gravity of these residues is approximately 45 unit 2 (very coarse grained sublithic-arenite) or occasionally as
 API, which is that of very light oils. This correlates with the sur- post-cementation extensional fractures.
rounding gas fields in production wells adjacent to Apium-1. The
deeper interval shows a stylolite acting as a local barrier (Fig. 8), 4.4. Microstructural analysis of stylolites in Apium-1
indicating that this stylolite formed and reduced adjacent vertical
permeability prior to hydrocarbon migration. Immediately below All stylolites have strongly anastomosing, sutured quartz grains
this, a series of stylolites is overprinting hydrocarbon shows, indi- along their interfaces indicative of quartz dissolution, and authigenic
cating that these stylolites initiated after hydrocarbon migration or quartz is precipitated in rock volumes on both sides of stylolites. Ul-
that adjacent porosity was still sufficiently high during migration. traviolet fluorescence, optical microscopy and AMA show that
136 H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147

2/3

2/3

Figure 5. Stereographic projections of fracture and stylolite orientations in Apium-1 using GEOrientÓ. The trend of a major regional fault is denoted by a black line, striking NNW.
The smaller scale NW-striking regional fault is also marked on with a grey line. (A) Poles to fractures in Apium-1; (B) Present day principal stress direction, inferred from shmax, after
King et al. (2008); (C) Poles to stylolites in Apium-1, and; (D) Poles to stylolites in Centella-1.

stylolites consist of mainly organic carbon and illite (Fig. 10AeD). equivalent to 50e70 mm of quartz lost vertically in sample 13 and
Automated mineral analysis shows concentrations of accessory pha- 200e300 mm of quartz lost vertically in sample 6. In two clay-rich
ses in stylolites, including TiO2 (rutile/anatase/brookite), zircon and sandstone samples (18 and 23), zircon and titanium oxides were
pyrite (Fig. 10CeE). Titanium dioxide (TiO2) phases have a bimodal enriched between 30 and 57 times in stylolites as adjacent to host
grain size distribution; the larger, rounded phase (50e1000 mm) is rock through quartz dissolution.
probably detrital rutile and the smaller TiO2 phase (3e15 mm) is
thought to be diagenetic anatase or brookite (Fig. 10E). Pyrite identi- 4.5. Diagenetic minerals and fabrics
fied under optical and SEM microscopy is located in stylolites, typi-
cally subhedral and 50e100 mm in diameter. It is also located as an Detrital quartz grain boundaries were identified by thin hae-
anhedral pore-filling phase and fine remnant of incomplete alteration matitic coatings in optical photomicrographs, which reveal that
to siderite near siderite-cemented fracture zones. optically continuous, authigenic overgrowth cements are common.
Local mass balance associated with quartz dissolution and The detrital to authigenic quartz ratio varies from 4:1 in fine-grained
precipitation around stylolites can be investigated by considering sandstone, 16:1 in medium-grained sandstone, and 18:1 to 28:1 in
the dissolution required to explain enrichment of insoluble min- coarse-grained sandstone. Up to four stages of quartz overgrowth
erals within stylolites versus the volume of cement in the sur- zones were identified by CL, and quartz grains typically show one or
rounding halo, thereby facilitating assessment of open or closed two overgrowth stages (Fig. 11). Crack-fill quartz cement cuts across
system behaviour. If it is assumed that in homogenous sandstone quartz overgrowth cements (Fig. 9C). Crack-fill quartz truncates
(units 3 and 4) detrital clays and heavy minerals were originally anastomosing stylolite boundaries (Fig. 9C) and is truncated by
evenly distributed, then zircon and titanium oxides were enriched stylolites (Fig. 10G). Diagenetic clay comprises 1e5% pore-filling
between 6 and 35 times in stylolites as compared to adjacent host kaolinite ‘books’ that pre-date quartz overgrowth cements and mi-
rock. The mechanism for this enrichment is via quartz dissolution nor amounts of diagenetic illite (Fig. 11B). Other diagenetic minerals
and removal. This would require one to two orders of magnitude include pyrite, siderite and anatase/brookite. Large diagenetic pyrite
of quartz to have been dissolved, when compared to the area of nodules (up to 80 mm across) occur between 2765 m and 2770 m
the image, to produce this enrichment of detrital minerals. This is (units 3 and 4). These are occasionally cross-cut by stylolites. Optical
H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147 137

Figure 6. Core photographs (A, C, E) and respective sample photomicrographs in transmitted plane-polarized light (B, D, F) of various thickness/amplitude stylolites. (A) and (B) are
from sample location 3 (2760.85 m) in medium-grained sandstone; C and D are from sample location 23 (2779.25 m) in coarse-grained sandstone, and; E and F are from sample 17
(2774.10 m) in fine-grained sandstone. Red arrows show stylolites in core and respective photomicrographs. (For interpretation of the references to colour in this figure legend, the
reader is referred to the web version of this article.)

microscopy and EBSD mapping shows that the siderite cement in of poor sample preparation. These two samples were not included
micro-breccia zones along shear fractures is red-brown and very in the cross plot between core plug- and thin section-derived
fine-grained (Fig. 9E). Residual pore spaces adjacent to micro- porosity (Fig. 7C). The primary control on porosity away from
breccia zones are also filled with siderite. Here and in the limited fractures and stylolites domains is grain size, with higher values
pore space in the micro-breccia, small siderite grains fringed the from coarser grained units, with a small proportion of the porosity
pore walls, and grew larger in the centre of pores via competitive partially occluded by diagenetic kaolinite (and minor illite). There
grain growth, forming a drusy texture. Siderite wraps around early- are strong variations in porosity within sandstone units of similar
formed kaolinite and has formed after quartz overgrowth and grain size. This variation is due to the presence of fractures and
brecciation. Anatase/brookite is typically subhedral, confined to stylolites. Significant porosity reduction in fractures has resulted
stylolites and very fine-grained, ranging from 3 to 15 mm. from siderite, pyrite, and minor quartz cementation. Domains
immediately adjacent to fractures typically do not show significant
4.6. Characterisation of porosity and permeability reduction in porosity. However, stylolites exhibit lateral porosity
reduction by authigenic quartz cementation between 1 and 20 mm
Porosity measurements of sandstones from optical photomi- wide, or more, outside the thin section area (Fig. 7D). These ’halos’
crographs vary between 0 and 20% (Fig. 7C). These porosity mea- of reduced porosity are commonly asymmetrically developed.
surements correlate moderately with core plug porosity There is a strong correlation between the thickness of stylolites and
(R2 ¼ 0.558), but are typically understated. Limited grain plucking width and intensity of the porosity reduction halo (Fig. 10I). Thicker
prior to resin impregnation has generated anomalously high and higher amplitude stylolites have wider halos of cementation.
porosity values from image analysis in two samples (and a high Stylolites formed along pre-existing clay-rich laminae are thicker
clear to partially-occluded porosity ratio), and is entirely an artefact and have higher amplitudes, typically with >20 mm halos of
138 H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147

A B Stylolites C Porosity D Permeability


Depth (m) Grain size Maximum Stylolite Amplitude (mm) (mD)
vcs 0.1 1 10
25

Tin section 2D porosity (%)


100
Legend
Legend 20
Kockatea Shale

Permeability (mD)
10
Hydrocarbon 15
residue
10 <10 cm from fractures
1 <10 cm from stylolites
Fractures 5 >10cm from styl./frac.
0 0.1
0 5 10 15 20 25 0 5 10 15 20 25
Helium porosity 3D (%) Helium porosity 3D (%)

Undeformed
f Core plug fluid
Styloltized permeability
Fractured and
stylolitized
f Clay-occluded
Dongara Sandstone

Figure 7. Graphical representation of the effect of lithology and stylolites on porosity and permeability: (A) Simplified sedimentary Log; (B) Stylolite distribution (thickness of bars
represents average thickness of stylolites, to scale); (C) Thin section porosity, with cross-plot between thin section and helium core plug porosity, and; (D) Core plug permeability,
with cross-plot between helium core plug porosity and permeability. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of
this article.)

decreased porosity. Stylolites in quartz arenite have cemented halos proximal to the major fault (Apium-1) to distal wells (Centella-1
between 1 and 20 mm wide. and Mondarra-4) infers that the regional fault has a significant
The permeability within lithological units varies by two orders damage zone (Fig. 3).
of magnitude and does not correlate with variations in ‘primary’ Steeply-dipping to sub-vertical shear fractures have a NW strike,
sedimentological characteristics (Fig. 7D). The maximum perme- parallel to the regional fault and surrounding minor faults. This
ability observed in each lithological unit away from deformation implies that they were formed in the wall or linkage damage zones
structures was assumed to represent the residual ‘primary’ in the same stress field as the regional fault (Fisher and Knipe,
permeability of undeformed rock, and was used as a benchmark for 2001; Kim et al., 2004). Porosity in shear fractures is now
relative comparison (Fig. 7D). The correlation between porosity and occluded, and this severely reduces the permeability across them.
permeability (Fig. 7D) shows that porosity is significantly reduced This compartmentalizes the basin, prohibiting fluid flow pathways
in the vicinity of cemented stylolites, but unaffected near fractures. across shear fractures, but promoting pathways NW-SE. Northe
Core plugs taken at or in close proximity (5 cm) to stylolites typi- South trending holes along jogs are indicative of releasing bends,
cally show an order of magnitude reduction from the maximum promoting linear, lateral NeS trending additional fluid flow path-
permeability in undeformed samples. Core plugs at fractures show ways. This late-stage reverse dip-slip reactivation of existing shear
permeability increases because they tend to be broken due to core fractures is thought to occur in the present-day stress field (King
extraction and handling. Increased silica overgrowths, character- et al., 2008).
ized by vitreous rims on quartz grains in core samples, reduce Stylolites in Apium-1 are most commonly bedding-parallel
permeability by up to two orders of magnitude. (sub-horizontal) implying that they formed by pressure solution
independent of other deformation features along existing bedding
5. Discussion planes (Tada and Siever, 1989). Alternatively, some sub-horizontal
stylolites could be fault-related, having developed during normal
5.1. Mesoscopic mechanisms for damage accommodation faulting (where s1 was vertical). A smaller proportion of stylolites
are steeply dipping, only adjacent to fractures, and are therefore
Meso-structural deformation is characterized by shear fractures thought to have formed in the tip-zone of propagating shear frac-
and stylolites. The decrease of fractures and stylolites from wells tures (Kim et al., 2000, 2004). Both types are predominantly filled
H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147 139

creep microstructures are pre-depositional (Fig. 9D). Dauphiné


twins form predominantly at the edge of grains, with some twins
formed at vein-detrital quartz boundaries, indicating that veins
formed syntaxial to pre-formed twins and undeformed grains. In
laboratory experiments, Hartley and Wilshaw (1973) applied load
rates of 106 msec1 on the crosscut X- and Y-axis surfaces of quartz
(perpendicular to the c-axis cut), which formed dauphiné twins at
critical loads of w100e150 N under ‘wet’ conditions (air) and
w200e400 N in ‘dry’ conditions (dry nitrogen) between 25  C and
150  C. The temperature conditions correspond to the known range
of burial depths, for which the maximum temperature was about
150  C (Thomas and Barber, 2004). Due to DMT processes creating
‘wet’ dissolved quartz, critical loads are only likely to be 100e150 N.
This gives an indication of the stress magnitudes and strain rates
required to form dauphiné twins in sedimentary basins.
Extensional and shear microfractures are also common in quartz
grains in Apium-1. These can occur with a significant component of
DMT in low-temperature deformation systems, leading to
cementation-related brecciation (Knipe, 1991; Onasch et al., 2010).
However, cataclasis in the studied samples is likely to have been
derived primarily from brittle shearing, without a significant
component of cementation and migration of grains because: (a)
Deformation bands have discrete to mildly gradual boundaries to
their host rock, which is consistent with a shear-fracture model
(Onasch et al., 2010), and; (b) Extensional fractures (now micro-
veins) occur between 25 and 40 to shear fractures, perpendic-
ular to s3 (Onasch et al., 2010).
Figure 8. (A) & (B) Core photograph of stylolite bounding oil residue, under normal Grain boundary elongation by dissolution and stylolitization
light and ultraviolet light, respectively. (C) Specific gravity of hydrocarbon residue, forms primarily by pressure solution creep, which involves dis-
showing that hydrocarbons in Apium-1 are w45  API indicative of very light oils.
solving, transporting and re-precipitating dissolved quartz (Canole
et al., 1997; Gratier et al., 2005; Tada and Siever, 1989). Stylolites are
with organic carbon and clay minerals (up to 95%), but there are most common in the fine- to medium-grained sandstone units
also concentrations of heavy minerals. This mineralogy together because of short pathways for DMT and higher density of graine
with the enhanced diagenetic cementation adjacent to stylolites grain contacts relative to coarse-grained units.
resulted in a drastic vertical permeability reduction and associated The stylolites in Apium-1 are characterized by concentration of
halos of porosity reduction (Fig. 10I). Hydrocarbon-rich domains heavy minerals and clay minerals (Fig. 10CeE), as is commonly
are sharply bounded by stylolites, implying that the initial forma- observed elsewhere (Heald, 1955; Tada and Siever, 1989). It is also
tion of some stylolites preceded hydrocarbon migration (Fig. 8). observed in Apium-1 that thicker, higher amplitude stylolites have
However, stylolites are also found within hydrocarbon-rich do- wider halos of cementation, indicating local source and sink of
mains, implying that hydrocarbon migration was contemporaneous quartz cementation. If stylolites nucleated homogenously in sand-
with either continuous or multiple-stage stylolite formation. Sty- stone, then a linear relationship between thickness of residual
lolites change orientation to steeper dips adjacent to fractures, phases and 2D porosity reduction halos would exist, which was not
which imply that stylolites and fractures interacted, and are most observed (Fig. 10I). This implies that stylolites nucleated on heavy
probably contemporaneous. mineral-rich and clay-rich layers, having a higher initial heavy
mineral/clay proportion and therefore skewing the results. Clay at
5.2. Microscopic mechanisms for damage accommodation quartz grainegrain contacts are known to promote pressure solu-
tion, and hence the formation of stylolites (Canole et al., 1997; Tada
Microstructural deformation is localized primarily near stylo- and Siever, 1989). This also helps explain the discrepancy between
lites and faults. Strain is accommodated by shear and tensile frac- zircon and TiO2 phases, where the difference in enrichment
turing, quartz dauphiné twinning, crystal plasticity and diffusive quantities is caused by probable authigenic anatase production,
mass transfer (DMT) through pressure solution. Stylolitization and enriching TiO2, or the non-uniform initial distribution of zircon and
fracturing are believed to be caused by predominantly DMT and rutile.
microfracturing, respectively (Knipe, 1989; Labaume et al., 2004; The quartz lost in the stylolites calculated from residual
Lloyd and Knipe, 1992; O’Kane et al., 2007). enrichment does not balance with quartz overgrowth cements in
Shear and tensile fractures, brecciation and dauphiné twinning the surrounding halos in the four analysed samples. There are
are observed in fractured regions of Apium-1. Impingement-related higher enrichments of residual phases than required from volumes
deformation at quartz grain contacts can occur by four mecha- of neighbouring authigenic quartz overgrowths. The excess dis-
nisms: (a) crystal-plastic dislocation activity away from impinge- solved quartz may have escaped through: (a) diffusion further into
ment; (b) densification by forming polymorphs of a-quartz; (c) the host sandstone outside the scale of the thin sections, or (b)
dauphiné twinning, or; (d) microfracturing (Hartley and Wilshaw, migration along shear fractures prior to sealing by siderite
1973; Lloyd, 2000; Lloyd and Knipe, 1992). Densification and cementation. However, core plug porosity and permeability mea-
crystal-plastic dislocation is not possible in low-temperature surements taken adjacent to stylolites do not show a significant
sedimentary basins (Hartley and Wilshaw, 1973; Peacock et al., reduction further than 10 cm away from any stylolites. This in-
1998). Some quartz grains do preserve crystal-plastic microstruc- dicates that fracture-related migration is the most likely cause for
tures; however, textural relationships indicate that dislocation the loss of dissolved quartz from the local system. This implies that
140 H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147
H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147 141

shear fractures were conduits to fluid flow pathways during the phases, such as zircon and rutile, in stylolites can be used to
evolution of the basin. further test this hypothesis. Automated mineral analysis
showed that zircon and rutile concentrations in stylolites are
5.3. Relative timing of structure development 1e2 orders of magnitude (up to 35 times) higher than in the
surrounding quartz arenite. This implies that these heavy
The relative timing of diagenetic phases and structures and the mineral phases are residual from the removal of large vol-
source of diagenetic phases has been used to establish a diagenetic umes of quartz by pressure solution (Hamilton et al., 2011).
and structural history that involved six distinct events: (1) Pre- The presence of SieTieO bearing grains indicates that the
depositional deformation; (2) Early diagenesis; (3) Pressure solu- likely source of titanium to form anatase is rutile needles in
tion, stylolitization and kaolinite formation; (4) Permian-Jurassic quartz grains because a single mineral does not exist con-
faulting; (5) Siderite cementation, and; (6) Present-day reac- taining only these elements. This also implies anatase
tivation (Fig. 12). formed during stylolite formation.
Kaolinite formation occurred after the onset of quartz
(1) Pre-depositional deformation is characterized by low-angle dissolution and re-precipitation (Fig. 11B). Kaolinite partially
misorientation boundaries that indicate crystal-plastic fills large secondary pore spaces, most likely produced by in
deformation. Crystal-plastic microstructures are at the situ alteration of feldspar. The most likely origin for this
edges of quartz grains, cannot be traced across multiple book-form kaolinite is in situ from dissolution of feldspar for
grains and cannot have formed during compaction in low- the following reasons: (a) the pore shapes that host kaolinite
temperature sedimentary basins (Knipe, 1989; Lloyd and retain pseudo-habits of feldspars with similar size to sur-
Knipe, 1992), implying that these formed prior to trans- rounding framework grains, and; (b) ‘books’ of 5e30 plate-
portation and deposition of the grains. lets are randomly oriented (Górniak, 1997; Wilson and
(2) Early diagenesis is characterized by nodular pyrite formation Pittman, 1977). Kaolinite is also not found within breccia
and the onset of quartz dissolution and re-precipitation, zones, which implies that feldspar alteration to kaolinite was
although the timing relationship between these two phases complete before brecciation.
is uncertain. The open texture of included clasts within the (4) Normal faulting occurred during Permian to Jurassic rifting,
pyrite nodules indicated that pyrite formed early during probably creating subsidiary, core-scale transgranular frac-
diagenesis, prior to significant mechanical compaction (Xiao tures (Song and Cawood, 2000). Steeply-dipping stylolites
et al., 2010). Nodules are thought to instigate through mi- are found splaying out from fracture surfaces. These stylolites
crobial reduction of sulphates to sulphide, typically in the either: (a) post-date the Permian to Jurassic normal faulting
form of hydrogen sulphide, which subsequently reacts with event, using the deformed fracturing surfaces as lower en-
free iron to form pyrite (Schenau et al., 2002; Worden et al., ergy nucleation sites (Gratier et al., 2005; Song and Cawood,
2003). The amoeboid shape of some detrital quartz cores 2000); or (b) stylolites nucleated and propagated syn-
prior to quartz overgrowths (Fig. 11B), long and sutured grain faulting, forming in the tip-zone of soft-linked fractures
contacts, kaolinite cementation and microfracturing indicate (Canole et al., 1997; Labaume et al., 2004; Rispoli, 1981). The
that quartz dissolution initiated before any cementation and latter mechanism is more likely, considering steeply-dipping
deformation. stylolites, those expected to form at tip zones, vary 40e50 in
(3) An initial quartz overgrowth cement phase is consistently strike to sealed fractures. This strike variation is expected in
truncated by stylolites, shear fractures within breccia cores stylolites located in tip damage zones of normal and strike-
and extensional fractures. Long and sutured grain contacts slip faults (Peacock et al., 1998). Tip-zone stylolites with
(Fig. 11A) throughout the host sandstone indicate that pres- non-zero dips formed due to a local perturbation of the stress
sure solution was triggered during compaction. Individual field, where the maximum principal stress, s1, is not vertical
quartz grains locally display up to four separate syntaxial (i.e., the suggested strike slip fault stress regime). Locally,
overgrowths, indicating that quartz cementation was quartz cementation continued during stylolite development
spatially and temporally heterogeneous, most probably and continued pressure-solution in the matrix. Sub-
derived from heterogeneous pressure solution (cf. Knipe, horizontal stylolites also occasionally truncate microveins
1989; Lloyd and Knipe, 1992). (Fig. 10G), indicating that stylolitization initiated prior to
The initiation of stylolitization is commonly between faulting, yet still occurred syn- and post-faulting. It is
w1000 and 2500 m depth in quartzose sandstone in sedi- possible some of these stylolites initiated during normal
mentary basins, implying that stylolites are probably pre- faulting, during which it is assumed that s1 was vertical,
ceded by heterogeneous compaction-related pressure promoting horizontal stylolite development. Regional
solution (Baron and Parnell, 2007; Gratier et al., 2005; normal faulting also resulted in the onset of dauphiné
Labaume et al., 2004; Rispoli, 1981). This supports iso- twinning, low-angle grain boundary slip sub-parallel to
chemical models of quartz cementation in reservoir sand- faulting, extensional fractures and shear fractures. Dauphiné
stones, and it is unlikely that quartz cement originated from twins are clearly crosscut by quartz veins (representing low-
externally-derived fluids (Fisher et al., 2000). Stylolites are angle grain boundaries) and a subsequent phase of quartz
also occasionally truncated by quartz-filled microcracks cements. Extensional microfractures are spatially and
(microveins), indicating that stylolitization begun prior to geometrically related to shear fractures, but are not filled
microfracturing (Fig. 9C). The concentration of insoluble with siderite cement; both sets of fractures are also observed

Figure 9. Fractures exhibit different microstructures that are shown to reduce porosity. (A) Siderite-filled breccia zone with significant quartz comminution, in thin section, under
plane polarized light. (B) Zoom-in of siderite-filled breccia zone in thin section. (C) CL image of (B), showing syntaxial quartz veins, oblique to breccia zone. (D) EBSD image of (B),
delineating detrital quartz boundaries, syntaxial quartz veins, low-angle misorientation boundaries and dauphiné twins. (E) A different breccia zone, coloured for grain size, with
quartz (dark red) and fine grained siderite, coloured for grain size. (F) Rose diagram showing orientations of tensile fractures (both open and quartz-filled) in relation to the breccia
zone margin in a breccia zone transect. (G) Extensional fracture frequency distribution across two breccia zones, measured as fractures per 2.5 mm by 2.5 mm areas, perpendicular
to the shear fracture. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
142 H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147
H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147 143

to be filled with quartz cement. An explanation for these respectively. Siderite cementation of the micro-breccia reduced the
observations is that extensional fractures (and shear frac- remaining porosity in the shear fractures from 13.9% to 1.6%, with
tures) formed rapidly-annealing crack-seal systems, where concomitant reduction in permeability of several orders of
newly-formed extensional and shear fractures were sealed magnitude.
with locally-derived quartz cement soon after microcracking,
eventually inhibiting migration of silica-bearing fluids 5.5. Implications for ancient and present-day fluid flow
(Ramsay, 1980). Quartz cement hardened both sets of frac-
tures, so upon further critical loading, micro-breccias formed The porosity and permeability evolution associated with
along shear fracture planes, opening fractures for infiltration different deformation structures allows analysis of the geometry
by exotic fluids that precipitated siderite. of fluid flow pathways over time (Fig. 14). It is highly likely that
It is also probable that hydrocarbon migration occurred ancient permeability in the sandstone units prior to structural
post-faulting. Two key observations were made regarding deformation, compaction and quartz overgrowth cementation
the timing of stylolite migration: (1) Ultraviolet images show was approximately isotropic, and decreased uniformly by
one occurrence of a stylolite acting as a barrier (Fig. 8), which compaction-related deformation (Fig. 2). An initial porosity of
locally prevented migration of hydrocarbons; (2) Stylolites 30%, estimated from 2D images (Fig. 13), is equivalent to a 3D
are seen to grow into hydrocarbon shows. The observations porosity of w15% (Fig. 7C) and can be correlated to an isotropic
imply that the onset of stylolitization occurred prior to hy- permeability of w50 mD, using porosityepermeability relation-
drocarbon migration, but stylolitization was also active post- ships in undeformed sandstones from Apium-1 (Fig. 7D),
migration. Geohistory modelling has shown that source (Mitchell, 2004).
rocks reached oil-generation window in the earliest Creta- The onset of stylolite formation early during diagenesis began to
ceous (Mory and Iasky, 1996), implying that hydrocarbon locally reduce porosity by quartz re-precipitation into surrounding
migration post-dated Permian-Jurassic faulting. pore space adjacent to sub-horizontal stylolites (Fig. 14). This effect
(5) Micro-brecciated fracture core zones are filled with siderite plus the residual layers of insoluble phases (such as clay minerals
cement. There are two possible sources for the iron and and heavy minerals) leave an impermeable, sub-horizontal layer
carbonate that comprises the siderite. The iron could be that severely impeded vertical fluid flow. The dissolved quartz is re-
derived locally from nearby cm-scale biogenic pyrite nodules precipitated into vertically-adjacent pore space. Although the finite
or Fe-rich shale. The carbonate could originate from car- length of individual stylolite bands cannot be established in drill
bonate alteration or exotic CO2-rich fluids sourced from core, laterally continuous stylolites may severely reduce vertical
organic matter degradation outside the sandstone beds. The fluid pathways. This resulted in stylolite-induced compartmental-
latter is likely due to the lack of local carbonate minerals in ization, which affected subsequent vertical hydrocarbon migration
the local stratigraphy (Schenau et al., 2002; Xiao et al., 2010). (Fig. 8), with preferential lateral fluid flow pathways and baffled
(6) Present-day stresses (transitional reverse to strike-slip) vertical pathways. Using porosityepermeability relationships be-
reactivated steeply-dipping sealed fractures to create cm- tween thin section and core plugs, 3D porosity and permeability
scale jogs in releasing bends (Kim et al., 2004; King et al., values have been reduced from 15% and 50 mD to w2.5% and
2008). Quartz veins were not observed in these fractures, << 0.01 mD after quartz cementation (Fig. 7).
indicating that stylolitization and pressure solution had When critical stress loads were achieved to brecciate shear
ceased before the present-day stress state had been fractures, sub-vertical and lateral fluid pathways were generated
established. along fracture planes (Fig. 14). Micro-breccia cores remained open
to fluid flow during this process, permitting precipitation of exotic
cements (e.g., siderite, and even pyrite) along transmissive shear
5.4. Porosity and permeability evolution fractures after their formation. Extension fractures in the associated
damage zone remained sealed, shown by the lack of siderite
The evolution of porosity was quantified in 2D for an unde- cementation, implying that fluid pathways were only along discrete
formed and a damaged domain by using EBSD and CL images to micro-breccia cores. Sealing of shear fractures by siderite post-
back-strip the effects of deformation and diagenetic events (Fig. 13). dated quartz cementation because breccia clasts within the frac-
The area percent of porosity at each stage was quantified using tured zone are only cemented with siderite (Fig. 9). The initial
ImageJ software (Fig. 13). Two main diagenetic porosity- and fracture porosity and permeability is unknown, but the final
permeability-altering events were identified: (1) authigenic quartz porosity after siderite cementation is less than 1%. Given pressure
overgrowth and fracture-fill cementation, and; (2) siderite solution rates from experimental studies, reduction of fracture
cementation following micro-brecciation in shear fractures. porosity from approximately 10% to approximately 0% by pressure
Detrital feldspar volumes altered to microporous kaolinite were solution of quartz could take on the order of one month to seal
sufficiently small to have caused only minimal enhancement of 10 mm wide fractures, and 1000 years to seal 100 mm wide fractures
porosity and permeability. (Gratier and Gueydan, 2007).
Initial 2D porosity was 34.7% and 30.0% in the fractured and Where coeval stylolites and fractures intersect, ‘crack seal sys-
undeformed samples of medium-grained quartz arenite, respec- tems’ were established that completely compartmentalize the
tively (Fig. 13). Authigenic quartz production reduced porosity to reservoir on a metre-scale, effectively reducing the finite perme-
12.6% and 4.7% for the fractured and undeformed regions, ability to almost zero (Fig. 14). Using porosityepermeability

Figure 10. Stylolites are shown to reduce porosity on the micro-scale. Note that the mineralogy legend is applicable for (C), (D) and (E). (A) & (B) Core photograph of fluorescent,
illite-bearing stylolite, under normal light and ultraviolet light, respectively. (C) AMA of clay-poor stylolite shown to be filled with muscovite/illite, pyrite, rutile/anatase/brookite,
zircon and siderite. (D) AMA of clay-rich stylolite enriched heavily in muscovite/illite, kaolinite and accessory minerals. (E) AMA of (D), showing only zircon and TiO2 phases,
highlighting enrichment in stylolite as opposed to surrounding host rock. (F) CL image showing highly sutured quartz grains at quartzestylolite interfaces (red arrows). (G) Syntaxial
quartz veins (green arrows) overprinting early (red arrows) and later (yellow arrows) quartz overgrowth phases. (H) Maximum amplitude of stylolites is proportional to stylolite
width. (I) Stylolite thickness is inversely proportional to porosity. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this
article.)
144 H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147

Figure 11. (A) Quartz (Q) has multiple overgrowth stages, divided into two mains phases: an early phase (dark grey arrows) and later phase (light grey arrows). (B) Kaolinite (K)
partially-occludes pore spaces that followed both early (dark grey arrows) and later (light grey arrows) quartz (Q) phases.

Inheritance of crystal- PRE-DEPOSITIONAL


plastic deformation DEFORMATION

Onset of quartz dissolution EARLY


and cementation DIAGENESIS

Initial quartz Kaolinite formation PRESSURE SOLUTION,


overgrowth from dissolution STYLOLITIZATION AND
cements of feldspar KAOLINITE FORMATION

PERMIAN-
Subsequent JURASSIC
quartz overgrowth NORMAL
cements
FAULTING
?

Hydrocarbon
migration
SIDERITE
?
CEMENTATION

PRESENT-DAY
REACTIVATION

Figure 12. Flow chart showing interpreted burial history of Apium-1.

relationships for sandstones in Apium-1 (Mitchell, 2004), a the northern Perth Basin would be compartmentalized by stylolites
porosity of 1% would match to a permeability of << 0.01 mD. Given and fractures on a metre-scale into elongate NW-SE trending
the interconnected structures observed in the core, they are likely blocks, which are aligned 26e30 to the present-day shmax in a
to form a 3D network of steeply-dipping fractures and shallowly transitional strike to slip-reverse fault stress regime.
dipping stylolites. The bulk permeability of the damage zone would
reflect the permeability of the fractures and stylolites. Therefore, it 6. Conclusions
is likely that ‘fossil’ (i.e., inactive) faults have impermeable fault
cores and an interlocking network of impermeable structures in  Meso-structural deformation is characterized by shear fractures
their damage zones. A significant caveat for this interpretation is and stylolites. Micro-structural deformation is accommodated
that the rock/cements are not overcome by the magnitude of the by shear and tensile fracturing, quartz dauphiné twinning and
present-day stress field (King et al., 2008). diffusive mass transfer through pressure solution. Stylolitization
The orientation of this 3D network would compartmentalize and fracturing are believed to be caused by predominantly DMT
present-day fluid flow into elongate, NW-SE trending blocks, with (pressure solution creep) and microfracturing.
significantly inhibited trans-fracture and trans-stylolite fluid  Six stages of diagenetic and structure development are recog-
migration (Fig. 14). The dense fault network identified from 3D nized: (1) Pre-depositional inheritance of crystal-plastic defor-
seismic (limited to faults with displacements >10 m), infers that mation. (2) Early diagenesis characterized by the onset of quartz
most of the Mesozoic rocks of similar mineralogy and grain size in dissolution and re-precipitation, and biogenic pyrite nodule
H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147 145

Figure 13. Quantitative porosity of: (A) the host sandstone, and; (B) a brecciated fracture zone. Note that the breccia zone in (B) only formed between the quartz overgrowth and
siderite phases, and hence is excluded in (i) to (iii). (C) Porosity evolution for the undeformed zone (A) and the brecciated zone (B).

formation. (3) Stylolitization, kaolinite formation and further development is uncertain. (5) Mobilization of CO2-rich fluids
pressure solution. (4) PermianeJurassic normal faulting char- through Fe-rich shale or pyrite nodules to form siderite that
acterized by dauphiné twinning at quartz grainegrain contacts, sealed fractures. (6) Present-day transitional strike-slip to
followed by development of extensional fractures and shear reverse-slip reactivation of breccia zones resulting in the for-
fractures with brecciation, tip-zone stylolites and possible fault- mation of cm-scale open jogs.
induced sub-horizontal stylolites. Hydrocarbon migration post-  Two main diagenetic porosity-altering events were identified:
dated faulting, but its timing with respect to siderite authigenic quartz overgrowths and quartz veining, and; siderite
146 H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147

A Mesozoic evolution Compaction

NNW
SSE
Fault

Stylolitization

Fracturing

B Present-day
NNW
SSE
Fault

Fracture sealing

NNW

Figure 14. Conceptual model of hanging wall damage zone network surrounding a major NNW-SSE trending fault. (A) Mesozoic evolution of fluid pathway anisotropy, shown by
arrows, through compaction, stylolitization, fracturing and fracture sealing. (B) Present-day fluid pathway anisotropy is compartmentalized by NNW-trending, sub-vertical sealed
fractures and sub-horizontal stylolites.

cementation following micro-brecciation in shear fractures. Blenkinsop, T.G., Doyle, M.G., 2010. A method for measuring the orientations of
planar structures in cut core. J. Struct. Geol. 32, 741e745.
Porosity was reduced from approximately 30% to <1%, with
Bonson, C.G., Childs, C., Walsh, J.J., Schöpfer, M.P.J., Carboni, V., 2007. Geo-
concomitant reduction in permeability of several orders of metric and kinematic controls on the internal structure of a large normal
magnitude. fault in massive limestones: the Maghlaq Fault, Malta. J. Struct. Geol. 29,
 The orientation of a 3D network of sub-horizontal stylolites and 336e354.
Canole, P., Odonne, F., Polve, M., 1997. Heterogeneous strain associated with normal
steeply dipping fractures would compartmentalize present-day faulting: evidence of mass transfer by pressure solution associated with fault
fluid flow into elongate, NW-SE trending blocks, with signifi- displacement. Tectonophysics 283, 129e143.
cantly inhibited trans-fracture and trans-stylolite fluid migra- Champ, P., 2010. Building a Robust Groundwater Model for the Southern Perth
Basin, WA: 3D Visualization and Modelling of Geology Using PetrelTM.
tion if stress magnitudes do not exceed the strength of the rock Groundwater, Canberra.
framework and cement. Childs, C., Watterson, J., Walsh, J.J., 1996. A model for the structure and development
of fault zones. J. Geol. Soc. (London) 153, 337e340.
Cockbain, A.E., 1990. Perth Basin, Geology and Mineral Resources of Western
Australia. In: Western Australia Geological Survey Memoir, pp. 495e524.
Acknowledgements
Comer, J.J., 1972. Electron microscope study of dauphiné microtwins formed in
synthetic quartz. J. Cryst. Growth 15, 179e187.
The authors would like to thank Ammtec Ltd. for automated Crostella, A., Backhouse, J., 2000. Geology and Petroleum Exploration of the Central
mineral analysis work and A. Vartesi for enhancing the figure and Southern Perth Basin, Western Australia. GSWA Report 57. Geological
Survey of Western Australia.
aesthetics. This study was supported by the Western Australia Dilks, A., Graham, S.C., 1985. Quantitative mineralogical characterisation of
Geothermal Centre of Excellence. R. King, D. Sanderson and an sandstones by back-scattered electron image analysis. J. Sediment. Petrol. 55,
anonymous reviewer are thanked for their constructive comments 347e355.
Ennis-King, J., Wu, G., 2005. Simulation of geological Storage of carbon dioxide in
during the peer review process. the onshore Perth Basin. Cooperative Research Centre for Greenhouse Gas
Technologies (CO2CRC). Report No: RPT05e0087.
Faulkner, D.R., Jackson, C.A.L., Lunn, R.J., Schlische, R.W., Shipton, Z.K.,
References Wibberley, C.A.J., Withjack, M.O., 2010. A review of recent developments con-
cerning the structure, mechanics and fluid flow properties of fault zones.
J. Struct. Geol. 32, 1557e1575.
Augustsson, C., Bahlburg, H., 2003. Cathodoluminescence spectra of detrital quartz
Fisher, Q.J., Knipe, R.J., 1998. Fault sealing processes in siliciclastic sedi-
as provenance indicators for Paleozoic metasediments in southern Andean
ments. In: Geological Society, London, Special Publications, vol. 147,
Patagonia. J. South Am. Earth Sci. 16, 15e26.
pp. 117e134.
Baron, M., Parnell, J., 2007. Relationships between stylolites and cementation in
Fisher, Q.J., Knipe, R.J., 2001. The permeability of faults within siliciclastic petroleum
sandstone reservoirs: examples from the North Sea, U.K. and East Greenland.
reservoirs of the North Sea and Norwegian Continental Shelf. Mar. Pet. Geol. 18,
Sediment. Geol. 194, 17e35.
Billi, A., Salvini, F., Storti, F., 2003. The damage zone-fault core transition in car- 1063e1081.
Fisher, Q.J., Knipe, R.J., Worden, R.H., 2000. Microstructures of deformed and non-
bonate rocks: implications for fault growth, structure and permeability.
deformed sandstones from the North Sea: implications for the origins of
J. Struct. Geol. 25, 1779e1794.
H.K.H. Olierook et al. / Marine and Petroleum Geology 50 (2014) 130e147 147

quartz cement in sandstones. In: Worden, R.H., Morad, S. (Eds.), Quartz Mozley, P.S., 1989. Complex Compositional zonation in concretionary siderite: im-
Cementation in Sandstones, Special Publication of the International Association plications for geochemical studies. J. Sediment. Petrol. 59, 815e818.
of Sedimentology, pp. 129e146. O’Kane, A., Onasch, C.M., Farver, J.R., 2007. The role of fluids in low temperature,
Folk, R.L., 1965. Petrology of Sedimentary Rocks. Hemphill, Austin, Texas. fault-related deformation of quartz arenite. J. Struct. Geol. 29, 819e836.
Goldie Divko, L.M., O’Brien, G.W., Harrison, M.L., Hamilton, P.J., 2010. Evaluation of Odling, N.E., Harris, S.D., Knipe, R.J., 2004. Permeability scaling properties of fault
the regional top seal in the Gippsland Basin: Implications for geological carbon damage zones in siliclastic rocks. J. Struct. Geol. 26, 1727e1747.
storage and hydrocarbon prospectivity. APPEA J., 463e486. Odonne, F., 1990. The control of deformation intensity around a fault: natural and
Górniak, K., 1997. The role of diagenesis in the formation of kaolinite raw materials experimental examples. J. Struct. Geol. 12, 911e913, 915e921.
in the Santonian sediments of the North-Sudetic Trough (Lower Silesia, Poland). Onasch, C.M., Farver, J.R., Dunne, W.M., 2010. The role of dilation and cementation
Appl. Clay Sci. 12, 313e328. in the formation of cataclasite in low temperature deformation of well
Gratier, J.-P., Gueydan, F., 2007. Deformation in the presence of fluids and mineral cemented quartz-rich rocks. J. Struct. Geol. 32, 1912e1922.
reactions: effect of fracturing and fluid-rocks interaction on seismic cycle. In: Owad-Jones, D., Ellis, G., 2000. Western Australia Atlas of Petroleum Fields. In:
Handy, M.R., Hirth, G., Hovius, N. (Eds.), The Dynamics of Fault Zones. Dahlem Onshore Perth Basin, vol. 1. Department of Minerals and Energy, Petroleum
Workshop. MIT Press, Cambridge, Massachusetts, USA, pp. 319e356. Division, Perth.
Gratier, J.P., Muquet, L., Hassani, R., Renard, F., 2005. Experimental microstylolites in Peacock, D.C.P., Fisher, Q.J., Willemse, E.J.M., Aydin, A., 1998. The relationship be-
quartz and modeled application to natural stylolitic structures. J. Struct. Geol. tween faults and pressure solution seams in carbonate rocks and the implica-
27, 89e100. tions for fluid flow. In: Geological Society, London, Special Publications, vol. 147,
Hamilton, J., Reddy, S.M., Olierook, H., Timms, N.E., 2011. Stylolites; their origin and pp. 105e115.
impact on reservoir quality. In: AAPG Annual Conference and Exhibition: Playford, P.E., Cockbain, A.E., Low, G.H., 1976. Geology of the Perth Basin, Western
Making the Next Giant Leap in Geosciences. American Association of Petroleum Australia. Geological Survey of Western Australia, Perth.
Geologists, Houston, Texas, U.S.A. Prior, D.J., Boyle, A.P., Brenker, F., Cheadle, M.C., Austin, D., Lopez, G., Peruzzo, L.,
Harris, L.B., 1994. Structural and tectonic synthesis for the Perth basin, Western Potts, G.J., Reddy, S.M., Spiess, R., Timms, N.E., Trimby, P., Wheeler, J.,
Australia. J. Pet. Geol. 17, 129e156. Zetterström, L., 1999. The application of electron backscatter diffraction and
Hartley, N.E.W., Wilshaw, T.R., 1973. Deformation and fracture of synthetic a-quartz. orientation contrast imaging in the SEM to textural problems in rocks. Am.
J. Mater. Sci. 8, 265e278. Mineral. 84, 1741e1759.
Heald, M.T., 1955. Stylolites in sandstones. J. Geol. 63, 101e114. Pryer, L., Loutit, T., Gardner, P., Petrovich, S., Teasdale, J., Stuart-Smith, P., Romine, K.,
Johansen, T.E.S., Fossen, H., Kluge, R., 2005. The impact of syn-faulting porosity Shi, Z., de Vries, S., Cathro, D., Etheridge, M., Foss, C., Munroe, M., Hamm, A.,
reduction on damage zone architecture in porous sandstone: an outcrop Ebrahim, M., Vizy, J., Henley, P., Mills, A., Guy-Villon, M., 2005. OZ SEEBASEÔ
example from the Moab Fault, Utah. J. Struct. Geol. 27, 1469e1485. Structural GIS. FrOG Tech Pty Ltd, Canberra, p. 184.
Johnson, L., 2002. Microstructural and Diagenetic History of a Hydrocarbon Reser- Ramsay, J.G., 1980. The crack-seal mechanism of rock deformation. Nature 284,
voir. Applied Geology. Curtin University, Perth, p. 96. 135e141.
Kim, Y.-S., Andrews, J.R., Sanderson, D.J., 2000. Damage zones around strikeeslip Rasband, W.S., 2011. ImageJ. U. S. National Institute of Health, Bethesda, Maryland,
fault systems and strikeeslip fault evolution, Crackington Haven, southwest U.S.A.
England. Geosci. J. 4, 53e72. Reddy, S.M., Timms, N.E., Pantleon, W., Trimby, P., 2007. Quantitative character-
Kim, Y.-S., Peacock, D.C.P., Sanderson, D.J., 2003. Mesoscale strike-slip faults and ization of plastic deformation and geological implications. Contrib. Mineral.
damage zones at Marsalforn, Gozo Island, Malta. J. Struct. Geol. 25, 793e812. Petrol. 153, 625e645.
Kim, Y.-S., Peacock, D.C.P., Sanderson, D.J., 2004. Fault damage zones. J. Struct. Geol. Reid, A., 1989. Measuring the unmeasureable e recent advances in applied
26, 503e517. mineralogy. AusIMM Bull. Proc. 294, 20e24.
Kim, Y.-S., Sanderson, D.J., 2010. Inferred fluid flow through fault damage zones Rispoli, R., 1981. Stress fields about strikeeslip faults inferred from stylolites and
based on the observation of stalactites in carbonate caves. J. Struct. Geol. 32, tension gashes. Tectonophysics 75, 29e36.
1305e1316. Rutter, E.H., Elliott, D., 1976. The Kinetics of rock deformation by pressure solution
King, R.C., Hillis, R.R., Reynolds, S.D., 2008. In situ stresses and natural fractures in [and discussion]. Phil. Trans. R. Soc. 283, 203e219.
the Northern Perth Basin, Australia. Aust. J. Earth Sci. 55, 685e701. Schenau, S.J., Passier, H.F., Reichart, G.J., de Lange, G.J., 2002. Sedimentary pyrite
Knipe, R.J., 1989. Deformation mechanisms e recognition from natural tectonites. formation in the Arabian Sea. Mar. Geol. 185, 393e402.
J. Struct. Geol. 11, 127e146. Shipton, Z.K., Cowie, P.A., 2003. A conceptual model for the origin of fault damage
Knipe, R.J., 1991. Microstructural analysis and tectonic evolution in thrust systems: zone structures in high-porosity sandstone. J. Struct. Geol. 25, 333e344.
examples from the Assynt Region of the Moine Thrust Zone, NW Scotland. In: Song, T., Cawood, P.A., 2000. Structural styles in the Perth Basin associated with the
Barber, D.J., Meredith, P.G. (Eds.), Deformation Processes in Minerals, Rocks and Mesozoic break-up of Greater India and Australia. Tectonophysics 317, 55e72.
Ceramics. Unwin Hyman, London, pp. 228e261. Tada, R., Siever, R., 1989. Pressure solution during diagenesis. Annu. Rev. Earth
Kwon, Y.-I., Boggs, S., 2002. Provenance interpretation of tertiary sandstones from Planet. Sci. 17, 89e118.
the Cheju Basin (NE East China Sea): a comparison of conventional petrographic Thomas, B.M., Barber, C.J., 2004. A re-evaluation of the hydrocarbon habitat of the
and scanning cathodoluminescence techniques. Sediment. Geol. 152, 29e43. northern Perth Basin. J. Aust. Pet. Explor. Assoc. 44, 13e57.
Labaume, P., Carrio-Schaffhauser, E., Gamond, J.-F., Renard, F., 2004. Deformation Timms, N.E., Corbel, S., Olierook, H., Wilkes, P., Delle Piane, C., Sheldon, H., Alix, R.,
mechanisms and fluid-driven mass transfers in the recent fault zones of the Horowitz, F., Wilson, M., Evans, K.A., Grifiths, C., Stütenbecker, L., Israni, S.,
Corinth Rift (Greece). Compt. Rend. Geosci. 336, 375e383. Hamilton, P.J., Esteban, L., Cope, P., Evans, C., Pimienta, C., Dyt, C., Huang, X.,
Landini, G., 2010. Software: Threshold_colour, 1.12 ed. Hopkins, J., Champion, D., 2012. Project 2: Geomodel. WA Geothermal Centre of
Lemon, N.M., Cubitt, C.J., 2009. Illite Fluorescence Microscopy: a New Technique in Excellence. CSIRO Report, p. 188.
the Study of Illite in the Merrimelia Formation, Cooper Basin, Australia. In: Volk, H., Kempton, R.H., Ahmed, M., Gong, S., George, S.C., Boreham, C.J., 2009.
Worden, R.H., Morad, S. (Eds.), Clay Mineral Cements in Sandstones. Blackwell Tracking petroleum systems in the Perth Basin by integrating microscopic,
Publishing Ltd., Oxford, U.K., pp. 409e424. molecular and isotopic information on petroleum fluid inclusions. J. Geochem.
Lloyd, G.E., 2000. Grain boundary contact effects during faulting of quartzite: an Explor. 101, 109.
SEM/EBSD analysis. J. Struct. Geol. 22, 1675e1693. Wenk, H.-R., Janssen, C., Kenkmann, T., Dresen, G., 2011. Mechanical twinning in
Lloyd, G.E., Knipe, R.J., 1992. Deformation mechanisms accommodating faulting of quartz: shock experiments, impact, pseudotachylites and fault breccias. Tecto-
quartzite under upper crustal conditions. J. Struct. Geol. 14, 127e143. nophysics 510, 69e79.
Messent, B.E.J., Farmer, L.E., 2008. Basker field, Gippsland basin: assessing Wentworth, C.K., 1922. A scale of grade and class terms for clastic sediments. J. Geol.
reservoir connectivity with the aid of complementary technologies. In: 30, 377e392.
Blevin, J.E., Bradshaw, B.E., Uruski, C. (Eds.), Eastern Australasian Basins Wilde, S.A., Nelson, D.R., 2001. Geology of the Western Yilgarn Craton and Leeuwin
Symposium III, Petroleum Exploration Society of Australia, Special Publica- Complex, Western Australia d a Field Guide. Record 2001/15. Western Australia
tion, pp. 29e44. Geological Survey, p. 41.
Mitchell, J., 2004. Apium-1 Interpretive Well Completion Report, L1/L2 Western Wilson, M.D., Pittman, E.D., 1977. Authigenic clays in sandstones; recognition and
Australia. ARC Energy. influence on reservoir properties and paleoenvironmental analysis. J. Sediment.
Mørk, M.B.E., Moen, K., 2007. Compaction microstructures in quartz grains and Petrol. 47, 3e331.
quartz cement in deeply buried reservoir sandstones using combined petrog- Worden, R.H., Smalley, P.C., Barclay, S.A., 2003. H2S and diagenetic pyrite in North
raphy and EBSD analysis. J. Struct. Geol. 29, 1843e1854. Sea sandstones: due to TSR or organic sulphur compound cracking? J. Geochem.
Mory, A.J., Haig, D.W., McLoughlin, S., Hocking, R.M., 2005. Geology of the northern Explor. 78e79, 487e491.
Perth Basin, Western Australia e a Field Guide. Western Australia Geological Xiao, S., Schiffbauer, J.D., McFadden, K.A., Hunter, J., 2010. Petrographic and SIMS
Survey, Perth, p. 71. pyrite sulfur isotope analyses of Ediacaran chert nodules: Implications for mi-
Mory, A.J., Iasky, R.P., 1996. Stratigraphy and Structure of the onshore northern Perth crobial processes in pyrite rim formation, silicification, and exceptional fossil
Basin, Western Australia. Report 46. Western Australia Geological Survey. preservation. Earth Planet. Sci. Lett. 297, 481e495.

You might also like