You are on page 1of 24

Available online at www.sciencedirect.

com

Geochimica et Cosmochimica Acta 72 (2008) 2510–2533


www.elsevier.com/locate/gca

Late Palaeozoic hydrocarbon migration through the Clair


field, West of Shetland, UK Atlantic margin
Darren F. Mark a,b,*, Paul F. Green c, John Parnell a, Simon P. Kelley d,
Martin R. Lee e, Sarah C. Sherlock d
a
Department of Geology & Petroleum Geology, Meston Building, Meston Walk, University of Aberdeen, Aberdeen, AB24-3UE, UK
b
NERC Argon Isotope Facility, Scottish Universities Environmental Research Centre, Rankine Avenue, Scottish Enterprise Technology
Park, East Kilbride, G75-0QF, UK
c
Geotrack International Ltd, 37 Melville Road, Brunswick West, Vic. 3055, Australia
d
Centre for Earth, Planetary, Space and Astronomical Research (CEPSAR), Department of Earth Sciences, Open University,
Walton Hall, Milton Keynes, MK7-6AA, UK
e
Department of Geographical and Earth Sciences, Lilybank Gardens, University of Glasgow, Glasgow, G12-8QQ, UK

Received 9 July 2007; accepted in revised form 28 November 2007; available online 26 February 2008

Abstract

Geochemical analysis of bitumen- and hydrocarbon-bearing fluid inclusions from the Devonian-Carboniferous Clair field
indicates that the reservoirs contain a mixture of oils from different marine and lacustrine sources. Reconstruction of the Clair
field oil-charge history using fluid inclusion petrography show that oil-charging occurred at times of K-feldspar, quartz and
calcite cementation. Temperature–composition–time data yielded from the integration of fluid inclusion microthermometry
with high-resolution Ar–Ar dating, date hydrocarbon-bearing K-feldspar overgrowths at 247 ± 3.3 Ma. These data show that
in order for oil to be trapped within primary fluid inclusions in K-feldspar overgrowths, hydrocarbon migration throughout
the UK Atlantic margin must have been taking place during the Late Palaeozoic and as such, current industry oil-play models
based solely on oil charging from Jurassic-Cretaceous marine sources are clearly incomplete and need revision. Apatite fission
track analysis and vitrinite reflectance data were used to reconstruct thermal burial histories and assess potential oil genera-
tion from Middle Devonian lacustrine source rocks. Thermal history data from wells along The Rona Ridge adjacent to the
Clair field show that the Palaeozoic section was heated to greater than 100 °C at some time between 270 and 230 Ma, con-
firming that Devonian source rocks were mature and expelling oil during the Late Palaeozoic at the time that authigenic K-
feldspar overgrowths were growing in the Clair field.
Ó 2008 Elsevier Ltd. All rights reserved.

1. INTRODUCTION rately constrain the timing of hydrocarbon generation and


whilst studies (Scotchman et al., 1998; Holmes et al.,
The West of Shetland (WOS) region of the UK Atlantic 1999) have shown the presence and effectiveness of hydro-
margin has been of exceptional interest for its hydrocarbon carbon source rocks in the adjacent Faeroe-Shetland Basin
prospectivity. Petroleum system analysis has failed to accu- (FSB), determination of the amounts, type and timing of
hydrocarbon generation have been problematical. A critical
issue upon which exploration and appraisal depend is the
* timing of hydrocarbon migration throughout the margin,
Corresponding author. Address: NERC Argon Isotope Facil-
ity, Scottish Universities Environmental Research Centre, Rankine and although a series of models show postulated timings
Avenue, Scottish Enterprise Technology Park, East Kilbride, of oil generation and expulsion (Iliffe et al., 1999; Lamers
G75-0QF, UK. Fax: +44 1355229898. and Carmichael, 1999; Carr and Scotchman, 2003; Scotch-
E-mail address: d.mark@suerc.gla.ac.uk (D.F. Mark). man et al., 2006), it is very difficult to determine absolute

0016-7037/$ - see front matter Ó 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.gca.2007.11.037
Hydrocarbon migration through the Clair field 2511

dates for oil generation, expulsion and migration. Mark throughout the WOS region. Using core samples from with-
et al. (2005) using temperature–composition–time data in the WOS region, the Devonian-Carboniferous Clair field
determined from authigenic K-feldspar (Mark et al., (block 206; Fig. 1), we applied the methodology developed
2006) managed to directly date hydrocarbon migration, res- by Mark et al. (2005, 2006) to determine temperature–com-
ervoir charging and reservoir breach within the UK Atlan- position–time (TXt) data from authigenic K-feldspar to
tic margin to within 3 Ma. reconstruct the oil-charge history of the Clair reservoir.
Hydrocarbon generation is dependant upon burial-re- Apatite fission track analysis (AFTA) and vitrinite reflec-
lated increases in temperature, but may also be affected tance (VR) data were also utilised to provide a regional
by pulses of hot fluid flow, evidence for which is widespread framework for understanding the thermal history and
across the UK Atlantic margin (Parnell et al., 2005). The hydrocarbon generation history of Orcadian Basin lacus-
hydrocarbon source rocks for a large portion of oil shows trine facies throughout Caithness, Orkney and Shetland
throughout WOS are Middle Jurassic and Kimmeridge and the adjacent offshore region.
Clay Formation (KCF) equivalent facies that reside in the
FSB (Bailey et al., 1987). Industry models suggest that oil 1.1. Clair field
generation within the FSB began during the Aptian-Albian
(115 Ma) transition with WOS reservoir-charging taking The Devonian-Carboniferous reservoired Clair field is
place since the Late Cretaceous (Carr and Scotchman, located 75 km west of the Shetland Isles (Fig. 1) within a
2003; Scotchman et al., 2006). Although pre-Mesozoic remnant of the Devonian Orcadian Basin. The reservoir is
source rocks have been considered as viable hydrocarbon 800 m thick and overlies a NE-SW trending ridge (The
sources within the WOS region (Scotchman et al., 1998), Rona Ridge) of metamorphic basement (Fig. 1). The reser-
a lack of geochemical evidence (i.e. the absence of Devo- voir is composed of Upper ORS–Lower Carboniferous aeo-
nian biomarkers) in reservoirs has resulted in rejection of lian–fluviatile–lacustrine red-bed facies (Allen and Mange-
the hypothesis (Scotchman et al., 1998). Rajetzky, 1992) deposited between 330 and 380 Ma. The
The study described in this paper was undertaken in or- two lower cycles of the Clair Group are separated from
der to determine the timing of the first oil-charge event the upper cycle by an important field-wide unconformity.

Fig. 1. (A) Map showing regional extent of the Devonian Orcadian Basin (palaeogeographical reconstruction modified from Duncan and
Buxton, 1995) and position of the Clair field and Beatrice field relative to mainland Scotland. (B) Area enlarged from (A) showing Clair field
outline and sample well locations relative to the field. (C) Cross-section (line of section from previous panel) showing fractured nature of the
Devonian-Carboniferous reservoir and well locations relative to field structure.
2512 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

Late Cretaceous shales unconformably overlie the Carbon- where Jurassic source rocks are still immature. Fractured
iferous strata and seal the reservoir. The Clair field is dom- crystalline basement found on Orcadian basin boundary
inated by two main structural sets, NE-SW oblique faults highs, such as that found throughout the Clair field (The
which truncate The Rona Ridge and ENE-WSW normal Rona Ridge), also becomes a much more prominent explo-
fault segments. Due to the extensively cemented nature of ration target throughout the UK Atlantic margin.
the reservoir, open fractures that trend NNW-SSE and
WNW-ESE are integral to the field’s reservoir potential 1.2. Orcadian basin
(Coney et al., 1993). The reservoirs contain an estimated
3–5 billion barrels of undersaturated oil in place (Scotch- In the north of Scotland collapse of over-thickened Cal-
man et al., 1998). edonian lithosphere following the Scandian episode of the
Rooney et al. (1998) and Scotchman et al. (1998) used Caledonian Orogeny resulted in normal reactivation of Cal-
gas-chromatography, biomarkers, and compound specific edonian thrust zones culminating in northwest–southeast
isotope analysis to demonstrate that the low API gravity extension. Rifting produced many linked half-graben basins
(22–26°) Clair hydrocarbons are a mixture of different oils that are collectively known as the Devonian Orcadian Ba-
that originated from various sources. The origin of an oil sin. The Orcadian Basin extends from the borders of the
accumulation is difficult to establish when the oil is a mix- Moray Firth in the south to Shetland in the north and Nor-
ture from multiple sources, or when more than one episode way in the east (Fig. 1). Offshore reconstruction of the Or-
of charging is involved, but these studies identified two cadian Basin boundaries is extremely poor and currently
main groups. The first oil is heavy, extensively degraded, based on lithological data from wells that have penetrated
and originated from type I non-marine, lacustrine kerogens Devonian sedimentary rocks. On average, the Orcadian Ba-
at low maturities. The second oil is relatively undegraded sin contains a 3–4 km fill of Devonian continental deposits
and was generated from a more refractory type II/III kero- that are collectively termed the Old Red Sandstone (ORS).
gen from a KCF equivalent source rock. The first oil was Broadly the ORS can be divided into three groups: Lower,
deposited in a more oxic environment with a stronger ter- Middle and Upper. Lower ORS conglomeratic facies
restrial input than the second oil. Preliminary geochemical unconformably reside on metamorphic basement and were
data based on isotopic signals of n-alkanes generated by deposited by alluvial fans that developed at fault scarps
hydropyrolysis (hypy) of asphaltenes (Russell et al., 2004) along basin margins (Trewin, 1989). Middle ORS facies at
isolated from Clair crudes that are most tightly bound to Caithness and Orkney are dominated by a thick sequence
the oil stained cores (well 206/8-8 and 206/8-7), show that of organic-rich cyclic lacustrine deposits, whereas in Shet-
there is at least one additional group of oils within the Clair land and Moray Firth, fluvial sedimentary rocks predomi-
reservoir and potentially several more. It is not known if the nate. The fluvial sediments are punctuated by numerous
isolated fraction is representative of the first oil charge to lacustrine intervals reflecting periodic lake transgression
enter the Clair reservoir; it is however representative of across surrounding alluvial plains (Trewin, 1989). The most
the first oil to be bound to the reservoir detritus (Russell widely developed of the lacustrine intervals is the Achanar-
et al., 2004). The n-alkanes generated by the asphaltene ras horizon which contains distinctive fish fauna and due to
hypy for both the oil and the reservoir cores are signifi- its wide regional extent, is the key to basin-wide correlation
cantly lighter than would be expected if, as currently be- (Hillier and Marshall, 1992). Upper ORS facies are domi-
lieved, the early charges were solely sourced from the nantly fluvial deposits that were deposited by braided river
Jurassic-Cretaceous source rocks in the FSB. The prelimin- system that flowed from the southwest (Trewin, 1989).
ary data suggest that oil from pre-Jurassic terrestrial Middle ORS organic-rich lacustrine facies are of great
sources has contributed to the charging history of the Clair interest as potential hydrocarbon source rocks especially
reservoir. in the Orkney and Caithness regions. Several studies have
A similar story charting exploration-discovery unfolded mapped present day maturation distributions of organic-
with respect to the Beatrice oil field within the Inner Moray rich lacustrine sedimentary rocks throughout the Orcadian
Firth (Fig. 1). Although assumed to be charged with oil Basin in an attempt to understand basin evolution and as-
from Jurassic-Cretaceous sources, geochemical analysis of sess potential for hydrocarbon generation. These studies
bitumens within the Beatrice field reservoir showed that (Parnell, 1985; Irwin and Meyer, 1990; Hillier and Mar-
Beatrice-oil could not have been generated solely from shall, 1992; Duncan and Buxton, 1995) have shown an ex-
KCF source rocks (Peters et al., 1989). The oil is in fact a treme range in maturity across the Orcadian Basin where
mixture of products derived from effective Devonian lacus- VR values (records of maximum palaeotemperature) range
trine facies and Middle Jurassic source rocks. Recognition from 0.5% to 10.5% Ro (Hillier and Marshall, 1992). Within
of effective Devonian source rocks in the Inner Moray Firth this range in maturity, the overall pattern is one of sharply
has many implications for regional hydrocarbon explora- juxtaposed regions of high and low maturity.
tion strategies within areas surrounding the Orcadian Ba- Rift basins such as the Orcadian Basin are characterised
sin, notably the UK Atlantic margin. For example, the by high but variable geothermal gradients and above aver-
realisation that Devonian source rocks have passed through age levels of organic maturation at shallow depths. For
the oil window during the Late Palaeozoic, Mesozoic and example, Tertiary rocks buried between 300 and 1500 m
Cenozoic, raises the possibility that if such source rocks in the Rhynie Graben witness present day geothermal gra-
are as extensive offshore as suspected, hydrocarbons might dients that range between 30 and 90 °C km 1 and have VR
be found in very old trap structures throughout regions values of 0.8% Ro (Roberts, 1985). In basins with normal
Hydrocarbon migration through the Clair field 2513

levels of heat flow (i.e. 20–30 °C km 1), 0.8% Ro would ments was noted by Pay et al., 2000, this was not evident
correspond to burial depths of approximately 5000 m (Rob- in samples analysed throughout his study). All samples con-
erts, 1985). With generally high geothermal gradients, re- tain various clay minerals including smectite, chlorite, illite
gions within the Orcadian Basin must have undergone and kaolinite as well as calcite, quartz, K-feldspar, fluorite,
very early maturation. Reservoired hydrocarbons which pyrite and marcasite cements. Clay minerals and calcite are
have been baked in situ by Permian dykes in Orkney (Astin, the dominant cementing phases. Quartz and K-feldspar ce-
1990) suggest that oil generation and migration had already ments mainly occur as overgrowths on detrital quartz and
occurred by Late Carboniferous times (290 Ma), the dyke K-feldspar grains, respectively (Fig. 2). Quartz overgrowths
swarms themselves having little effect on maturation. Direct occasionally contain small inclusions of K-feldspar cement
evidence for the timing of cooling is provided in Caithness at the contact between detrital quartz grains and their over-
and Orkney by AFTA (well 12/16-1) which reveal that cool- growths. This relationship suggests that K-feldspar cemen-
ing from palaeotemperatures in excess of 110 °C began dur- tation predated quartz overgrowths. In all samples
ing the Late Carboniferous (Green et al., 1995). This study authigenic quartz and K-feldspar pre-date calcite cementa-
also confirms that geothermal gradients at this time were tion but post-date deposition of clay cements. Calcite ce-
elevated. ments are coarsely-crystalline and in places envelop
detrital grains, producing a poikilitopic texture. Based on
2. SAMPLE DESCRIPTION cathodoluminescence properties, calcite veins and cements
can be divided into two types: (1) dull to moderately or-
The samples used for fluid inclusion work during this ange-luminescing calcite (C1) which is slightly ferroan in
study come from wells 206/8-8, 206/8-7, and 206/8-1a composition; and (2) bright orange-luminescing calcite
(Fig. 1). Baron et al. (2008) provides a comprehensive (C2) which is non-ferroan (Fig. 2). C1 always predates C2
description of samples from well 206/8-8, 206/8-7 and (Fig. 2). Small deformation bands (up to 1 cm) are com-
206/8-1a and a summary is presented here. Rock samples posed of fine angular detrital grains whose sizes are signif-
are pale orange-colored subarkosic sandstones. Most of icantly smaller than in the surrounding rock. Detrital grains
the detrital grains are quartz and feldspar with rare lithic with fractured overgrowths occur within some of the defor-
fragments (although the presence of abundant lithic frag- mation bands, indicating that K-feldspar and quartz

Fig. 2. (A) Transmitted light photomicrograph showing a K-feldspar overgrowth (arrow) enveloping a detrital feldspar grain (well 206/8-8,
1950 m). (B) SEM-SE image showing authigenic quartz overgrowth (white arrow) developing on a detrital quartz grain that is enveloped by
illite (black arrow; well 206/8-7, 2048.75 m). (C) Cold-CL image showing fracture-fill relationship between ferroan-rich C1 calcite and C2
calcite (well 206/8-7, 2025.5 m). (D) TEM image showing subgrain structure of K-feldspar overgrowth at sub-micron scale. A dislocation-rich
(semi-coherent) boundary (arrow) separating two subgrains (well 206/8-8, 2134.75 m).
2514 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

cementation pre-dated their formation. The deformation induced correction factors used were: (39Ar/37Ar)Ca = 0.00065,
bands are cemented by unbrecciated C1, indicating that (36Ar/37Ar)Ca = 0.000264, (40Ar/39Ar)K = 0.0085.
the deformation bands pre-dated C1 precipitation. Veins Mark et al. (2005, 2006) provide further details concerning the
filled by C2 also cross-cut, and clearly post-date, the defor- integration of fluid inclusion and Ar–Ar data.
mation bands.
3.3. Apatite fission track analysis

3. METHODS The principles involved in using AFTA to constrain thermal


histories in sedimentary basins are described in detail elsewhere
3.1. Fluid inclusion petrography and microthermometry (Green et al., 1989a,b, 2001, 2002, 2004; Parnell et al., 1999, 2005).
In brief, optical microscope analysis of radiation damage features
Hundred micrometers thick doubly polished fluid inclusion in detrital apatite grains separated from clastic units allows deter-
wafers were characterised using a standard Nikon ECLIPSE E600 mination of the maximum temperature to which a sedimentary unit
microscope fitted with an ultra-violet (UV) light source. Microth- has been heated (up to a maximum of around 110–120 °C,
ermometry was performed using a calibrated (associated error, depending on apatite composition) and the time at which the
±1 °C) Linkam TH-600 fluid inclusion stage with a heating rate of sample began to cool from that palaeothermal maximum. Inte-
10 °C/min 1 and employing standard methods (Shepherd et al., gration of AFTA and VR data allows extension of the thermal
1985). The fluid inclusions had consistent sizes and shapes of va- history constraints to higher palaeotemperatures.
pour bubbles and vapour–liquid ratios (Goldstein, 2001). Inclu- Definition of the variation of palaeotemperature with depth
sions that showed stretching were excluded from the study (Munz, provides insights into mechanisms of heating and cooling, distin-
2001). Aqueous fluid inclusions occur coevally with hydrocarbon guishing between heating due to deeper burial, elevated basal heat
inclusions and therefore are gas saturated and require no correlated flow and lateral flow of hot fluids (Bray et al., 1992). Linear pal-
pressure correction (Hanor, 1980). Homogenization temperatures aeotemperature profiles provide direct constraints on palaeogeo-
(Th) are therefore indicative of the true temperatures of fluid thermal gradients and extrapolation to an appropriate palaeo-
entrapment. Final ice melting temperature (Tm) measurements, surface temperature allows determination of the corresponding
which are dependent on the quantity of salt present in solution, amounts of removed section required to explain the observed pal-
were determined using a heating rate of 1 °C/min 1. Salinities were aeotemperatures (Green et al., 2004; Japsen et al., 2005; Holford
estimated using the methods of Bodnar (1993). et al., 2005).
It is important to appreciate that it is very uncommon for fis-
3.2. Ar–Ar geochronology sion track ages to directly indicate the timing of a discrete cooling
event. Rather, the fission track age represents an integrated mea-
In addition to conducting fluid inclusion analysis and mic- sure of the balance between the production of tracks by sponta-
rothermometry (Parnell et al., 1999), the natural radiogenic decay neous fission and the reduction in track length due to the thermal
of 40K to 40Ar in K-feldspar permits high resolution Ar–Ar age history. In order to extract thermal history constraints from the
determination (Mark et al., 2005). Integration of fluid inclusion data, a forward modelling approach is used to define the range of
analysis and microthermometry with UV laser ablation Ar–Ar histories for which predicted parameters (fission track age and the
dating (Kelley et al., 1994) will yield TXt data (Mark et al., 2006). distribution of track lengths and their variation with wt.% Cl) are
TXt data can constrain the timing, duration, temperature and consistent with measured data. It is however, not possible to extract
composition of fluid flow (Mark et al., 2007), providing that the explicit thermal history solutions directly from AFT data because
Ar-isotope system has not been thermally disturbed (Lee and of the high degree of redundancy in the data—i.e. many histories
Parsons, 2003). result in the same measured age and length parameters. Instead, we
Despite the fact that previous investigations (Hagen et al., 2001) focus on determining the maximum palaeotemperature and the
have yielded meaningful Ar–Ar ages for authigenic K-feldspar, it onset of cooling, using an approach based on likelihood theory
has been suggested that authigenic K-feldspar does not quantita- similar to that described by Gallagher (1995) to define best-fit
tively retain Ar at low temperatures (Lee and Parsons, 2003). values and corresponding 95% confidence intervals.
Therefore, the retentiveness of authigenic K-feldspar with respect
to Ar had to be quantified in detail in order to support the Ar–Ar 3.4. Vitrinite reflectance
data. The microstructures of authigenic K-feldspar overgrowths
were characterised by high-resolution imaging including transmis- VR based on the increasing reflectance of organic macerals
sion electron microscopy (TEM), and Ar-loss was modelled with with increasing temperature, is commonly used in hydrocarbon
respect to subgrain structure and Ar diffusion using reconstructed exploration as an indicator of thermal maturity (Tissot and Welte,
thermal histories. 1984). VR data can be used to provide independent estimates of
Prior to irradiation, a 100 lm thick doubly polished fluid maximum palaeotemperatures, through knowledge of the reaction
inclusion wafer was cleaned ultrasonically in methanol and kinetics, thereby providing a useful complement to the informa-
deionised water. The sample was cadmium shielded and irradiated tion derived from AFTA. A VR value of 0.7% corresponds clo-
for 50 h in the Canadian McMaster reactor. Neutron flux was sely to total annealing of fission tracks in apatite (Duddy et al.,
monitored with biotite standard GA1550 (98.8 ± 0.5 Ma; Renne 1994).
et al., 1998); a calculated J value of 0.01175 ± 0.00006 was used. A In recent years it has become apparent that VR data generated
New Wave Research UP-213 nm pulsed Nd-YAG laser (266 nm) by different analysts are not equivalent (see discussion in Green
with a 12 lm spot size was used for Ar-extraction. Extracted gases et al., 2002). We find a high degree of consistency between thermal
were cleaned using three SAES AP10 getters, two operated at history interpretations from AFTA and VR data generated using
450 °C and the other at room temperature. A MAP 215-50 noble an approach (Cook, 1982) recommended by the International
gas mass spectrometer analysed Ar-isotope compositions. The data Commission on Organic and Coal Petrography (www.iccop.org).
were corrected for blanks, mass spectrometer discrimination, 37Ar This involves measurement of maximum reflectance under oil
decay and reactor induced interferences. Quoted Ar–Ar errors are (Romax) in polished thick sections, with identification of the
2r and include a 0.5% error assigned to the J value. Reactor indigenous vitrinite population being made on textural grounds.
Hydrocarbon migration through the Clair field 2515

These data then enable an independent assessment to be made of 0.97) of liquid. Most hydrocarbon inclusions fluoresce in
the possible presence of reworked vitrinite populations from pet- UV light. These inclusions were confirmed as hydrocar-
rographic evidence, as well as allowing identification of material bon-bearing by freezing to less than 80 °C. The dominant
introduced by caving from higher levels in the borehole. The data fluorescence colour of the hydrocarbon inclusions is yel-
presented here were generated using the ICCOP potocols.
low–orange, with subordinate quantities of blue- and
In contrast, VR data encountered in hydrocarbon industry re-
ports are often based on measurements of random reflectance
white-fluorescing inclusions also occurring. Only blue-fluo-
(Rorand) in strewn slides of organic concentrates, with the indige- rescing primary hydrocarbon inclusions occur within K-
nous vitrinite population often identified by inspection of histo- feldspar overgrowths. These inclusions are believed to be
grams of measurements and separation into perceived sub- representative of the first recorded oil charge to enter the
populations. This approach tends to give lower values compared to Clair Group. Primary aqueous fluid inclusions occur coev-
measurement of Romax as described above (see Green et al., 2002 ally with the hydrocarbon fluid inclusions. Aqueous inclu-
for further discussion). Therefore, in converting VR data to max- sions are two-phase (liquid and vapour), range in size
imum palaeotemperatures it is essential to use a kinetic model from 2 to 10 lm and have high volumetric proportions
which is appropriate to the analytical approach adopted. Experi- (0.90–0.97) of liquid. The complete fluid inclusion microth-
ence in a variety of different settings suggests that the Romax ap-
ermometric results are summarised in Table 1 and Fig. 3.
proach described above yields estimates of maximum
palaeotemperature that are highly consistent with those from
Blue-, yellow–orange-, white-, and non-fluorescing hydro-
AFTA, if the Burnham and Sweeney (1989) algorithm for the ki- carbon inclusions within authigenic quartz and calcite yield
netic response of VR is used. very similar Th values. Th values of primary aqueous inclu-
sions hosted in quartz overgrowths range from 111.3 to
4. RESULTS 170.5 °C. Th values determined from fluid inclusions within
K-feldspar overgrowths, which are considered from petro-
The focus of this study is determination of TXt data graphic evidence to pre-date entrapment of the quartz-
from authigenic K-feldspar via the integration of fluid hosted inclusions, are slightly lower. These values range
inclusion petrography and microthermometry with high- from 90.2 to 116.3 °C. Th values of primary inclusions
resolution Ar–Ar dating. Hence, only a detailed description hosted in C1 cements and veins range from 129.4 to
of the authigenic K-feldspar is presented below. A brief 170.0 °C. Th values of primary inclusions hosted in C2 ce-
summary of the petrography of the quartz and calcite ce- ments and veins are slightly lower, ranging from 118.6 to
ments is given in Section 2, and readers are directed to Bar- 159.4 °C. The average Th values of the majority of the
on et al. (2008) for further detailed descriptions. hydrocarbon fluid inclusions are lower than the associated
coeval aqueous fluid inclusions. The range of salinities of
4.1. Authigenic K-feldspar the aqueous inclusions indicates that mixing of fluids with
differing compositions occurred during diagenesis. The low-
Samples contain 20 to 40 lm K-feldspar overgrowths er salinity aqueous fluids are similar in composition to
that are developed around detrital grains of plagioclase meteoric (<0.5 wt.% NaCl) and/or marine (3.5 wt.% NaCl)
and K-feldspar (Fig. 2). The overgrowths are untwinned waters. The higher salinity aqueous fluid is typical of basi-
and in optical discontinuity with the enveloped detrital nal brines, whose salinity has been derived from extensive
grain. Complex extinction patterns when viewed in trans- water–rock interactions or precipitated directly from deep
mitted light between cross polarisers suggest that the over- saline fluids. Rare 3-phase inclusions that are filled by a
growths are composed of slightly misorientated subgrains. hydrocarbon liquid phase, an aqueous liquid phase and a
Diffraction contrast TEM imaging shows that K-feldspar vapour phase, occur within primary and secondary settings
overgrowths are composed of mosaics of misaligned sub- in K-feldspar, quartz and calcite cements and fractured
grains separated by dislocation-rich boundaries and detrital grains in the Clair reservoir. The proportions of
micropores (Fig. 3). Subgrains are finely mottled and have oil, vapour and aqueous liquid within the 3-phase inclu-
consistent sizes of between 1 and 2 lm and so are compara- sions is extremely variable. The volumes of oil contained
ble to subgrains reported in previous investigations (Wor- within fluid inclusions ranges from 3% to 74% (n = 31) total
den and Rushton, 1992; Lee and Parsons, 2003; Mark inclusion volume. The extreme variability indicates that the
et al., 2005). precipitating fluid was not a homogenous mix of oil and
water, but a heterogeneous fluid (Goldstein, 2001). At
4.2. Fluid inclusion petrography and microthermometry Earth surface conditions (pressure and temperature) oil
and aqueous fluids do not readily mix and hence a suite
Both aqueous and hydrocarbon inclusions occur in all of fluid inclusions containing oil and aqueous fluids would
samples analysed. Hydrocarbon fluid inclusions occupy pri- possess variable oil–water ratios.
mary settings (the fluid was entrapped during mineral
growth) within K-feldspar overgrowths (Fig. 3), quartz 4.3. Ar–Ar geochronology
overgrowths, C1 cement and veins, C2 veins, but not C2 ce-
ment, suggesting that during C2 precipitation fluid predom- Ar–Ar ages determined from detrital feldspar grains
inantly migrated through open fracture systems rather than range from 424 ± 19 to 1743 ± 428 Ma (n = 5). Variability
inter-granular pore networks. Hydrocarbon inclusions are in the detrital feldspar ages is potentially due to a combina-
mainly 2-phase (liquid and vapour), range in size from 2 tion of factors: (1) significant replacement of the detrital
to 50 lm and have high volumetric proportions (0.90– grains by feldspar that crystallized later and at lower tem-
2516 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

Fig. 3. (A) Transmitted light fluid inclusion photomicrograph. Primary 2-phase aqueous fluid inclusion (arrow) trapped within a K-feldspar
overgrowth (OG) that is enveloping a detrital feldspar grain. (B) Enlargement of (A). Both the liquid and vapour phases within the fluid
inclusion are clearly visible. (C) Transmitted light fluid inclusion photomicrograph. Primary 2-phase oil inclusion (arrow) trapped within a K-
feldspar overgrowth. (D) UV light photomicrograph. Oil inclusion from (C) enlarged and clearly showing a blue fluorescence (arrow). (E)
Cross-plot showing total aqueous fluid inclusion data set for Clair reservoir; temperature of fluid entrapment (homogenization temperature)
versus fluid salinity (converted from final ice melting temperature). Plot shows salinity of meteoric water, sea water, and basinal brines relative
to the data. Clearly mixing of fluids from different reservoirs is required to produce the distribution observed.

peratures; (2) potential Ar-loss from subgrained micro- 0.73). Scatter within the K-feldspar overgrowth Ar–Ar data
structures within the feldspar crystal; (3) extraction of Ar set may also be attributed, to a lesser degree, to ablation of
gas from the detrital grains via laser ablation in volumes populations of subgrains of different size ranges, as deter-
too small to yield reliable ages (Table 2) and (4) real varia- mined from TEM imaging (Fig. 2). The temperature at
tion in source ages. Ar–Ar ages from optically identifiable which Ar-loss occurs is directly related to subgrain size,
K-feldspar overgrowths (Table 2) range from 230 ± 70 to with Ar closure temperatures being lower for the smaller
283 ± 58 Ma (n = 9). Owing to the small size of the over- subgrains. For example, an ablation area containing ten
growths it was extremely difficult to extract sufficiently large 2 lm subgrains and only two 1 lm subgrains would poten-
volumes of argon from them to ensure a mass spectrometer tially yield a slightly older Ar–Ar age than an ablation area
signal greater than 10 times that of the blank levels (Mark containing ten 1 lm subgrains and two 2 lm subgrains be-
et al., 2006) which would minimise analytical error. These cause more Ar will be lost from the sample with the greater
potential analytical difficulties are the source of the large er- ratio of small to large subgrains.
rors associated with most of the K-feldspar overgrowth Ar–
Ar ages in Table 2. Although only small amounts of Ar gas 4.4. AFTA and VR
were extracted from some K-feldspar overgrowths, the ratio
of 40Ar to 39Ar is clearly consistent regardless of gas vol- AFTA and VR data from samples of four wells in the
umes (Table 2). The data yield a weighted mean of vicinity of the Clair field (206/8-8, 1750 m; 206/8-7,
247.7 ± 3.3 Ma (2r, mean square weighted deviation of 2075 m; 206/10-2, 366–2918 m; 206/9-2, 658–2429 m;
Hydrocarbon migration through the Clair field 2517

Table 1
Fluid inclusion microthermometric data from primary inclusions
Well Depth (m) Authigenic mineral phase Homogenization temperature (°C) Salinity (wt.% NaCl eq.)
H/C n Aq n
206/8-7 2000 K-feldspar 88.1-96.7 20 93.2–111.5 12 4.03–6.59
Quartz 107.3–134.8 6 118.4–144.6 7 0.53–2.41
Calcite (C1c) 133.6–144.2 4 140.6–163.4 3 1.4–3.23
Calcite (C1v) 148.7–167.8 8 135.3–161.9 6 2.57–4.65
Calcite (C2c) — 0 118.6–143.5 7 2.07–3.87
Calcite (C2v) 121.3–154.6 5 120.8–149.3 6 2.41–4.03
206/8-la 2041.4 K-feldspar 94.3–108.7 11 90.2–116.3 14 3.06–6.01
Quartz 105.4–143.2 8 111.3–151.6 6 0.71–1.91
Calcite (C1c) 128.3–156.1 6 148.5–163.2 8 1.74–4.03
Calcite (C1v) 136.7–162.8 4 143.9–170.0 7 2.24–5.26
Calcite (C2c) — 0 124.6–148.2 9 1.91–3.73
Calcite (C2v) 128.6–141.9 7 119.3–151.4 8 2.57–4.65
206/8-8 2134.75 K-feldspar 94.5–99.6 6 90.5–101.2 11 6.06–6.74
Quartz 132.5–148.6 8 139.8–170.5 7 1.57–4.80
Calcite (C1c) 138.7–151.2 5 148.3–164.9 6 1.91–3.23
Calcite (C1v) 132.4–153.2 7 138.6–161.4 5 2.74–4.8
Calcite (C2c) — 0 126.3–159.4 5 1.05–3.39
Calcite (C2v) 119.7–137.6 6 124.8–143.3 5 2.9–4.49
206/8-8 2158.5 K-feldspar 78.6-101.4 18 93.7–108.2 12 4.49–7.02
Quartz 137.3–151.4 6 123.6–144.5 6 2.07–2.9
Calcite (C1c) 124.6–144.8 7 133.3–156.8 4 1.91–4.03
Calcite (C1v) 134.6–159.1 5 129.4–155.3 6 1.4–3.26
Calcite (C2c) — 0 141.4–153.8 6 1.74–3.87
Calcite (C2v) 125.3–140.0 4 130.9–137.4 5 2.74–5.11
206/8-8 2160.85 K-feldspar 84.5–97.9 19 93.6–104.8 7 5.56–6.74
Quartz 134.5–155.5 4 134.2–165.1 6 2.57–4.03
Calcite (C1c) 132.9–146.0 6 128.4–151.5 5 1.05–2.41
Calcite (C1v) 141.1–163.2 5 145.8–163.4 5 1.91–3.55
Calcite (C2c) — 0 136.2–155.7 7 1.91–2.57
Calcite (C2v) 130.6–145.2 3 138.4–149.2 8 2.9–5.28
Salinity calculated using equations of Bodnar (1993). Homogenization occurred via bubble point transition. H/C, hydrocarbon-bearing
inclusion data; Aq, aqueous fluid inclusion data; x, no fluid inclusions present.

Fig. 1) are summarised in Tables 3 and 4, respectively. Fis- the values predicted from the respective default thermal his-
sion track ages and mean track lengths are plotted against tories, and detailed assessment shows that the track length
depth and present-day temperature in Fig. 4. The variation data in these samples can only be explained if these samples
of stratigraphic age with depth is also shown for each well, have been hotter in the past. In contrast, fission track ages
together with the variation in fission track age and mean measured in all samples from the Palaeozoic section in these
track length predicted from the default thermal history sce- four wells are younger than the respective stratigraphic age
nario in each well. As explained in more detail elsewhere and are also significantly less than the values predicted from
(Green et al., 2004; Parnell et al., 2005), this represents the respective default thermal history, as are the mean track
the situation in which all sedimentary units are currently lengths in most cases. These observations show that the
at their maximum post-depositional temperatures, and data cannot be explained by the default thermal history sce-
forms the basis for extraction of information on the palae- nario, and therefore the sampled sedimentary units must
o-thermal history from the AFTA and VR data. have been hotter than the present-day temperatures at some
Measured fission track ages in samples of Mesozoic and time after deposition.
Cenozoic units in the 206/9-2 and 206/10-2 wells are older
than respective stratigraphic ages, showing that these sam- 4.4.1. Thermal history constraints from AFTA and VR
ples are dominated by tracks formed prior to deposition of Maximum palaeotemperatures attained by each sample
the host sediments. In most of these samples, the AFTA and the time at which the sample began to cool from that pal-
data show no definite evidence of post-depositional heating, aeo-thermal maximum have been determined using proce-
and provide only an upper limit to the maximum post- dures explained in detail elsewhere (Green et al., 2004;
depositional palaeotemperature (Table 3). However, in Parnell et al., 2005), with results summarised in Table 4.
two samples from the Late Cretaceous section in well Two episodes of cooling can be resolved from the data in most
206/10-1, mean track lengths are more than 2 lm less than samples from the Palaeozoic section. Resolution of multiple
2518 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

Table 2
Ar–Ar isotopic data for K-feldspar overgrowths (OG), detrital grains (DG), hybrid analyses (Hy) and average blanks (1012 cm3 STP)
40 39 38 37 36 40
Well & depth (m) Sample No Ar ArK ArCl ArCa Ar Ar*/39Ar Age (Ma) ±
206/8-7 (2000) OG1 37.493 1.201 0.008 0.181 0.076 12.601 249.1 39.1
206/8-7 (2000) OG2 181.247 14.524 0.206 0.000 0.000 12.480 246.9 3.5

Average blank (n = 3) 7.535 0.076 0.004 0.004 0.013 — — —

206/8-8 (2134.75) OG3 20.899 1.772 0.000 0.609 0.000 11.805 233.8 39.2
206/8-8 (2134.75) OG4 16.888 1.163 0.000 2.503 0.000 14.495 283.1 57.9
206/8-8 (2134.75) OG5 40.186 3.083 0.059 1.835 0.000 13.034 256.5 22.4
206/8-8 (2134.75) OG6 33.810 2.562 0.055 1.415 0.000 13.195 259.4 19.9
206/8-8 (2134.75) OG7 8.526 0.735 0.034 1.684 0.000 11.595 229.9 70.5
206/8-8 (2134.75) OG8 12.827 0.907 0.013 0.239 0.000 14.112 276.1 56.1

Average blank (n = 7) 7.732 0.063 0.105 2.495 0.008 — — —

206/8-la (2041.4) OG9 20.761 1.613 0.050 0.021 0.000 12.879 253.6 20.2
206/8-8 (2134.75) Hy1 17.363 1.096 0.055 0.441 0.000 15.817 307.3 29.1
206/8-8 (2134.75) Hy2 41.332 2.071 0.034 0.109 0.000 19.970 379.6 23.5
206/8-8 (2134.75) Hy3 52.744 2.633 0.109 0.441 0.000 20.041 381.7 24.0

Average blank (n = 4) 6.817 0.076 0.059 2.575 0.004 — — —

206/8-7 (2000) DG1 385.186 4.498 0.063 0.475 0.000 85.603 1253.4 12.8
206/8-7 (2000) DG2 90.980 4.024 0.097 0.290 0.000 22.619 424.1 19.3

Average blank (n = 3) 7.505 0.118 0.071 2.352 0.008 — — —

206/8-8 (2134.75) DG3 6.296 0.164 0.021 2.873 0.000 38.694 675.7 159.2
206/8-8 (2134.75) DG4 12.050 0.088 0.046 0.017 0.000 138.835 1743.6 428.0

Average blank (n = 3) 5.947 0.227 0.092 2.848 0.004 — — —

206/8-la (2041.4) DG5 13.570 0.584 0.000 1.142 0.000 23.166 433.8 54.1

Average blank (n = 2) 6.397 0.181 0.101 0.386 0.008 — — —

episodes in this way depends on the annealing of tracks the maximum palaeotemperatures estimated from AFTA
formed subsequent to the onset of cooling in an are highly consistent with those from VR data in adjacent
early cooling episode. The extraction of thermal history infor- samples.
mation from AFTA data in this way is illustrated in Fig. 5. Maximum palaeotemperatures between 100 and
Comparison of the thermal history constraints derived 125 °C derived from VR data in the Palaeozoic section in
from each AFTA sample in Table 4 shows that results from both wells are highly consistent with those from AFTA cha-
all samples from the Palaeozoic section in all four wells are racterising the Late Permian to Early Triassic palaeo-ther-
consistent with cooling from maximum palaeotemperatures mal episode, confirming that the Palaeozoic section in
above 100 °C at some time in the interval 270–230 Ma (Late these wells cooled from maximum post-depositional palae-
Permian to Early Triassic). Similarly, estimates for the tim- otemperatures in this episode (i.e. any hydrocarbon source
ing of the later cooling episode from all AFTA samples in rocks within this section reached their maximum maturity
the 206/8-7, 206/8-8 and 206/10-1 wells, as well as two sam- levels in this episode). AFTA and VR data are available
ples from the 206/9-2 well are consistent with cooling begin- only over a limited depth range in these two wells, and from
ning at some time between 25 and 10 Ma (Miocene). In only single samples in the 206/8-7 and 206/8-8 wells, but
contrast, the three deepest samples from the Palaeozoic sec- overall the available constraints for this episode in the
tion in well 206/9-2 suggest a distinct cooling episode began 206/9-2 and 206/10-1 wells are consistent with linear pro-
between 150 and 50 Ma (Table 4). files, as illustrated in Fig. 6. The form of these profiles sug-
Maximum palaeotemperatures derived from VR data in gests that a major contribution to heating in this episode
the 206/9-2 and 206/10-1 wells are also listed in Table 4. No was due to deeper burial, possibly combined with some de-
VR data are available from the 206/8-7 and 206/8-8 wells. gree of elevated basal heat flow (see later discussion).
Palaeotemperature constraints from both AFTA and VR AFTA data in most post-Palaeozoic samples provide
data in the 206/9-2 and 206/10-1 wells are plotted against only upper limits to the magnitude of Miocene palaeotem-
depth in Fig. 6, where it is clear that throughout each well peratures, but in two samples from the 206/10-1 well
Table 3
AFTA sample details and apatite fission track age and track length data summary for samples from four WOS wells
Sample Depth Stratigraphic Present temperature qD (106 qs (106 qi (106 Fission track ± (Ma) P(v2) (%)d Mean track l ± (lm)e
(m) age (Ma)a (°C) tracks/cm2)b tracks/cm2)b tracks/cm2)b age (Ma)c ength (lm)
206/8-8
RD87-1 1750 385–330 69 1.145 (1779) 2172 (1352) 2.026 (1261) 226 22.7 <1 (20) 11.26 0.14 (100)
206/8-7
RD87-2 2075 385–330 73 1.147 (1779) 1.151 (791) 1.205 (828) 190 19.5 <1 (20) 10.71 1.52 (100)
206/9-2

Hydrocarbon migration through the Clair field


GC540-89 1088 65–57 40 1.413 (1091) 0.962 (285) 1.235 (366) 208.2 17.9 <1 (15) 12.27 0.18 (74)
GC540-97 1798 363–323 65 1.417 (1091) 1.817 (884) 2.561 (1246) 290.4 10.5 21 (20) 10.52 0.17 (108)
GC540-98 1943 363–323 70 1.42 (1091) 2.187 (1211) 3.104 (1719) 289.5 9.5 <1 (20) 10.99 0.14 (111)
GC540-99 2195 409–363 78 1.423 (1091) 1.636 (1193) 2.687 (1960) 264.5 8.2 <1(20) 10.08 0.18 (112)
GC540-101 2345 >540 84 1.426 (1091) 1.408 (416) 2.443 (970) 116.6 7.9 <1 (20) 9.73 0.20 (100)
GC540-102 2429 >540 86 1.43 (1091) 0.732 (283) 2.109 (815) 94.8 7.3 24 (20) 9.25 0.22 (109)
206/10-1
GC540-109 1105 65–57 22 1.389 (1091) 0.439 (164) 0.559 (209) 206.3 22.6 39 (17) 11.42 0.36 (24)
GC540-112 1341 74–65 40 1.383 (1091) 2.169 (1104) 2.176 (1108) 259.8 14.1 27 (20) 12.76 0.14 (102)
GC540-119 1981 87–74 57 1.378 (1091) 2.306 (1164) 2.746 (1386) 218.8 11.4 <1 (17) 11.51 0.17 (82)
GC540-122 2188 89–74 63 1.372 (1091) 1.608 (508) 2.431 (768) 172.2 11.4 <1 (20) 10.53 0.26 (75)
GC540-123 2217 89–87 64 1.367 (1091) 2.299 (311) 3.178 (430) 187.3 15.3 17 (7) 9.71 0.49 (33)
GC540-126 2339 409–290 67 1.361 (1091) 1.269 (594) 1.595 (747) 204.8 13.2 <1 (20) 10.32 0.17 (100)
GC540-127 2454 409–290 70 1.356 (1091) 1.423 (590) 2.85 (1182) 128.8 7.8 <1 (18) 9.77 0.20 (107)
GC540-129 2067 409–290 75 1.35 (1091) 1.324 (708) 2.028 (1085) 167.2 9.8 <1 (20) 9.59 0.23 (104)
GC540-130 2918 409–290 82 1.344 (1091) 0.673 (339) 1.585 (799) 108.8 7.9 <1 (20) 10.56 0.19 (100)
NB, All analytical details are as described by Green (1986), with the exception that some thermal neutron irradiations show a significant flux gradient, in which case the appropriate value of PD
was determined by linear interpolation through the grains.
a
Stratigraphic ages taken from composite log. All numerical values for stratigraphic ages assigned following (Harland et al., 1989).
b
Numbers in parentheses show the number of tracks counted.
c
Central age (Galbraith and laslett, 1993) used for samples containing significant spread in single grain ages, otherwise pooled age and conventional (Poisson) error (Green, 1981). Ages
calculated using CN-5 reference glass with zeta values (Hurford and Green, 1982) value of 380.4 ± 5.7 for samples RD87-1,2 and 384.6 ± 5.5 for GC540 samples.
d
Numbers in parentheses show the number of grains counted.
e
Numbers in parentheses show the number of track lengths measured.

2519
Table 4

2520
Thermal history interpretation summary of AFTA and VR data in samples from four WOS samples
Sample Depth Stratigraphic Present Measured Paleozoic & Mesozoic episodes Cenozoic episodes
(m) age (Ma)a temperature VR (%)b Max palaeo- Onset of cooling Max palaeo- Onset of cooling
(°C)
temperature (°C)c from AFTA (Ma) temperature (°C)c from AFTA (Ma)
206/8-8
RD87-1 1750 385–330 69 – 90–125 325–160 75–90 160–0
206/8-7
RD87-2 2075 385–330 73 – 100–110 300–215 85–95 100–0
206/9-2

D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533


GC540-86.1 658 35–23 25 0.3 (25) – – <50 –
GC540-87.1 756 57–35 29 0.29 (25) – – <50 –
GC540-89 1088 65–57 40 – – – <95 90–0
GC540-89.1 1116 65–57 41 0.36 (20) – – 59 –
GC540-90.1 1256 74–65 46 0.41 (6) – – 66 –
GC540-92.1 1378 74–65 50 0.5 (5) – – 83 –
GC540-94.1 1500 74–65 54 0.56 (5) – – 93 –
GC540-95.1 1588 98–74 57 0.52 (5) – – 86 –
GC540-96.1 1646 89–74 59 0.61 (10) – – 100 –
GC540-97.1 1783 363–323 64 0.64 (1) 106 – – –
GC540-97 1798 363–323 65 – 95–105 270–120 80–90 50–0
GC540-98 1943 363–323 70 – 100–110 270–200 80–90 120–0
>120 310–230
GC540-99 2195 409–363 78 – 90–100 160–50 – –
GC540-100.1 2286 409–363 82 0.78 (2) 125 – – –
>110 400–200
GC540-101 2345 >540 84 – 90–105 150–50 – –
>100 500–200
GC540-102 2429 >540 86 – 90–100 150–110 – –
206/10-1
GC540-105.1 671 65–2 14 0.23 (25) – – <50 –
GC540-107.1 792 65–2 22 0.29 (20) – – <50 –
GC540-108.1 1006 65–57 25 0.3 (26) – – <50 –
GC540-109.1 1076 65–57 31 0.34 (26) – – 53 –
GC540-109 1105 65–57 22 – – – <100 post-depositional
GC540-109.2 1158 65–57 33 0.33 (14) – – 50 –
GC540-110.1 1177 74–65 35 0.36 (27) – – 59 –
GC540-112.1 1280 74–65 35 0.33 (26) – – 50 –
GC540-112 1341 74–65 40 – – – <85 post-depositional
GC540-112.3 1402 74–65 38 0.43 (8) – – 72 –
GC540-114.1 1554 74–65 41 0.48 (12) – – 80 –
GC540-116.1 1682 74–65 46 0.47 (21) – – 79 –
GC540-118.1 1878 87–74 49 0.49 (7) – – 81 –
GC540-119.1 1978 87–74 54 0.55 (8) – – 91 –
GC540-119 1981 87–74 57 – – – <105 post-depositional
GC540-119.2 2051 87–74 57 0.54 (11) – – 90 –
(continued on next page)
Hydrocarbon migration through the Clair field 2521

from AFTA (Ma)


(GC540-122 and -123) which provide a discrete range of
Onset of cooling palaeotemperatures in this episode, these are highly consis-
tent with those defined by VR data in adjacent samples (Ta-
ble 4). While there is no timing constraint on the rest of the

50–10
palaeotemperatures defined by VR data in Mesozoic and
50–0

70–0

30–0

25–0

40–0
Cenozoic units, it seems reasonable to assume that all these



palaeotemperatures represent the Miocene cooling episode
defined from AFTA. These palaeotemperatures, together
Cenozoic episodes

with Miocene values defined by AFTA data in the Palaeo-


temperature (°C)c

zoic section, clearly define arcuate profiles in Fig. 6, sugges-


Max palaeo-

tive of heating due to passage of hot fluids through the


Mesozoic section (cf. Ziagos and Blackwell, 1986). The re-
80–100

95–105
80–90

80–90

85–95

85–95
sults from the 206/9-2 well have been discussed previously
88
94

91

81

by Duddy et al. (1998) who proposed a similar interpreta-



– tion, and reported similar observations from the nearby


205/23-1 well, while the results from well 206/10-1 were re-
ported by Green et al. (1999). Further evidence for palaeo-
from AFTA (Ma)
Onset of cooling

thermal effects due to passage of hot fluids during Cenozoic


times in the WOS region has been presented by Parnell
Stratigraphic ages taken from composite log. All numerical values for stratigraphic ages assigned following (Harland et al., 1989).

et al. (1999, 2005), so such an interpretation is consistent


270–120

270–180

300–50

with regional evidence.


Paleozoic & Mesozoic episodes

>100

Mesozoic palaeotemperatures are identified from AFTA








only in the deepest samples from the 206/9-2 well. These


palaeotemperatures are interpreted as representing an inter-
mediate cooling episode during the long period of time rep-
temperature (°C)c

resented by the top-Palaeozoic unconformity in these wells


Max palaeo-

(Carboniferous to Late Cretaceous). The most likely expla-


nation for their recognition only in the 206/9-2 well is that
105–115
95–105

90–110

Thermal history interpretation has been carried out using methods outlined by Green et al. (2001).

while this cooling episode affected the Palaeozoic section in


>100
104

111

all four wells, its effects have been overprinted by the Late
99
97





Miocene hot fluid-related heating in the other three wells.


Thus, we interpret this episode as a regional cooling event
experienced by the Palaeozoic section along The Rona
Measured
VR (%)b

Ridge.
0.53 (1)
0.57 (4)

0.55 (7)

0.49 (6)

0.63 (1)

0.59 (1)

0.67 (1)
0.6 (2)

4.4.2. Thermal history reconstruction


Using the methods outlined by Bray et al. (1992), we


have estimated the range of palaeogeothermal gradients
temperature

and amounts of additional burial on the top-Palaeozoic


unconformity in the 206/9-2 and 206/10-1 wells that are
Present

Numbers in parentheses show the number of fields measured.

consistent with the Late Permian to Early Triassic palaeo-


(°C)

59
61
63
62
64
63
67
66
70
70
72
75
74
82

temperature constraints derived from AFTA and VR data


in these wells. Results are illustrated in Fig. 7. Given the
similarity between the palaeotemperatures characterising
this episode in the 206/8-7 and 206/8-8 wells and those at
Stratigraphic

similar levels in the other two wells, together with the sim-
age (Ma)a

409–290
409–290
409–290
409–290
409–290
409–290
409–290
409–290

ilarity of the geological setting of each well, we suggest that


87–74
89–87
89–74
89–74
89–74
89–74

a common interpretation is likely to apply to all four wells.


On the basis of results from Fig. 7, we favour an interpre-
tation of the Late Permian to Early Triassic palaeotemper-
atures in terms of deeper burial by around 2 km of
additional section, corresponding to a palaeogeothermal
Depth

2109
2167
2188
2207
2217
2316
2339
2444
2454
2533
2600
2067
2682
2918

gradient of 40 °C/km. This section must have been of


(m)

Carboniferous to Permian age, subsequently removed by


exhumation beginning in the interval 270–230 Ma.
Table 4 (continued)

The Mesozoic palaeotemperatures defined in the 206/9-2


GC540-120.1
GC540-122.1
GC540-122.1
GC540-122.3

GC540-125.1

GC540-127.1

GC540-128.1
GC540-129.1
GC540-129.1
GC540-129.2

well are interpreted as representing an intermediate phase


GC540-123

GC540-126

GC540-127

GC540-130

of cooling prior to the onset of Late Cretaceous deposition,


Sample

although we note that previous studies in the region have


revealed Early Cretaceous palaeo-thermal effects attributed
b
c
a
2522 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

Fig. 4. Fission track age and mean track lengths in samples from four WOS wells, plotted against depth and present-day temperatures. Also
shown for each well are the variation of stratigraphic age with depth, and the pattern of fission track age and mean track length predicted from
the default thermal history for each well, for apatites containing 0.0–0.1, 0.4–0.5, 0.9–1.0 and 1.5–1.6 wt% Cl. The default thermal history is
the history that would pertain if all units in each well are currently at their maximum post-depositional palaeotemperatures, which provides an
initial reference for interpreting the AFTA data in terms of likely palaeo-thermal effects. Present-day temperatures in each well were calculated
on the basis of corrected BHT values, as explained by Parnell et al. (2005).
Hydrocarbon migration through the Clair field 2523

Fig. 5. Thermal history solutions are extracted from the AFTA data by comparing modelled parameters from a range of thermal history
scenarios with measured values to define the range of conditions (maximum palaeotemperature and onset of cooling) for which the predictions
match the measured data within 95% confidence limits. The default thermal history for each sample provides the basic framework for this
analysis, with episodes of heating and cooling added as required in order to obtain a good match. An acceptable thermal history solution
should match both the measured fission track age and length distribution for the sample as well as the variation of these parameters with wt.%
Cl within each sample, as shown.

to hot fluid movements (Parnell et al., 1999), so this alterna- the Mesozoic and Cenozoic section due to hot fluid move-
tive explanation should also be borne in mind. ments during the Miocene are shown schematically in
The corresponding reconstructed thermal histories are Fig. 8. Note that since heating in this episode is attributed
illustrated in Fig. 8. The effects of transient heating within to transient fluid movements, the timescale of heating was
2524 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

Fig. 6. Palaeotemperature profiles in wells 206/9-2 and 206/10-1 derived from the AFTA and VR data in each well (palaeotemperatures are
listed in Table 4). The convex profiles defining the Cenozoic palaeotemperatures suggest that heating was due to the flow of hot fluids. In
contrast, Late Palaeozoic palaeotemperatures in each well can be described by linear profiles, consistent with heating due to a combination of
deeper burial and elevated basal heat flow.

probably much shorter than assumed in estimating the pal- to localised movements of hot fluids (Ziagos and Blackwell,
aeotemperatures quoted in Table 4, and for this reason the 1986), producing transient heating effects which are too
real palaeotemperatures are likely to have been much high- rapid to produce a detectable effect on the AFTA and VR
er than the quoted values. As a rough ‘‘rule of thumb”, each systems (Parnell et al., 2005).
order of magnitude increase in heating rate corresponds to Kinetic modelling of fission track annealing and VR
a 10 °C increase in the palaeotemperature required to pro- evolution (Parnell et al., 2005) shows that fluids with tem-
duce a given degree of annealing (Green et al., 1989b). peratures up to 200 °C and pulse durations less than
10 5 Ma have no detectable effect on AFTA or VR data,
4.4.3. Comparison of fluid inclusion and AFTA temperatures but are recorded and preserved by fluids entrapped in
Fluid entrapment temperatures at the time for K-feld- inclusions. The precipitation of both quartz and calcite ce-
spar authigenesis within the Palaeozoic section in wells ment have been linked to periods of rifting and magmatic
206/8-7 and 206/8-8 are consistent with temperatures deter- activity within the UK Atlantic margin (Wycherley et al.,
mined from AFTA. Fluid inclusion homogenization tem- 2003).
peratures range from 90.2 to 116.3 °C (Table 1) and
temperatures from AFTA and VR range from 90 to 5. DISCUSSION
125 °C (Table 4). This high degree of consistency between
all three techniques strongly suggests that K-feldspar 5.1. TXt data
cementation and entrapment of the inclusions occurred at
or close to the Late Palaeozoic palaeo-burial maximum. Laser ablation Ar–Ar dating of authigenic K-feldspar
In contrast, fluid inclusion homogenization tempera- cement containing characterised fluid inclusions allows inte-
tures from 111.3 to 170.5 °C for inclusions within quartz gration (Mark et al., 2006) of the fluid inclusion microther-
cement (Table 1) and values of 129.4 to 170 °C and 118.6 mometric and Ar–Ar geochronological data sets (Fig. 9).
to 159.4 °C for the first and second phase of calcite cements The integrated data shows that the timing of K-feldspar
(Table 1) are much higher than the maximum palaeotem- authigenesis in all three wells was similar, as was the salinity
peratures derived from AFTA and VR data within the Pal- of the precipitating fluids. However, fluid entrapment tem-
aeozoic section (Table 4). Similar observations have been peratures were greater in well 206/8-1a than in well 206/8-8
reported in a number of studies of Paleogene and Mesozoic and 206/8-7 (Fig. 9). The simplest and most reasonable
sandstones from the UK Atlantic margin (Parnell et al., explanation for this range of fluid entrapment temperature
1999; Green et al., 1999; Parnell et al., 2005) and can be is that the rocks in 206/8-1a had undergone enhanced burial
understood in terms of the rapid timescale of heating due and subsidence relative to those in 206/8-7 and 206/8-8,
Hydrocarbon migration through the Clair field 2525

model (Wheeler, 1996). No excess Ar was inputted into


the model. Spherical subgrains with radii of 1 lm and
500 nm, sizes determined from TEM, were tested
(Fig. 10). The Ar-diffusion model shows that the 2 lm sub-
grains are reset from 247 to 238 Ma, a 9 Ma difference, and
1 lm subgrains are reset from 247 to 222 Ma, a 25 Ma dif-
ference. The TEM results do not permit the ratio of large to
small subgrains within each laser ablation area to be deter-
mined due to petrographic scaling issues. However, the like-
lihood is that ablation areas contained varying ratios of 2
and 1 lm subgrains. Assuming a 1:1 ratio of 2 and 1 lm
subgrains we can estimate an overall resetting of Ar–Ar
ages by approximately 17 Ma, giving true growth ages of
the K-feldspar overgrowths in the Clair well of 264 Ma.
Relative to the individual Ar–Ar error for each data point,
the maximum effect of Ar-loss from subgrains on the indi-
vidual determined mineral growth ages is not significant.
Fig. 7. From the variation of Late Palaeozoic palaeotemperatures Furthermore, because the high temperature fluid pulses re-
with depth in the 206/9-2 and 206/10-1 (Figure 6, Table 4), the corded by fluid inclusions in authigenic quartz, calcite ce-
range of palaeogeothermal gradients and corresponding amounts
ment and calcite veins have not been recorded by the
of additional burial (subsequently removed by erosion) can be
AFTA, these episodes are inferred to have started rapidly
defined (Bray et al., 1992). Higher palaeo-gradients correspond to
lower amounts of removed section, and vice versa, resulting in a and to have been short-lived (Parnell et al., 2005). The
hyperbolic ellipsoid of allowed (at the 95% confidence level) values Ar-diffusion model uses the integrated AFTA-fluid inclu-
for each well, as shown by the shaded zones. sion thermal history for which periods of 0.5 Myr have been
assigned to represent the duration of the fluid pulse events
at 90 and 60 Ma (timing discussed later). If the duration
producing higher temperatures in line with the geothermal of the fluid pulses were shorter than 0.5 Myr, for example
gradient. in the order of 0.1 Myr, then the Ar–Ar ages would only
have been reset by 3 to 4 Ma, assuming a 1:1 ratio of 2–
5.2. Ar-loss and excess Ar 1 lm subgrains.
Ablation volumes contained varying numbers of fluid
Ar–Ar data show that K-feldspar authigenesis occurred inclusions. From visual two-dimensional analysis and
at 247 ± 3.3 Ma. As K-feldspar subgrains commonly leak mathematical conversion to three-dimensions (Mark
Ar at low temperatures (Lee and Parsons, 2003), the possi- et al., 2005), it has been estimated that the K-feldspar over-
bility that the Ar–Ar data does not represent K-feldspar growths contain between 0.1 and 1% (by volume) fluid
growth, but records a later thermal resetting or closure inclusions. Crustal fluids typically contain excess Ar con-
age must be considered. In order to test the Ar–Ar ages centrations that range from 0.01 to 1 ppm (Kendrick
for thermal resetting, a thermal history derived from the et al., 2002), thus it is possible to estimate the contribution
AFTA and fluid inclusion microthermometric data (well of excess Ar to measured ranges in Ar–Ar ages of the K-
206/8-7) was entered into a piecewise linear Ar-diffusion feldspar overgrowths, and approximation for a ‘worse case

Fig. 8. Reconstructed thermal histories for the 206/9-2 and 206/10-1 wells, based on the AFTA and VR data in each well. See text for details.
2526 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

Fig. 9. TXt cross-plot showing fluid inclusion homogenization temperature data versus Ar–Ar age for both aqueous- and oil-bearing fluid
inclusions in K-feldspar overgrowths. Salinity data for individual aqueous fluid inclusions is referenced to the data points.

scenario’ is required. Assuming that all ablation areas con- implies that basinal brines occupied pore spaces within
tain 1% (by volume) fluid inclusions and the fluid trapped the Clair reservoir at the time of K-feldspar cementation.
within the inclusions contains 1 ppm excess Ar, the growth With respect to K-feldspar authigenesis, the critical
age would be increased by 1.29 Ma (Mark et al., 2005). experiments describing isochemical pore fluid evolution
Hence, in summary, due to the absence of data describ- were performed by Orville (1963). Orville measured the
ing fluid episode durations (for which we modelled a worst composition (Na:K ratio) of 2 M brines (typical oilfield flu-
case scenario) and considering the possible effects of fluids ids) in equilibrium with alkali feldspars at 2 kbar PH2O. The
trapped within fluid inclusions containing 1 ppm excess experiments most relevant to diagenesis are the low-temper-
Ar, the true age of the K-feldspar overgrowths lies between ature runs down to 450 °C and although these tempera-
247 and 264 Ma. The variation in Ar–Ar age is still well tures are higher than those witnessed during diagenesis, it
within the analytical error of the individual Ar–Ar ages has been shown that extrapolation to low temperatures
(Table 2). It is important to note that the K-feldspar over- and pressures more consistent with diagenesis is possible
growth Ar–Ar age, 247 Ma, is the minimum age of feldspar (Kastner and Siever, 1979). In low temperature runs, two
precipitation and potential effects from thermal Ar-loss and feldspars (albite and K-feldspar, typical of an arkose) were
excess Ar would increase the K-feldspar overgrowth age. stable due to the alkali feldspar solvus and this is an appro-
priate comparison to the subarkose rocks found within the
5.3. Authigenesis Clair Group. For a large range of bulk feldspar composi-
tions the Na:K ratio in the fluid is buffered (fixed) by ion
Authigenic K-feldspar was precipitated within the reser- exchange of alkali metals between the brine and the feld-
voir between 247 and 264 Ma. Fluid entrapment temper- spar. The result is that fluids at lower temperatures have
atures determined from aqueous inclusions (Table 1) show higher Na:K ratios than at higher temperatures and so, as
that the precipitating fluids were of moderate temperature the fluid cools it becomes more sodic and it does this by dis-
(90 to 116 °C) and moderate salinity (3 to 7 wt.% NaCl solving albite and precipitating K-feldspar, even though the
eq.). The thermal history curve for well 206/8-7 (Fig. 5) fluid has an extremely high Na:K ratio already. If the fluid
shows that determined fluid entrapment temperatures are is heated it becomes more potassic, so it precipitates albite
approximately in line with temperatures determined from in place of K-feldspar. Albitization is a common process
AFTA. A simple temperature-burial depth conversion, that occurs in the deeper parts of sedimentary basins (Lee
assuming an elevated continental rift basin geothermal gra- and Parsons, 2003). This isochemical mechanism shows
dient of 30–90 °C km 1 (Roberts, 1985), indicates that the that there is no need to transport large amounts of Al2O3
Clair Group was buried to between 1.14 and 3.43 km and SiO2 around the system, a local Na–K replacement
(assuming an arithmetic mean value for Th). These esti- reaction can operate effectively enough to account for depo-
mates of burial depth are also compatible with the predic- sition of authigenic K-feldspar. The consistency between
tions of recent computer modelling that shows K-feldspar palaeo-fluid temperatures determined from AFTA and
authigenesis takes place at burial depths of 1 to 3 km VR data and temperatures calculated directly from fluid
(Makowitz et al., 2006). Marine waters (3.5 wt.% NaCl) inclusion microthermometry, have been attributed to burial
are insufficiently saline to account for the measured salini- and heating due to the regional geothermal gradient, not
ties of inclusions within the K-feldspar overgrowths. This flushing of hot fluids. This observation is consistent with
Hydrocarbon migration through the Clair field 2527

Fig. 10. Ar loss (DIFFARG) model showing the response of 1 and 2 lm subgrains to a reconstructed thermal history. DIFFARG is a
forward projection model, 0 Ma corresponds to 247 Ma and 247 Ma corresponds to present day. (A) Thermal history determined from AFTA
(Fig. 5) and fluid inclusion data (Table 1) against which K-feldspar subgrain Ar-retention is tested. The two spikes at 90 and 60 Ma are
discussed in text. (B) Cross-plot showing modelled Ar concentration gradient in a typical 2 lm subgrain. (C) Cross-plot showing that 2 lm
subgrains are reset from 247 Ma on the time axis to 238 Ma on the apparent age axis. The subgrains have potentially been reset by 9 Ma. (D)
Cross-plot showing modelled Ar concentration gradient in a typical 1 lm subgrain. (E) Cross-plot showing that 1 lm subgrains are reset from
247 Ma on the time axis to 222 Ma on the apparent age axis. The subgrains have potentially been reset by 25 Ma. See text for further
discussion.

an isochemical reaction controlling pore fluid composition of high fluid temperatures in the Clair field (Green et al.,
as proposed above. 1999; Parnell et al., 1999; Wycherley et al., 2003). Migration
Precipitation of authigenic quartz followed K-feldspar. of hot fluids within WOS could be related to either Late
Fluid entrapment temperatures determined from aqueous Jurassic-Early Cretaceous rifting or Tertiary magmatism
inclusions (Table 1) suggest that the temperatures of the (Nadin et al., 1997; Clift, 1999; Dean et al., 1999; Naylor,
precipitating fluids ranged from moderate to high (111 1999). Deformation that produced the deformation bands
to 170 °C) and salinities were low (0.5 to 4 wt.% NaCl within the Clair reservoir post-dates quartz cementation
eq.). Although Aplin et al. (1993) suggest it is unusual for and has been confidently related to Late-Cretaceous fault-
quartz cementation to occur within reservoirs at low tem- ing (85 to 65 Ma; Coney et al., 1993). The data shows that
peratures and fluid salinities, numerous studies since have quartz cementation occurred prior to deformation, proba-
clearly demonstrated that these processes can occur over bly during Cretaceous rifting (90 Ma). Although hot fluid
a wide range of burial depths, temperatures, pressures pulses were migrating through the region, the low solubility
and fluid salinities (Walderhaug, 1994; Worden and Morad, of silica at diagenetic temperatures and the large fluxes of
2000 and references within). Thermal histories (Fig. 8) show water correspondingly required to transport sufficient silica
that the Clair wells never reached burial temperatures as to support continued authigenesis (108 cm3 water passing
high as indicated by the fluid inclusions. Anomalously through each cm2 of rock; Worden and Morad, 2000), a lo-
high-temperature fluids may have mediated diagenesis in cal source of silica is more likely than large-scale transport
other basins of the WOS region and are the likely sources from elsewhere in the basin. The replacement of K-feldspar
2528 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

by kaolinite and/or illite yields quartz (Harris and Fullagar, from work on synthetic fluid inclusions (Palmer et al.,
1989), as does the illitization and chloritization of smectite 1999), in which trapping of single oil in the laboratory pro-
(Boles and Franks, 1979), pressure dissolution within duced oil inclusions of variable fluorescence colours. This
quartz-rich sandstones (Bjorkum, 1996; Oelkers et al., work revealed the presence of blue-fluorescing inclusions
2000) and dissolution of amorphous silica (Vagle et al., with small quantities of red-, yellow-, and orange-fluoresc-
1994; Hendry and Trewin, 1995). Any or all of these pro- ing inclusions. The commonly invoked explanation for this
cesses may have been catalysed or enhanced by invasion phenomenon is a local generation of subordinate low matu-
of the Clair Group by pulses of hot fluids. rity oil, with main charge high maturity oil represented by
Two phases of calcite authigenesis and calcite veining the dominant inclusion population (Lisk et al., 1993; Lisk
followed quartz cementation, the first ferroan-rich. Fluid and Eadington, 1994). The effects of trapping fractionation
entrapment temperatures determined from C1 aqueous may potentially vary for different mineral hosts and contin-
inclusions (Table 1) show that the precipitating fluids ran- ues to be investigated (George et al., 2001; Oxtoby, 2001).
ged from moderate to high temperatures (129 to 170 °C) O’Brien et al. (1996) used samples collected from the Timor
and low to moderate salinities (1.4 to 5.2 wt.% NaCl Sea to show that water washing of oil samples in reservoirs
eq.). Fluid entrapment temperatures determined from C2 produces molecular and gross fractionation of oils, lower-
aqueous inclusions (Table 1) show that the precipitating flu- ing specific gravity and therefore also altering the fluores-
ids ranged from moderate to high temperatures (118 to cence properties of the oil. Another potential control on
160 °C) and low to moderate salinities (1 to 5.3 wt.% the fluorescence properties of oil is the biodegradation of
NaCl eq.). Because deformation bands contain unbrecciat- hydrocarbons prior to entrapment. Bailey et al. (1973) dem-
ed C1 cement, calcite authigenesis commenced post-defor- onstrated that biodegradation decreases the aliphatic/aro-
mation. Thermal histories show that the Clair field never matic hydrocarbon ratio and increases the proportion of
reached burial temperatures of 160–170 °C (Fig. 8) and polar compounds. Any aliphatic/aromatic ratio changes
without stable oxygen isotope the source of the fluid that would affect UV fluorescence emission colour. The poten-
precipitated C1 and C2 can not be determined. Pay et al. tial fluorescence characteristics of oil inclusions can also
(2000) and Baron et al. (2008) show that calcite comprises be influenced by inorganic fluorescence from the host min-
up to 30% of the whole rock composition of Clair reservoir eral. Trace amounts of transition metals and rare earth ele-
sands and thus requires an extensive source of carbonate. ments in mineral hosts may contribute to inorganic
However, because source material for calcite is limited fluorescence which could be a further control on the fluo-
within the WOS, it is likely that either intrusion of a gabbro rescence character of oils inclusions (Wang et al., 1997; Bez-
pluton at 68 Ma (Hitchen and Ritchie, 1987) and/or ter- ouska et al., 1998).
tiary magmatic activity (60 Ma; Clift, 1999) induced ba- The presence of hydrocarbon fluid inclusions with differ-
sin-scale fluid migration. The extensive fracture network ent fluorescence colours (yellow, orange, blue, white) en-
within block 206 may have directed the fluids into the Clair trapped coevally in quartz and calcite in the Clair field is
reservoir where calcite was precipitated. A change in com- thus open to interpretation. However, the multiple fluores-
position (ferroan to non-ferroan) probably reflects exhaus- cence colours coupled with crush-leach analysis on hydro-
tion of the iron source or change in Eh of the fluids. carbon fluid inclusions from the Clair field (Scotchman
et al., 1998), which showed the presence of oils of varying
5.4. Oil inclusion fluorescence API gravity, suggests that mixing of hydrocarbons of differ-
ent maturities from different sources (Bailey et al., 1987;
Oil inclusion fluorescence colours have been used as a Rooney et al., 1998; Scotchman et al., 1998) was occurring.
qualitative guide for the thermal maturity of migrating oils The effect of oil emplacement on authigenesis rates with-
(Stasiuk and Snowdon, 1997; Bradshaw et al., 1998; O’Reil- in sandstones has been debated over previous years, partic-
ly et al., 1998) and are directly related to oil density (API ularly with reference to the retardation of quartz
gravity; Stasiuk and Snowdon, 1997; Parnell et al., 1999). precipitation during oil emplacement (Walderhaug, 1994;
Studies have shown that oil inclusions of higher API gravity Bjorkum et al., 1998; Worden et al., 1998; Marchand
and maturity fluoresce in the blue end of the visible spec- et al., 2001). The presence of oil-filled fluid inclusions within
trum, whereas those of lower API gravity and maturity primary settings in feldspar, quartz and calcite cement with-
fluoresce in the red region (McLimans, 1987; Bodnar, in samples from the Clair reservoir, suggest that cementa-
1990). Largely the relationship is based on observations tion continued after oil emplacement. However, these
made concerning the fluorescence properties of crude oils occurrences only indicate that oil was present within the
(Downare and Mullins, 1995; Ralston et al., 1996). Aro- reservoir during authigenesis and it is not possible to deter-
matic compounds are the main fluorescing components of mine if cementation rate was inhibited during oil emplace-
oil and although fluorescence in the first instance depends ment from the study data.
on chemical composition, which is primarily controlled by
maturity, several other processes can affect chemical com- 5.5. Hydrocarbon charging
position (George et al., 1997). Numerous studies have dem-
onstrated that oils fractionate during adsorption onto Hydrocarbon charging in the Clair field was synchro-
charged mineral surfaces, leading to enrichment in polar nous with K-feldspar, quartz and calcite cementation. The
compounds relative to the true parent oil composition data records only episodes of oil-charging that were accom-
(Pang et al., 1998). Support for a trapping control comes panied by pore fill cementation or veining. Current under-
Hydrocarbon migration through the Clair field 2529

standing concerning charging of the Clair field is based on Current petroleum system models for WOS are therefore
models that show multi-stage charging from the Jurassic- incomplete and a source rock was generating oil during the
Cretaceous source rocks that reside to the NW in the Late Palaeozoic. In the absence of geochemical data from
FSB (Bailey et al., 1987; Rooney et al., 1998; Scotchman the oil entrapped within the inclusions that reside in K-feld-
et al., 1998; Carr and Scotchman, 2003; Scotchman et al., spar overgrowths, it is not possible to directly correlate the
2006). It is important to note that K-feldspar overgrowths oil charge to a source rock. However, the timing of thermal
contain only blue fluorescing hydrocarbons, and this is maturation of Middle ORS lacustrine facies throughout
interpreted to represent the first recorded oil charge event Orkney and Shetland is consistent with the TXt data deter-
in the Clair field. However, as K-feldspar overgrowths con- mined from K-feldspar overgrowths in the Clair field.
tain primary hydrocarbon fluid inclusions and precipitated Hydrocarbon charging of the Clair reservoir at the time
between 247 and 264 Ma, data clearly indicate that hydro- of K-feldspar authigenesis can be linked to maturation of a
carbon generation, migration and accumulation within the Palaeozoic source rock as discussed above. However,
UK Atlantic margin must have occurred prior to deposition charging of the reservoir at times of quartz and calcite
of the Jurassic-Cretaceous source rocks in the FSB. Thus an cementation and veining, show from the fluorescence char-
oil-mature Palaeozoic source rock must have been present acters of the oil in the fluid inclusions and geochemical
within the WOS region whilst K-feldspar authigenesis was analysis of bitumens (Rooney et al., 1998; Scotchman
occurring. et al., 1998), that mixing of oil from different sources had
The only Palaeozoic source rocks identified within the occurred. Hydrocarbon charging of the Clair field during
region are the Devonian Middle ORS, East Greenland quartz and calcite cementation and veining is compatible
Upper Carboniferous and Upper Permian. The Upper with the standard model depicting hydrocarbon genera-
Permian marine black shales deposited in the Zechstein tion-migration from a Jurassic source throughout the UK
sea (256–248 Ma) would have been immature at the time Atlantic margin (Bailey et al., 1987). Thermal burial history
of oil entrapment. Lacustrine shales deposited during the modelling using stratigraphical and VR data from well
Westphalian (315–305 Ma) are a feasible source rock. 206/5-1 (Mark et al., 2005), the closest well to the Clair
AFTA data and geochemical modelling (Clift and Turner, field that contains Late Jurassic-Early Cretaceous source
1998) show sediments offshore East Greenland began to rocks, shows that they became early oil mature (0.5–0.7%
cool from 90 to 100 °C at approximately 280 Ma. However, Ro) at 120 Ma. The Late Jurassic source rocks were
considering their distance from the Clair reservoir, the time exhausted with respect to oil generation at 70 Ma,
for maturation, generation and expulsion of hydrocarbons, whereas Early Cretaceous source rocks have remained oil
and the time for migration from source to trap, the Carbon- mature up to modern day (Mark et al., 2005). As source
iferous East Greenland source does not appear to be a rocks in the FSB have remained oil-mature from 120 Ma
strong candidate for supply of the oil-charge at 264– to modern day, the degree of thermal maturity has in-
247 Ma. The Devonian ORS lacustrine deposits are the only creased with continued burial and subsidence over time
sensible option remaining and the timing of thermal matu- (Bailey et al., 1987). Increased thermal maturity and mixing
ration is a much more attractive option when compared to of hydrocarbons from multiple Devonian and Jurassic
the East Greenland source rocks. Thermal models recon- source rocks can account for variability in oil-inclusion
structed from AFTA and VR data (well 206/9-2 and 206/ fluorescence and geochemistry (Rooney et al., 1998; Scotch-
10-1; Fig. 8) show that ORS sedimentary rocks WOS were man et al., 1998).
heated to a maximum of 130 °C during burial, temperatures
high enough to push organic-rich lacustrine facies into the 6. CONCLUSIONS
oil window (Trewin, 1989). Although palaeogeographic
reconstructions of Middle ORS strata offshore Shetland is The TXt data presented here show that oil was gener-
poor, the presence of Middle ORS lacustrine facies through- ated, then migrated throughout the UK Atlantic margin
out the Orkney and Shetland Isles show that potential during the Late Palaeozoic. Thermal history data suggest
Devonian source rocks were deposited throughout the re- that Middle ORS Devonian source rocks within the region
gion, relatively close to the Clair field. Two previously pub- were oil-mature prior to the Permian and their maturation
lished thermal burial histories (Astin, 1990; Parnell et al., could account for the presence of oil inclusions within K-
1998) also show that Devonian lacustrine facies throughout feldspar overgrowths in the Clair field between 264 and
Orkney (150 km south from the Clair field; Fig. 1) entered 247 Ma. From the TXt data it is not possible to determine
the oil window between 350 and 340 Ma, remaining oil-ma- if oil was trapped and accumulating in the Clair field during
ture until between 300 and 280 Ma. Furthermore, Clift and the Late Palaeozoic, or whether the Clair field and The
Turner (1998) conducted a detailed AFT study throughout Rona Ridge actually acted as migration pathways.
the Scottish Highlands, Orkney and Shetland Isles. Mod- K-feldspar authigenesis occurred in the Late Palaeozoic
elled geothermal histories for samples from Orkney and during peak burial of the Devonian-Carboniferous red-bed
Shetland showed cooling from temperatures in excess of sediments. Calculated temperatures of precipitating fluids
110 °C at 350 Ma, to approximately 80 °C by 200 Ma. are commensurate with temperatures predicted by the re-
The data supports the hypothesis put forward in this man- gions geothermal gradient and are not a product of flushing
uscript that oil was generated from the Devonian source of hot fluids throughout the margin. Quartz and calcite
rocks north of the British Isles and charged the Clair reser- authigenesis occurred during the Mesozoic and Cenozoic,
voir during the Early Permian. respectively. Authigenesis was related to transient heating
2530 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

by pulses of hot fluid. Thermal history data show that inter- Bodnar R. J. (1990) Petroleum migration in the Miocene Monterey
mittent flow of hot fluid throughout the UK Atlantic mar- Formation, California, USA: constraints from fluid-inclusion
gin was responsible for complex diagenetic sequences in studies. Mineral. Mag. 54, 295–304.
Tertiary reservoirs, and occurred during both the Cenozoic Bodnar B. J. (1993) Revised equation and table for determining the
freezing point depression of H2O–NaCl solutions. Geochim.
and Mesozoic. These events were presumably related to epi-
Cosmochim. Acta 57, 683–684.
sodic rifting of the UK Atlantic margin. Boles J. R. and Franks S. G. (1979) Clay diagenesis in the Wilcox
In summary, this study shows that by integrating TXt Sandstone of SW Texas: implications of smectite diagenesis on
data into a regional model incorporating thermal con- sandstone cementation. J. Sediment. Petrol. 49, 55–70.
straints from AFTA and VR data, it is possible to develop Bradshaw J., Sayers J., Bradshaw M., Kneale R., Ford C. and
complex oil-show models. The novel determination of TXt Spencer, L. (1998) Palaeogeography and its impact on the
data through analysis of authigenic K-feldspar represents a Petroleum systems of the NW Shelf, Australia. In Proc. Expl.
powerful tool for reconstructing palaeo-fluid flow within Soc. Australia Symp. (eds. P. G. Purcell and R. R. Purcell).
sedimentary basins and should find many applications in Perth, pp. 99–121.
regional exploration, especially as hydrocarbon resources Bray R. J., Green P. F. and Duddy I. R. (1992). Thermal history
reconstruction using apatite fission track analysis and virinite
continue to dwindle during the 21st Century.
reflectance: a case study from the UK East Midlands and
ACKNOWLEDGMENTS Southern North Sea. In Exploration Britain: Geological Insights
for the Next Decade (ed. R. P. F. Hardman). Geological Society
This work was supported by the Natural Environment Re- Special Publications No.67, pp. 3–25.
search Council (3220/GL021/ RGA0782). We thank John Still Burnham A. K. and Sweeney J. J. (1989) A chemical kinetic model
and James Schwanethal for technical assistance. We are also grate- of vitrinite reflectance maturation. Geochim. Cosmochim. Acta
ful to A. Craven for access to the FIB-TEM facilities in Glasgow 53, 2649–2657.
University and C. How for help with microscope operation. Asso- Carr A. D. and Scotchman I. C. (2003) Thermal history modelling in
ciate Editor E.H. Oelkers is thanked for comments leading to the southern Faroe-Shetland Basin. Petrol. Geosci. 9, 333–345.
improvement of this manuscript. Two reviewers, I. Scotchman Clift P. D. and Turner J. (1998) Paleogene igneous underplating
and H. Shaw are also thanked for detailed reviews, comments and subsidence anomalies in the Rockall–Faeroe–Shetland
and constructive suggestions. area. Mar. Petrol. Geol. 15, 223–243.
Clift P. D. (1999). The thermal impact of Paleocene magmatic
underplating in the Faeroe-Shetland-Rockall region. In Petro-
REFERENCES leum Geology of Northwest Europe: Proceedings of the 5th
Conference (eds. A. J. Fleet and S. A. R. Boldy). Geological
Allen P. A. and Mange-Rajetzky M. A. (1992) Devonian-Carbon- Society, London, pp. 585–593.
iferous sedimentary evolution of the Clair area, offshore north- Coney D., Fyfe T. B., Retail, P. and Smith, P. J. (1993) Clair
western UK; impact of changing provenance. Mar. Petrol. appraisal: the benefits of a co-operative approach. In Petro-
Geol. 9, 29–52. leum Geology of Northwest Europe: Proceedings of the 4th
Aplin A. C., Warren E.A., Grant S. M. and Robinson, A. G. (1993) Conference (ed. J. R. Parker). Geological Society, London,
Mechanisms of quartz cementation in North Sea reservoir pp. 1409–1420.
sandstones: constraints from fluid compositions. In Diagenesis Cook A. C. (1982). The Origin and Petrology of Organic Matter in
and basin development, vol. 36 (eds. A. D. Horbury and A. Coals, Oil Shales and Petroleum Source Rocks. University of
Robinson) American Association of Petroleum Geologists Wollongong, Wollongong, New South Wales.
Studies in Geology, pp. 7–22. Dean K., McLachlan K. and Chambers, A. (1999). Rifting and the
Astin T. R. (1990) The Devonian lacustrine sediments of Orkney, development of the Faeroe-Shetland Basin. In Petroleum
Scotland; implications for climate cyclicity, basin structure and Geology of Northwest Europe: Proceedings of the 5th Conference
maturation history. J. Geol. Soc. Lond. 147, 141–151. (eds. A. J. Fleet and S. A. R. Boldy). Geological Society,
Bailey G. W., Brown D. S. and Karickhoff S. W. (1973) London, pp. 533–544.
Competitive hydration of quinazoline at the montmorillio- Downare T. D. and Mullins O. C. (1995) Visible and near infrared
nite–water interface. Science 182, 819–821. fluorescence of crude oils. Appl. Spectrosc. 49, 470–474.
Bailey N. J. L., Walko P. and Sauer M. J. (1987) Geochemistry and Duddy, I. R., Green, P. F., Bray, R. J. and Hegarty, K. A. (1994)
source rock potential of the West of Shetlands. In Petroleum Recognition of the thermal effects of fluid flow in sedimentary
Geology of North West Europe (eds. J. Brooks and K. Glennie). Basins. In Geofluids: Origin, Migration and Evolution of Fluids
pp. 711–721. in Sedimentary Basins (ed. J. Parnell). Geological Society of
Baron M., Parnell J., Mark D. F., Przyjalgowski M. and Feely M. London, Special Publication 78, pp. 325–345.
(2008) Evolution of hydrocarbon migration style in a fractured Duddy, I. R., Green, P. F., Hegarty, K. A., Bray, R. J. and
reservoir deduced from fluid inclusion data, Clair field, West of O’Brien, G. W. (1998). Dating and duration of hot fluid flow
Shetland. Mar. Petrol. Geol. 25, 153–172. events determined using AFTAÒ and vitrinite reflectance-based
Bezouska J. R., Wang J. and Mullins O. C. (1998) Origins of thermal history reconstruction. In Dating and duration of fluid
limestone fluorescence. Appl. Spectrosc. 52, 1606–1613. flow and fluid-rock interaction (ed. J. Parnell). Geological
Bjorkum P. A. (1996) How important is pressure in causing Society, London, Special Publication 144, pp. 41–51.
dissolution of quartz in sandstones? J. Sediment. Res. 66, 147– Duncan W. I. and Buxton N. W. K. (1995) New evidence for
154. evaporitic Middle Devonian lacustrine sediments with hydro-
Bjorkum P. A., Oelkers E. H., Nadeau P. H., Walderhaug O. and carbon source potential on the East Shetland Platform, North
Murphy W. M. (1998) Porosity prediction in quartzose Sea. J. Geol. Soc. Lond. 152, 251–258.
sandstones as a function of time, temperature, depth, stylolite Galbraith R. F. and laslett G. M. (1993) Statistical models for
frequency and hydrocarbon saturation. Am. Assoc. Petrol. mixed fission track ages. Nucl. Tracks Radiat. Measure. 21,
Geol. Bull. 82, 637–648. 459–470.
Hydrocarbon migration through the Clair field 2531

Gallagher K. (1995) Evolving temperature histories from apatite Harris W. B. and Fullagar P. D. (1989) Comparison of Rb–Sr and
fission-track data. Earth Planet. Sci. Lett. 136, 421–435. K–Ar dates of middle Eocene bentonite and glauconite, South-
George S. C., Greenwood P. F., Logan G. A., Quezada R. A., Pang eastern Atlantic Coastal plain. Bull. Geol. Soc. Am. 101, 573–
L. S. K. and Lisk M. (1997) Comparison of palaeo oil changes 577.
with currently reservoired hydrocarbons using molecular and Hanor J. S. (1980) Dissolved methane in sedimentary brines:
isotopic analyses of oil-bearing fluid inclusions: Jabiru oil field, potential effect on the PVT properties of fluid inclusions. Econ.
Timor Sea. Austral. Petrol. Produc. Explor. Assoc. J. 37(1), Geol. 75, 603–609.
490–504. Hendry J. P. and Trewin N. H. (1995) Authigenic quartz
George S. C., Ruble A., Dutkiewicz A. and Eadington P. J. (2001) microfabrics in Cretaceous turbidites—evidence from silica
Assessing the maturity of oil trapped in fluid inclusions using transformation processes in sandstones. J. Sediment. Res. 65,
molecular geochemistry data and visual determined fluores- 380–392.
cence colours. Appl. Geochem. 16, 451–473. Hitchen K. and Ritchie J. D. (1987) Geological review of the West
Goldstein R. H. (2001) Fluid inclusions in sedimentary and Shetland area. In Petroleum Geology of North West Europe (eds.
diagenetic systems. Lithos 55, 159–193. J. Brooks and K. Glennie). Graham and Trotman, London, pp.
Green P. F. (1981) A practical method of estimating standard error 737–749.
of age in the fission track dating method. Nucl. Tracks Radiat. Hillier S. and Marshall J. E. A. (1992) Organic maturation, thermal
Measure. 5(3), 317–322. history and hydrocarbon generation in the Orcadian Basin,
Green P. F. (1986) The thermo-tectonic evolution of Northern Scotland. J. Geol. Soc. Lond. 149, 491–502.
England, evidence from fission track analysis. Geol. Mag. 123, Holford S. P., Green P. F. and Turner J. P. (2005) Palaeothermal
493–506. and compaction studies in the Mochras borehole (NW Wales)
Green P. F., Duddy I. R., Gleadow A. J. W. and Lovering J. F. reveal Early Cretaceous and Neogene exhumation and argue
(1989a) Apatite Fission Track Analysis as a palaeotemperature against regional Palaeogene uplift in Southern Irish Sea. J.
indicator for hydrocarbon exploration. In Thermal History of Geol. Soc. Lond. 162, 829–840.
Sedimentary Basins: Methods and Case Histories (eds. N. Holmes A. J., Griffith C. E. and Scotchman I. C. (1999) The
Naesaer and T. H. McCulloch). Springer-Verlag, New York, Jurassic petroleum system of the West of Britain Atlantic
pp. 181–195. margin—an integration of tectonics, geochemistry and basin
Green P. F., Duddy I. R., Laslett G. M., Hegarty K. A., Gleadow modelling. In Petroleum Geology of Northwest Europe: Pro-
A. J. W. and Lovering J. F. (1989b) Thermal annealing of ceedings of the 5th Conference (eds. A. J. Fleet and S. A. R.
fission tracks in apatite 4. Quantitative modelling techniques Boldy) Geological Society, London, pp. 1351–1365.
and extension to geological timescales. Chem. Geol. (Isotope Hurford A. J. and Green P. F. (1982) A users’ guide to fission track
Geosci. Sect.) 79, 155–182. dating calibration. Earth. Planet. Sci. Lett. 59, 343–354.
Green P. F., Duddy I. R. and Bray R. J. (1995) Low-temperature Iliffe J. E., Robertson A. G., Ward G. H. F., Wynn C., Pead S. D.
thermal history of Eastern Ireland—effects of fluid flow— M. and Cameron N. (1999) The importance of fluid pressure
discussion. Mar. Petrol. Geol. 12(7), 773–776. and migration to the hydrocarbon prospectivity of the faroe-
Green, P. F., Duddy, I. R., Hegarty, K. A. and Bray, R. J. Shetland White Zone. In Petroleum Geology of Northwest
(1999) Early Tertiary heat flow along the UK Atlantic Europe: Proceedings of the 5th Conference (eds. A. J. Fleet and
margin and adjacent areas. In Petroleum Geology of North S. A. R. Boldy). Geological Society, London, pp. 645–659.
West Europe, Proceedings of the 5th Conference (eds. A. J. Irwin H. and Meyer T. (1990) Lacustrine organic facies—a
Fleet and S. A. R. Boldy). Geological Society, London, pp. biomarker study using multivariate statistical analysis. Org.
348–357. Geochem. 16, 197–210.
Green P. F., Duddy I. R., Hegarty K. A., Bray R. J., Sevastopulo Japsen P., Green P. F. and Chalmers J. A. (2005) Separation of
G., Clayton G. and Johnston D. (2001) The post-Carboniferous Palaeogene and Neogene uplift on Nuussuaq, West Greenland.
evolution of Ireland: evidence from thermal history reconstruc- J. Geol. Soc. Lond. 162, 299–314.
tion. Proc. Geol. Assoc. 112, 192–194. Kastner M. and Siever R. (1979) Low temperature feldspars in
Green P. F., Duddy I. R. and Hegarty K. A. (2002) Quantifying sedimentary rocks. Am. J. Sci. 279, 435–479.
exhumation from apatite fission-track analysis and vitrinite Kelley S. P., Arnaud N. O. and Turner S. P. (1994) High spatial
reflectance data: precision, accuracy and latest results from the resolution 40Ar/39Ar investigation using an ultra-violet laser
Atlantic margin of NW Europe. In Exhumation of the North probe extraction technique. Geochim. Cosmochim. Acta 58(16),
Atlantic Margin: Timing Mechanisms and Implications for 3519–3525.
Petroleum Exploration, vol. 196 (eds. A. G. Doré, J. Cartwright, Kendrick M. A., Burgess R., Pattrick R. A. D. and Turner G.
M. S. Stoker, J. P. Turner and N. White). Geological Society (2002) Hydrothermal fluid origins in a fluorite-rich MVT
Special Publication, pp. 331–354. district: combined noble gas and halogen analysis of fluid
Green P. F., Crowhurst P. V. and Duddy I. R. (2004) Integration inclusions from the South Pennine ore field, UK. Econ. Geol.
of AFTA and (U-Th)/He thermochronology to enhance the 97, 435–451.
resolution and precision of thermal history reconstruction in Lamers E. and Carmichael, S. M. M. (1999) The Paleocene
the Anglesea-1 well, Otway Basin, SE Australia. In Eastern deepwater sandstone play West of Shetland. In Petroleum
Australian Basins Symposium II (eds. P. J. Boult, D. R. Johns Geology of Northwest Europe: Proceedings of the 5th Conference
and S. C. Lang). Petroleum Exploration Society of Australia, (eds. A. J. Fleet and S. A. R. Boldy). Geological Society,
Special Publication, pp. 117–131. London, pp. 645–659.
Hagen E., Kelley S. P., Dypvik H., Nilsen O. and Kjølhamar B. Lee M. R. and Parsons I. (2003) Microtextures of authigenic Or-
(2001) Direct dating of authigenic K feldspar overgrowths from rich feldspar in the Upper Jurassic Humber Group, UK North
the Kilombero Rift of Tanzania. J. Geol. Soc. Lond. 158, 801– Sea. Sedimentology 50, 597–608.
807. Lisk M., Hamilton J., Eadington P. and Kotaka T. (1993).
Harland W. B., Armstrong R. L., Cox A. V., Craig L. E., Smith A. Hydrocarbons and pore water migration history in relation to
G. and Smith D. G. (1989) A Geological Time Scale 1989. diagenesis in the Toro and Iagifu Sandstones, S. E. Gobe-2. In
Cambridge University Press, Cambridge, 263. Petroleum exploration and development in Papua New Guinea.
2532 D.F. Mark et al. / Geochimica et Cosmochimica Acta 72 (2008) 2510–2533

Proc. 2nd PNG Petroleum Convention (eds. G. J. Carman, Z. Parnell J., Carey P. and Monson B. (1998) Timing and temperature
Carman). Port Moresby, pp. 477–488. of decollement on hydrocarbon source rock beds in cyclic
Lisk M. and Eadington P.J. (1994) Oil migration in the Cartier lacustrine successions. Palaeogeogr. Palaeoclimatol. Palaeoecol.
Trough, Vulcan sub-basin. In The Sedimentary Basins of 140, 121–134.
Western Australia. Proc. West Australian Basins Symp. (eds. Parnell J., Carey P. F., Green P. and Duncan W. (1999)
P. G. Purcell, R. R. Purcell). Perth, pp. 301–312. Hydrocarbon migration history, West of Shetland; integrated
McLimans R. K. (1987) Mixing zone and deep burial diagenesis— fluid inclusion and fission track studies. In Petroleum geology of
relevance to reservoir porosity and oil migration, Middle northwest Europe: Proceedings of the fifth conference (eds. A. J.
Jurassic Great Oolite Formation, Wealden Basin, England. Fleet and S. A. R. Boldy). Geological Society, London, pp.
Am. Assoc. Petrol. Geol. 71, 591–598. 613–625.
Makowitz A., Lander R. H. and Millikin K. L. (2006) Diagenetic Parnell J., Green P. F., Watt G. and Middleton D. (2005) Thermal
modelling to assess the relative timing of quartz cementation history and oil charge on the UK Atlantic Margin. Petrol.
and brittle grain process during compaction. Am. Assoc. Petrol. Geosci. 11, 99–112.
Geol. Bull. 90(6), 873–885. Pay M. D., Astin T. R. and Parker A. (2000) Clay mineral
Marchand A. M. E., Haszeldine R. S., Smalley P. C., Macaulay C. distribution in the Devonian-Carboniferous sandstones of the
I. and Fallick A. E. (2001) Evidence for reduced quartz Clair field, West of Shetland, and its significance for reservoir
cementation rates in oil-filled sandstones. Geology 29, 915–918. quality. Clay Miner. 35, 151–162.
Mark D. F., Parnell J., Kelley S. P., Lee M., Sherlock S. C. and Peters K. E., Moldowan J. M., Driscole A. R. and Demaison G. J.
Carr A. (2005) Dating of multistage fluid flow in sandstone. (1989) Origin of Beatrice oil by co-sourcing from Devonian and
Science 309, 2048–2051. Middle Jurassic source rocks, Inner Moray Firth, United
Mark D. F., Parnell J., Kelley S. P. and Sherlock S. C. (2006) Kingdom. Am. Assoc. Petrol. Geol. Bull. 73(4), 454–471.
Temperature–composition–time (TXt) data from authigenic K- Ralston C. Y., Wu X. and Mullins O. C. (1996) Quantum yields of
feldspar: an integrated methodology for dating fluid flow crude oils. Appl. Spectrosc. 50, 1563–1568.
events. J. Geochem. Explor. 89, 259–262. Renne P. R., Swisher C. C., Deino A. L., Karner D. B., Owens T.
Mark D. F., Parnell J., Kelley S. P. and Sherlock S. C. (2007) L. and DePaolo D. J. (1998) Intercalibration of standards,
Resolution of regional fluid flow related to successive orogenic absolute ages and uncertainties in Ar/Ar dating. Chem. Geol.
events on the Laurentian margin. Geology 35(6), 547–550. 145, 117–152.
Munz I. A. (2001) Petroleum inclusions in sedimentary basins: Roberts P. (1985) Histoire geothermique et diagenesis organique.
systematics, analytical methods and applications. Lithos 55, ELF Aquitaine, France. English translation (1988) D. Reidel
195–212. Dordrecht, Holland.
Nadin P. A., Kusznir N. J. and Cheadle M. J. (1997) Early Tertiary Rooney M. A., Vuletich A. K. and Griffith C. E. (1998)
plume uplift of the North Sea and Faeroe-Shetland Basins. Compound-specific isotope as a tool for characterizing mixed
Earth Planet. Sci. Lett. 148, 109–127. oils: an example from the West of Shetlands area. Org.
Naylor P. H., Bell, B. R., Jolley D. W. Durnall P. and Fredsted R. Geochem. 29(1-3), 241–254.
(1999) Palaeogene magmatism in the Faeror-Shetland Basin: Russell C. A., Snape C. E., Meredith W., Love G. D., Clarke E. and
influences on uplift history and sedimentation. In Petroleum Moffatt B. (2004) The potential of bound biomarker profiles
Geology of Northwest Europe: Proceedings of the 5th Conference released via catalytic hydropyrolysis to reconstruct basin charg-
(eds. A. J. Fleet and S. A. R. Boldy). Geological Society, ing history for oils. Org. Geochem. 35(11-12), 1441–1459.
London, pp. 545–558. Scotchman I. C., Griffith C. E., Holmes A. J. and Jones D. M.
O’Brien G. W., Lisk M., Duddy I., Eadington P. J., Cadman S. and (1998) The Jurassic petroleum system north and west of Britain:
Fellows M. (1996) Late Tertiary fluid migration in the Timor a geochemical oil-source correlation study. Org. Geochem. 29(1-
Sea: a key control on thermal and diagenetic histories? Austral. 3), 671–700.
Petrol. Produc. Explor. Assoc. J. 36(1), 399–427. Scotchman I. C., Carr A. D. and Parnell J. (2006) Hydrocarbon
Oelkers E. H., Bjørkum P. A., Walderhaug O., Nadeau P. H. and generation modelling in a multiple rifted and volcanic basin: a
Murphy W. M. (2000) Making diagenesis obey thermodynam- case study in the Foinaven Sub-basin, Faroe-Shetland Basin,
ics and kinetics: the case of quartz cementation in sandstones UK Atlantic margin. Scot. J. Geol. 42(1), 1–19.
from offshore mid-Norway. Appl. Geochem. 15, 295–309. Shepherd T. J., Rankin A. H. and Alderton D. H. M. (1985) A
O’Reilly C., Shannon P. M. and Feely M. (1998) A fluid inclusion Practical Guide to Fluid Inclusions. Chapman and Hall.
study of cement and vein minerals from the Celtic Sea, offshore Stasiuk L. D. and Snowdon L. R. (1997) Fluorescence micro-
Ireland. Mar. Petrol. Geol. 15, 519–533. spectrometry of synthetic and natural hydrocarbon fluid
Orville P. M. (1963) Alkali ion exchange between vapour and inclusions: crude oil chemistry, density and application to
feldspar phases. Am. J. Sci. 261, 201–237. petroleum migration. Appl. Geochem. 12, 229–241.
Oxtoby N. H. (2001) Comments on: Assessing the maturity of oil Tissot B. P. and Welte D. H. (1984) Petroleum Formation and
trapped in fluid inclusions using molecular geochemistry data Occurrence. Springer-Verlag.
and visually-determined fluorescence colours. Appl. Geochem. Trewin N. H. (1989) The petroleum potential of the Old Red
17, 1371–1374. Sandstone of northern Scotland. Scot. J. Geol. 25(2), 201–225.
Palmer N., Eadington P. J. and George S. C. (1999) Effect of Vagle G. B., Hurst A. and Dypvik H. (1994) Origin of quartz
aromatic fractions on fluorescence colours of oil inclusions in cements in sandstones from the Jurassic of the Inner Moray
samples from Australian and Papuan oil fields. CSIRO Petro- Firth (UK). Sedimentology 41(2), 363–377.
leum Unrestricted Report No. 99-002. Walderhaug O. (1994) Precipitation rates for quartz cement in
Pang L. S. K., George S. C. and Quezada R. A. (1998) A study of sandstones determined by fluid inclusion microthermometry
the gross composition of oil-bearing fluid inclusions using high and temperature-history modelling. J. Sediment. Res. A64, 324–
performance liquid chromatography. Org. Geochem. 29, 1149– 333.
1161. Wang J., Wu X. and Mullins O. C. (1997) Fluorescence of
Parnell J. (1985) Hydrocarbon source rocks, reservoir rocks and limestones and limestone components. Appl. Spectrosc. 51,
migration in the Orcadian Basin. Scot. J. Geol. 21, 321–335. 1890–1895.
Hydrocarbon migration through the Clair field 2533

Wheeler J. (1996) DIFFARG: a program for stimulating argon Wycherley H. L., Parnell J., Watt G. R., Chen H. and Boyce A. J.
diffusion profiles in minerals. Comput. Geosci. 22(8), 919–929. (2003) Indicators of hot fluid migration in sedimentary basins:
Worden R. H., Oxtoby N. H. and Smalley P. C. (1998) Can oil evidence from the UK Atlantic Margin. Petrol. Geosci. 9, 357–
emplacement prevent quartz cementation in sandstones? Petrol. 374.
Geosci. 4, 129–137. Ziagos J. P. and Blackwell D. D. (1986) A model for
Worden R. H. and Rushton J. C. (1992) Diagenetic K-feldspar the transient temperature effects of horizontal fluid
textures: a TEM study and model for diagenetic feldspar flow in geothermal systems. J. Volcanol. Geotherm. Res. 27,
growth. J. Sediment. Petrol. 62(5), 779–789. 371–397.
Worden R. H. and Morad S. (2000) Quartz Cementation in
Sandstones. Blackwell Publishing. Associate editor: Eric H. Oelkers

You might also like