You are on page 1of 34

Marine and Petroleum Geology 21 (2004) 457484 www.elsevier.

com/locate/marpetgeo

A sedimentological approach to rening reservoir architecture in a mature hydrocarbon province: the Brent Province, UK North Sea
Gary J. Hampson*, Peter J. Sixsmith, Howard D. Johnson
Department of Earth Science and Engineering, Imperial College, Prince Consort Road, London SW7 2BP, UK Received 5 April 2003; received in revised form 26 July 2003; accepted 29 July 2003

Abstract Improved reservoir characterisation in the mature Brent Province of the North Sea, aimed at maximising both in-eld and near-eld hydrocarbon potential, requires a clearer understanding of sub-seismic stratigraphy and facies distributions. In this context, we present a regional, high-resolution sequence stratigraphic framework for the Brent Group, UK North Sea based on extensive sedimentological re-interpretation of core and wireline-log data, combined with palynostratigraphy and published literature. This framework is used to place individual reservoirs in an appropriate regional context, thus resulting in the identication of subtle sedimentological and tectono-stratigraphic features of reservoir architecture that have been previously overlooked. We emphasise the following insights gained from our regional, high-resolution sequence stratigraphic synthesis: (1) improved denition of temporal and spatial trends in deposition both within and between individual reservoirs, (2) development of regionally consistent, predictive sedimentological models for two enigmatic reservoir intervals (the Broom and Tarbert Formations), and (3) recognition of subtle local tectono-stratigraphic controls on reservoir architecture, and their links to the regional, Middle Jurassic structural evolution of the northern North Sea. We discuss the potential applications of these insights to the identication of additional exploration potential and to improved ultimate recovery. q 2003 Elsevier Ltd. All rights reserved.
Keywords: Brent Group; Broom formation; Tarbet formation; Sequence stratigraphy; Reservoir architecture; Ninian HuttonDunlin system

1. Introduction The Middle Jurassic Brent Group has been the most productive reservoir interval in the North Sea during the last 25 years, hosting over 50 elds that contained a total of approximately 15 billion barrels of recoverable oil (Fig. 1; Lee & Hwang, 1993). Most major Brent Group elds are now in production decline, thus recovery and value in these reservoirs needs to be maximised. One tool for this is the application of high-resolution sequence stratigraphic concepts for predicting detailed, 3D reservoir architecture. This approach relies largely on core, well-log and production history data due to low seismic resolution (typically ca. 20 m). To date, sequence stratigraphic studies of the Brent Group have erected either regional frameworks of low temporal resolution (1 10 Ma; e.g. Fjellganger, Olsen, & Rubino, 1996; Graue et al., 1987; Helland-Hansen, Badley,
* Corresponding author. Tel.: 44-20-759-46475; fax: 44-20-75946464. E-mail address: g.j.hampson@ic.ac.uk (G.J. Hampson). 0264-8172/$ - see front matter q 2003 Elsevier Ltd. All rights reserved. doi:10.1016/S0264-8172(03)00094-1

Lmo, & Steel, 1992; Mitchener, Lawrence, Partington, Bowman, & Gluyas, 1992), or high-resolution (, 1 Ma) frameworks that are specic to an individual reservoir or small group of reservoirs (Table 1; e.g. Flint, Knight, & Tilbrook, 1998; Jennette & Riley, 1996; Reynolds, 1995). Although these studies have undoubtedly been valuable exploration and production tools, there remain signicant discrepancies in the spatial and temporal scales considered by the various low-resolution, regional studies and highresolution, reservoir studies. As a result, regional trends in deposition have rarely been understood at a sufciently high temporal resolution to enable reliable predictions of reservoir-scale stratigraphic architecture within, and between, individual elds. In this article, we present a new regional, high-resolution (, 1 Ma) sequence stratigraphic framework that links regional sedimentological, tectonic and palaeogeographic evolution to reservoir-scale facies architecture. Our aim is to highlight the added value of this integrated regional-scale to reservoir-scale approach in identifying near-eld exploration potential and additional recovery opportunities in producing

458

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

Fig. 1. Location map of the Brent Province showing major Brent Group reservoirs and the study area (Figs. 2 and 4).

reservoirs. Based on this aim, we emphasise the following points using our regional, high-resolution sequence stratigraphic synthesis: (1) improved denition of temporal and spatial trends in deposition within and between reservoirs,

(2) development of regionally consistent, predictive sedimentological models for two hitherto-enigmatic reservoir intervals (the Broom and Tarbert Formations), and (3) recognition of subtle tectono-stratigraphic controls on

Table 1 Summary of published sedimentological and high-resolution (reservoir zone-scale) sequence stratigraphic work on individual elds in the study area Field Ninian Strathspey Brent Previous work High-resolution biostratigraphy and reservoir zonation scheme Detailed facies model, high-resolution sequence stratigraphy and reservoir zonation scheme Detailed facies model, high-resolution genetic stratigraphy and reservoir zonation scheme, production data tied to reservoir zonation Detailed facies model, high-resolution sequence stratigraphy Detailed facies model, high-resolution genetic stratigraphy and reservoir zonation scheme Detailed facies model, high-resolution sequence stratigraphy and reservoir zonation scheme Detailed facies model Detailed facies model, high-resolution sequence stratigraphy Detailed facies model, high-resolution sequence stratigraphy Detailed facies model Authors Sawyer and Keegan (1996) and Underhill et al. (1997) Morris et al. (2003) Bryant and Livera (1991) and Livera (1989)

Northwest Hutton Cormorant Thistle Murchison Tern Eider Don and Dee

Flint et al. (1998) and Richards and Brown (1986) Budding and Inglin (1981), Jennette and Riley (1996) and Livera and Caline (1990) Reynolds (1995) Brown and Richards (1989) Jennette and Riley (1996) Jennette and Riley (1996) Brown and Richards (1989)

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

459

reservoir architecture and their links to the regional, Middle Jurassic structural evolution of the northern North Sea. 2. Geological setting 2.1. Structural evolution of the northern North Sea The northern North Sea has undergone a complex structural evolution that encompasses three discrete episodes of rifting separated by periods of tectonic quiescence (Fig. 2A; Lee & Hwang, 1993; Rattey & Heyward, 1993). The rst rifting episode occurred as part of the Caledonian Orogeny during the Devonian, and resulted in the generation of NE SW trending structures that are exemplied by the Tern Eider Horst feature (Fig. 2A and B; Johnson & Dingwall, 1981; Lee & Hwang, 1993). The second rifting episode occurred during the Triassic (Fig. 2A and B; Frseth, 1996; Lee & Hwang, 1993; Roberts, Kusznir, Walker, & Dorn-Lopez, 1995), and has previously been difcult to interpret because of overprinting by late Jurassic structures. However, recent interpretation of high-quality 3D seismic volumes suggests that this rifting episode produced a series of half-grabens bounded by NNE SSW trending, west-facing

normal faults (John Underhill, pers. comm.). Although the spatial development of the Triassic basin is broadly coincident with later Jurassic depocentres in the northern North Sea (Lee & Hwang, 1993; Rattey & Heyward, 1993; Roberts et al., 1995), there are some signicant differences in fault orientation and spacing between the two rifting episodes (Fig. 2B and C; Frseth, 1996; John Underhill, pers. comm.). The Brent Group was deposited during the phase of postrift subsidence that followed Triassic rifting (Fig. 2A). Brent Group deposition is coincident with thermal doming above a transient mantle plume in the central North Sea and subsequent deation of this dome as the thermal anomaly produced by the mantle plume dissipated (Fig. 2A; Underhill & Partington, 1993). It has been suggested that deation of the thermal dome introduced the regional tensional regime that caused development of the trilete North Sea rift basin in the late Jurassic (Davies, Turner, & Underhill, 2001). In the northern North Sea, late Jurassic rifting produced a series of half-grabens bounded by N S trending normal faults (Fig. 2A and C; Davies et al., 2001). Detailed interpretation of these fault systems suggest that they were initiated during deposition of the youngest Brent Group strata, but they grew and linked predominantly during post-Brent times (Fig. 2B;

Fig. 2. (A) Chronostratigraphic column highlighting the main phases of structural evolution in relation to Brent Group deposition (see text for details). (B) Map of the UK Brent Province showing key structures that controlled Brent Group deposition: the TernEider Horst, a Caledonian basement feature (after Lee & Hwang, 1993); inactive Triassic rift faults (after Lee & Hwang, 1993; Roberts et al., 1995); and active Middle Jurassic extensional faults (after McLeod et al., 2000; Roberts et al., 1995). During Brent Group deposition, the NinianHuttonDunlin fault system had already formed a well-developed, hard-linked fault (Roberts et al., 1995; Underhill, pers. comm.), whereas the North Alwyn BrentStatfjord fault system initiated as a series of small, unlinked fault segments during latest Brent Group deposition (McLeod et al., 2000). (C) Map of the UK Brent Province showing late Jurassic rift faults, which formed the structural traps for most Brent Group reservoirs, and younger structures. Fig. 2B and C are shown at the same scale. Note that the Late Jurassic structure map (Fig. 2C) gives a poor impression of the structural controls on Brent Group deposition (Fig. 2B).

460

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

Davies, Dawers, McLeod, & Underhill, 2000; Dawers & Underhill, 2000; McLeod, Dawers, & Underhill, 2000). A good understanding of the complex structural evolution outlined above (Fig. 2) is critical in determining the structural setting of the northern North Sea during deposition of the Brent Group in the Middle Jurassic. Most importantly, many of the late Jurassic extensional faults that dene the structural traps of all current Brent Group reservoirs (Fig. 2C) were not present during Brent Group deposition (Dawers & Underhill, 2000; McLeod et al., 2000). Instead, Brent Group deposition was predominantly inuenced by passive differential subsidence across older, basement structures and active extension across a small number of pre-existing fault systems (Fig. 2B). The latter may have formed in response to thermal doming and relaxation during the Lower and Middle Jurassic (Underhill & Partington, 1993). Latest Brent Group deposition may have been inuenced locally by initiation of the fault arrays that grew and linked into major rift structures during the late Jurassic, particularly along the southern part of the North Alwyn Brent-Statfjord fault system (Fig. 2B; Davies et al., 2000; McLeod et al., 2000). Key structures controlling Brent Group deposition include: (1) the Tern Eider Horst; a basement-involved Caledonian trend whose northwestern boundary is marked by Triassic rift faults and whose southeastern boundary coincides with the Cormorant Field (Fig. 2B; Lee & Hwang, 1993; Roberts et al., 1995), and (2) the Ninian Hutton Dunlin fault system; an extensional fault system that was initiated during the lower Jurassic and which was active during the Middle Jurassic (Fig. 2B; Roberts et al., 1995; John Underhill, pers. comm.). 2.2. Brent Group lithostratigraphy

and Tarbert Formations (Fig. 3; Deegan & Scull, 1977). The Broom Formation and its Norwegian equivalent, the Oseberg Formation, comprise coarse-grained shallow marine deposits that are now widely considered to be genetically unrelated to the overlying Formations (Fig. 3; Fjellganger et al., 1996; Helland-Hansen et al., 1992; Mitchener et al., 1992). The upper four Formations record a major regressivetransgressive episode in which the Rannoch, Etive and Lower Ness Formations represent overall regression of a wave-dominated deltaic coastline and the Upper Ness and Tarbert Formations record subsequent transgression (e.g. Brown, Richards, & Thomson, 1987). More detailed interpretations of the regional Brent Group succession have recognised a variable number of low-frequency (110 Ma) stratigraphic cycles that are superimposed on the major regressivetransgressive episode (e.g. Fjellganger et al., 1996; Graue et al., 1987; Helland-Hansen et al., 1992; Mitchener et al., 1992). 2.3. Sediment provenance Heavy mineral suites and isotopic data both suggest that the Brent Group was derived from multiple provenance areas (Hamilton, Fallick, & MacIntyre, 1987; Mearns, 1992; Mitchener et al., 1992; Morton, 1992). These data indicate that Brent Group sediments were supplied by several river systems that drained the areas to the west of the UK Brent Province and to the east of the Norwegian Brent Province, in addition to a river system which drained the uplifted central North Sea Triple Junction area to the south (Mearns, 1992; Mitchener et al., 1992; Morton, 1992).

3. Database The Brent Group comprises shallow marine, marginal marine and non-marine deposits of Middle Jurassic (AalenianBathonian) age. Five lithostratigraphic units are recognised in the Brent Group: the Broom, Rannoch, Etive, Ness Detailed interpretation of the sedimentology and stratigraphy of the Brent Group is not possible using lowresolution, regional 2D seismic data. 3D seismic data offer

Fig. 3. Schematic lithostratigraphy of the Brent Group and its regional variability (after Brown et al., 1987; Deegan & Scull, 1977).

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

461

present a brief summary of previous facies analyses, combined with our own observations. The resulting facies association scheme is summarised in Table 2, and forms the basis for recognition of facies discontinuities across key sequence stratigraphic surfaces. The robust and consistent recognition of such surfaces in cores and wireline logs is essential for high-resolution (i.e. sub-seismic) application of sequence stratigraphic methods. 4.1. Facies associations A number of facies associations have been identied in core, based on lithology, primary sedimentary structures, bioturbation fabric and the nature of bedding contacts with underlying and overlying units (Table 2). Facies associations represent a variety of shallow marine, marginal marine and non-marine environments that may be classied into three broad groups (Table 2): (1) weakly waveinuenced shallow marine, (2) wave-dominated shallow marine and marginal marine, and (3) lagoonal and nonmarine. The last two groups of facies associations occur within the Rannoch, Etive and Ness Formations and can be readily accommodated in the widely used facies model developed for the Brent Group by Budding and Inglin (1981), Fig. 5). These facies associations have been comprehensively described by previous workers (e.g. Jennette & Riley, 1996; Livera, 1989; Livera & Caline, 1990; Richards & Brown, 1986; Scott, 1992) and they are not discussed further here. The rst group of facies associations is poorly documented in existing literature, and is described briey below. 4.2. Weakly wave-inuenced shallow marine deposits The weakly wave-inuenced shallow marine facies association occurs exclusively in the Broom and Tarbert Formations (Table 2), and is not represented in the conventional Brent Group facies model (Fig. 5). The association comprises three depositional elements in vertical section: (1) thick (10 25 m), upward-coarsening, silty sandstone successions (facies 1.2, 1.3; Table 2), (2) thin (, 5 m), upward ning, silty sandstone successions (facies 1.2; Table 2), and (3) erosively based blocky sandstones of variable thickness (2 20 m; facies 1.4; Table 2). These elements are each described and interpreted below. 4.2.1. Upward-coarsening, silty sandstone successions Upward-coarsening successions are characterised by intense bioturbation in their lower part (facies 1.2; Table 2; Fig. 6D). Bioturbation intensity decreases in the upper part of the successions, where bed thickness is greater and primary sedimentary structures are preserved (facies 1.3; Table 2). Here beds are normally graded and contain highangle trough and tabular cross-bedding (Fig. 6E), with some cross-beds exhibiting a bimodal grain size distribution and/ or clay drapes along their foresets (Fig. 6C). The intensely

Fig. 4. Map showing the distribution of wells within the studied dataset relative to major Brent Group reservoirs (Fig. 1) and key Middle Jurassic structural features (Fig. 2B). The wells and correlation panels shown in Figs. 7 and 11 are located. The area covered by isopach and palaeogeographic maps (Figs. 12 and 13) is also shown.

improved vertical resolution (e.g. approximately 15 20 m for fault analysis), but this remains too low for detailed stratigraphic analysis in the thin (, 150 m) Brent Group reservoirs of the eastern UK Brent Province (e.g. east of the Ninian Hutton Dunlin fault system, Fig. 2B). Instead, high-resolution data in the form of cores and wireline logs are essential for detailed study. A database of 62 key wells was selected for this study, on the basis of core coverage and geographical distribution (Fig. 4). 58 of these wells are cored, yielding over 5000 m of core that was examined in detail. Correlation between these wells was constrained by a regional, palynologically based biostratigraphic scheme (Simon Petroleum Technology Ltd, 1994) supplemented by additional biostratigraphic sampling.

4. Facies analysis and key sequence stratigraphic surfaces Many previous authors have described the detailed sedimentology of various parts of the Brent Group (e.g. Jennette & Riley, 1996; Livera, 1989; Livera & Caline, 1990; Richards & Brown, 1986; Scott, 1992). Here we

462

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

Table 2 Summary sedimentology of facies associations in the Brent Group Lithofacies association Lithology and sedimentary structures Bioturbation

1. Weakly wave-inuenced shallow marine facies associations (Broom Formation, Tarbert Formation) 1.1. Offshore transition (OT) Mudstone and siltstone with rare (, 25%) laminae and beds of very ne- to ne-grained sandstone. Parallel lamination, wave and current ripple cross-lamination 1.2. Distal tide-inuenced sheet Micaceous, lower ne-grained to lower mediumsandstone (dTSS) grained silty sandstones (80100%). Intense bioturbation almost completely obscures primary physical structures, but silty and micaceous laminations at 15 cm spacing record original bed tops. Rare current ripples and cross-beds. Rare marine body fossils (e.g. belemnites) 1.3. Proximal tide-inuenced sheet Micaceous, lower ne-grained to lower mediumsandstone (pTSS) grained silty sandstones (100%). Trough and tabular cross-beds with intensely bioturbated bed tops. Bimodal grain-size distribution and clay drapes in some cross-beds. Rare planar lamination and current ripples. Abundant drifted plant material 1.4. Tide-inuenced channel-ll Well-sorted to moderately sorted lower medium- to sandstone (TCS) coarse-grained sandstones (100%) in erosively based bodies 220 m thick. Trough and tabular cross-beds with sparsely bioturbated bed tops. Bimodal grain-size distribution and clay drapes in some cross-beds

Moderate to intense (BI 35; Planolites, Terebellina, Paleophycus, Rosselia, Thalassinoides, Teichichnus, Asterosoma, mud-lled Arenicolites) High to complete (BI 4 6: Anconichnus, Thalassinoides, Arenicolites, Skolithos, Planolites, Palaeophycus, Ophiomorpha, Teichichnus, Cylindrichnus). Teichichnus-Anchonichnus ichnofabric of Goldring et al. (1991).

Moderate to intense (BI 35; Thalassinoides, Arenicolites, Palaeophycus, Cylindrichnus, Ophiomorpha)

Absent to low (BI 02; Ophiomorpha, Thalassinoides, Arenicolites)

2. Wave-dominated shallow marine and marginal marine facies associations (Rannoch Formation, Etive Formation, Tarbert Formation) Generally sparse to low (BI 12; 2.1. Lower, wave-dominated shoreface (LSF) Upward-coarsening succession of micaceous, very Ophiomorpha, Arenicolites, Thalassinoide), ne to lower ne-grained sandstones (70100%) with mudstone and siltstone interbeds in its lower part. but moderate to intense at preserved bed tops (BI: 35; Planolites, Terebellina, Sandstone bed amalgamation increases upwards. Paleophycus, Rosselia, Thalassinoides, Low-angle and hummocky cross-stratication, minor Teichichnus, Asterosoma, Arenicolites) current and wave ripple cross-lamination in sandstone beds 2.2. Upper, wave-dominated shoreface (USF) Upper ne-grained sandstone (100%). Trough and Sparse to low (BI 12; Ophiomorpha, tabular cross-beds, planar lamination and swaly Skolithos) cross-stratication Absent (BI 0) 2.3. Foreshore (FS) Upper ne-grained sandstone (100%). Gently wedging to sub-parallel laminations, low amplitude current ripples. Enrichment in microscopic heavy mineral grains produces high (. 100 API) gamma ray values Sparse to moderate (BI 13; Skolithos, 2.4. Barrier (BAR) Extremely well sorted, upper ne-grained to Ophiomorpha, Arenicolites, Palaeophycus) lower medium-grained sandstone (100%). Indistinctly stratied, mottled, structureless and/or root-penetrated. Rare cross-beds, current ripples, planar-parallel lamination and minor dewatering structures 2.5. Tidal inlet/estuarine Well-sorted, lower medium-grained sandstone Generally absent (BI 0), but low at channel-ll (T) (80 100%) in erosively based bodies 2 15 m thick. preserved bed tops (BI 2; Arenicolites) Cross-beds and current ripples with rare clay drapes. Bodies may be amalgamated into thick (15 30 m) multistorey channel-ll complexes 2.6. Washover fan/ood-tidal Upward-coarsening succession of micaceous, Sparse to low in sandstone beds (BI 12; delta (WF) upper ne-grained to lower medium-grained Ophiomorpha, Palaeophycus) and sparse to sandstone (6090%) with mudstone interbeds. moderate in mudstones (BI 1 3; Planolites, Wave and current ripple cross-lamination, minor mud-lled Arenicolites) cross-beds and planar-parallel lamination in sandstone beds 3. Lagoonal and non-marine facies associations (Ness Formation) 3.1. Lagoon/lake (L) Mudstone and siltstone with rare (, 25%) laminae and beds of very ne- to ne-grained sandstone. Parallel lamination, wave and current ripple crosslamination and rare soft-sediment folding Sparse to moderate (BI 13; Planolites, mud-lled Arenicolites). Some successions contain additional traces (Thalassinoides, Teichichnus). (Continued on next page)

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484 Table 2 (continued) Lithofacies association 3.2. Fluvial/wave-inuenced delta (lagoonal shoreface, LS) Lithology and sedimentary structures Upward-coarsening succession of very ne to lower coarse-grained sandstones (50100%) with mudstone and siltstone interbeds in its lower part. Sandstone beds thicken upwards. Wave and current ripple cross-lamination, low-angle cross-lamination, climbing current ripples, planar-parallel lamination and cross-beds in sandstone beds Poorly to moderately sorted, very ne- to coarsegrained sandstone (80100%) in erosively based bodies 515 m thick. Amalgamation into thick (1550 m) multistorey channel-ll complex. Cross-beds and current ripples Mudstone and siltstone with very ne- to mediumgrained sandstone laminae and beds 1 200 cm thick (1080%). Cross-beds, current ripples, dewatering structures and soft-sediment folds. Isolated roots and/or pedogenic features are pervasive Coal and carbonaceous shale Bioturbation Generally sparse to moderate (BI 13; Skolithos, Thalassinoides, Arenicolites, Teichichnus)

463

3.3. Fluvial channel-ll (F)

Generally absent (BI 0), but locally sparse to moderate (BI 13; Taenidium, Planolites, mud-lled Arenicolites)

3.4. Aggradational oodplain (AF)

Absent to moderate (BI 0 3; Planolites, mud-lled Arenicolites)

3.5. Coal

Absent (BI 0)

Bioturbation is described using the bioturbation index (BI) scheme of Taylor and Goldring (1993).

bioturbated and upward-coarsening character of the successions implies shallowing-upward deposition on a prograding, weakly wave-inuenced shoreface, similar to those described from the Fulmar Formation of the Central Graben, North Sea (e.g. Howell, Flint, & Hunt, 1996; Johnson & Stewart, 1985; Taylor & Gawthorpe, 1993). The diverse trace fossil assemblages that characterise the successions indicate deposition under fully marine conditions (e.g. a Teichichnus Anchonichnus ichnofabric in facies 1.3; Goldring, Pollard, & Taylor, 1991; Taylor & Gawthorpe, 1993). However, primary sedimentary structures record

episodic sand deposition from unidirectional currents with no subsequent wave reworking. The bimodal grain size distribution and clay drapes in some cross-beds are interpreted, respectively, to record sorting by tidal processes and clay deposition from suspension during slack-water periods of the tidal cycle. Thus, tides are an inuential depositional process, although tidal indicators are not pervasive. The successions are typically 10 25 m thick and can commonly be correlated over several kilometres to tens of kilometres, and hence we refer to them as regressive tide-inuenced sandstone sheets.

Fig. 5. Conceptual model for deposition of the Rannoch, Etive and Ness Formations, after Budding and Inglin (1981). The model was originally conceived for the South Cormorant Field, but has since been widely applied to many other Brent Group reservoirs.

464

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

Fig. 6. Photographs showing sedimentological aspects of the weakly wave-inuenced shallow marine facies association (Table 2) and key sequence stratigraphic surfaces in the Tarbert Formation. (A) Coarse-grained lag containing angular intrabasinal and sub-angular extrabasinal clasts at the base of the Tarbert Formation, which marks a major sequence boundary (SB1000; 115330 in 211/27-10). The sequence boundary juxtaposes a tide-inuenced sheet sandstone (Tarbert Formation) and underlying lagoonal mudstones (Ness Formation). (B) Coarse-grained lag containing sub-angular extrabasinal clasts near the base of the Tarbert Formation (107100 in 2/5-17; Fig. 7B). Extrabasinal clasts have been transgressively reworked from the base-Tarbert sequence boundary (SB1000). The lag has been undergone post-depositional calcite cementation. (C) Coarse-grained lag containing sub-rounded extrabasinal clasts that have been transgressively reworked from the base-Tarbert sequence boundary (SB1000; 95200 in 3/4-12; Fig. 7A). Note the presence of clay drapes in sandstones above the lag. (D) Intensely bioturbated silty sandstone (facies 1.2; Table 2) within a distal tide-inuenced sheet sandstone (106790 in 2/5-17; Fig. 7B). The sandstone contains belemnite guards and sub-angular extrabasinal clasts, which were reworked from the base-Tarbert sequence boundary (SB1000). Trace fossils include Anconichnus, Terebellina, Teichichnus and Cylindrichnus. (E) Cross-bedded silty sandstone (facies 1.3; Table 2) within a proximal tide-inuenced sheet sandstone (94720 in 3/4-12; Fig. 7A). Cross-bedding is dened by micaceous laminations. (F) Carbonaceous roots capping a proximal tide-inuenced sheet sandstone (94600 in 3/4-12; Fig. 7A). The roots have been truncated and transgressively reworked at a ooding surface (FS1050).

4.2.2. Upward-ning, silty sandstone successions Upward-ning successions comprise only intensely bioturbated silty sandstones (facies 1.2; Table 2) and pass gradationally upwards into offshore mudstones. The diverse trace fossil assemblage within the successions (Teichichnus Anchonichnus ichnofabric of Goldring et al., 1991) indicates deposition in an offshore marine environment. Thus, these thin (, 5 m) successions record deepeningupward deposition that culminated in offshore mudstones. We regard these successions as the transgressive counterparts to the regressive tide-inuenced sandstone sheets described above. 4.2.3. Erosively based blocky sandstones Erosively based blocky sandstones (facies 1.4; Table 2) are dominated by medium-scale (ca. 30 cm thick) crossbeds, a few of which contain bimodal grain-size distributions and/or clay drapes. Bioturbation is sparse in this

facies. Coarse-grained lags of extrabasinal sediment occur at the base of several blocky sandstones, which lie at specic stratigraphic horizons (e.g. Fig. 6A D). The erosive bases of the sandstones and their laterally discontinuous character imply that they represent channel-ll successions. Cross-beds are interpreted to result from subaqueous dune migration in the channels, and the presence of sparse Ophiomorpha, Thalassinoides and Arenicolites at bed tops implies deposition in a shallow marine or marginal-marine environment. Where present, the bimodal grain size distribution and clay drapes in some cross-beds are interpreted as tidal indicators. These successions are similar to sand-dominated tidal inlet and/or estuarine channel-lls observed within the Etive Formation (facies 2.5; Table 2), which are associated with wavedominated shoreface and barrier deposits (facies 2.1 2.4, 2.6; Table 2). However, channel-ll sandstones in the Broom and Tarbert Formations cut into, and are interpreted

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

465

to pass down-dip into, regressive tide-inuenced sheet sandstones. We refer to these successions as tide-inuenced channel-ll sandstones. 4.3. Sequence boundaries and forced regressive deposits Major, low-frequency sequence boundaries within the Brent Group are interpreted at the base of the Broom Formation (SB100) and at, or near, the base of the Tarbert Formation (SB1000; Figs. 6A C and 7; Mitchener et al., 1992; Morton, 1992). Both surfaces are characterised by an abrupt inux of coarse-grained, extrabasinal material from the basin anks (Mitchener et al., 1992; Morton, 1992), and biostratigraphic data imply a signicant time gap (. 1 Ma) across both surfaces over much of the East Shetland Basin. The base-Broom sequence boundary (SB100) records a major basinward shift (. 100 km) in shallow marine sedimentation, and is interpreted to overlie a regionally extensive angular unconformity produced by thermal doming in the central North Sea (Underhill & Partington, 1993). The origin of the base-Tarbert sequence boundary (SB1000) is more cryptic and its regional extent and morphology are described later. Both sequence boundaries are overlain by depositional systems composed of the weakly wave-inuenced shallow marine facies association (Table 2), within which tidal processes are more inuential than waves and storms. We also interpret several higher frequency sequence boundaries and their correlatives within the Rannoch, Etive, Ness and Tarbert Formations. In the northern part of the Brent Province, the Rannoch shoreface system is locally sharp-based (sensu Plint, 1988) and its basal surface is marked by an abrupt inux of medium-grained sand and/or a mudclast lag (e.g. within well 211/19-6 in Fig. 7D; Reynolds, 1995). In a handful of wells, hummocky-cross-stratied sandstones of the proximal lower shoreface (facies 2.1; Table 2) are juxtaposed directly on offshore mudstones (facies 1.1; Table 2) across the base-Rannoch surface, implying signicant erosion and/ or non-deposition. These observations imply that the northerly part of the Rannoch Formation was deposited during a forced regression caused by a fall in relative sealevel (e.g. within well 211/19-6 in Fig. 7D; Reynolds, 1995). The sharp base is not laterally continuous on a eld- or a regional-scale, suggesting that the forced regression was not continuous, but was punctuated by episodes of normal regression during which relative sea-level rose. Similar vertical successions and lateral variability are documented in forced regressive, sharp-based shoreface deposits exposed at outcrop (Hampson, 2000). Thus the Rannoch Formation exhibits some evidence for deposition during the relative sea-level fall that produced laterally variable incision at the base of the Etive Formation (see review in Olsen & Steel, 2000). Incised valleys, with high-frequency sequence boundaries at their bases, are interpreted in the Etive and Ness Formations (SB300, SB350, SB400, SB500, SB550, SB600,

SB700; Fig. 7). In cores and wireline logs, each candidate valley ll is characterised by the stacking of single-storey channel-ll bodies into a considerably thicker (ca. 30 m), multistorey body, which commonly has a distinctive internal facies architecture that is interpreted to reect increasing accommodation space during valley lling. For example, candidate valley lls in the Etive and Ness Formations generally have a lower, uvial component (facies 3.3; Table 2) and an upper, tide-inuenced component (facies 2.5; Table 2). Candidate valley lls generally lack an abrupt basinward shift in facies across their bases, but in areas of dense well-spacing the candidate valleys can be demonstrated to be laterally discontinuous and deeply erosive (e.g. in the Cormorant Field; Jennette & Riley, 1996). In some cases, the valleys erode through underlying stratigraphic markers, such as eld-wide coal seams. Several candidate valley lls also have a basal lag of anomalously coarse-grained, extrabasinal sand and pebbles (Fig. 8A). A number of candidate valley lls in the Etive Formation comprise stacked, tidal inlet/estuarine channelll (facies 2.5; Table 2) and barrier sandstones (facies 2.4; Table 2) above a coarse-grained, extrabasinal lag. In these cases, the basal lag is the only feature that allows a valleyll origin to be interpreted, rather than an unconned aggradational barrier system. A valley-ll interpretation for these Etive successions implies that an idealised, Etive Ness valley ll comprises a wave-dominated barrier system in its seaward part and a tide-inuenced uvial system in its landward part. Thus, we interpret Etive Ness valleys to have been cut during episodes of relative sea-level fall and inlled by wave-dominated estuaries that were developed under microtidal or mesotidal conditions during subsequent relative sea-level rise (as portrayed in the models of Zaitlin, Dalrymple, & Boyd, 1994). Interuves to the Etive Ness incised valley lls are marked by palaeosols and can be condently recognised only in areas of high well-density, where they are directly overlain by the ooding surface or coal seam that caps the correlative valley ll. Most palaeosols in the Ness Formation (including interuve palaeosols at SB350, SB400) are medium- to dark-grey gley and semi-gley palaeosols that contain abundant carbonaceous material and which formed under permanent and semi-permanent waterlogged drainage conditions (Fig. 8B; e.g. Besly & Fielding, 1989). In contrast, several interuve palaeosols (SB550, SB600, SB700) are characterised by: (1) mottled, red and pale grey colouration, (2) roots and rootlets lacking preserved carbonaceous material, and (3) auto-brecciated mudstones with variable calcite content (Fig. 8C and D). Red mottling and non-preservation of carbonaceous rootlining material indicate partial oxidation (Retallack, 1990) and auto-brecciated mudstones imply complete dessication of the upper palaeosol horizons. Calcite within the palaeosols occurs either as sparse calcrete nodules, implying upward transport of carbonate-bearing groundwater during soil formation (Retallack, 1990), or as crack-lling cement

466

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

Fig. 7. Wireline logs and simplied core logs through four type wells: (A) well 3/4-12, from the Strathspey Field in the hangingwall of the NinianHutton Dunlin fault system, (B) well 2/5-17, from the Heather Field near the southwestern margin of the Brent Province, (C) well 211/16-6, from the Eider Field on the TernEider Horst, and (D) 211/19-6 from the Murchison Field near the northern margin of the Brent Province. The core and wireline log character of reservoir facies associations (Table 2) is illustrated, and biostratigraphic markers (Table 3) and key sequence stratigraphic surfaces (Fig. 10) identied in the well are highlighted. The location of the wells is shown in Figs. 4 and 11.

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

467

Fig. 7 (continued )

468

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

Fig. 7 (continued )

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

469

Fig. 8. Photographs showing sedimentological aspects of sequence boundaries in the Ness Formation. (A) Basal part of a multistorey, uvial channel-ll sandstone in the Upper Ness Formation (97080 in 3/4-12; Fig. 7A), containing granule-sized, lithic extraclasts (labelled e) and mudstone intraclasts (labelled i). This sandstone is interpreted to overlie a sequence boundary (SB600). (B) Carbonaceous root traces overprinting oodplain deposits (97410 in 3/4-12; Fig. 7A). The roots are interpreted as part of a poorly developed gley palaeosol, which is typical of oodplain deposits within the Ness Formation. (C) Distinctive palaeosol in the Upper Ness Formation (97350 in 3/4-12; Fig. 7A), which comprises mottled red-brown siltstone containing green-coloured root traces (labelled r) and calcite-lled, post-depositional fractures (labelled c) around rhizoconcretions. The high degree of mottling and abundance of rhizoconcretions implies prolonged soil formation, while the pervasive red-brown colour and scarcity of carbonaceous material suggests soil development under oxidising conditions above the water table. In the context of the coal-prone Ness Formation, this palaeosol records anomalous groundwater drainage conditions and is interpreted to mark a sequence boundary (SB600). (D) Another distinctive palaeosol, containing abundant red siderite rhizoconcretions and rare carbonaceous root traces overprinting pale grey sitstones, in the Upper Ness Formation (95090 in 3/4-9). This palaeosol is also interpreted to record pedogenesis under predominantly oxidising conditions at a sequence boundary (SB600), and is interpreted to correlate to the palaeosol in (C).

of uncertain origin. In combination, these distinctive features imply soil development under anomalously welldrained conditions above the regional water table, consistent with an interuve origin (e.g. Hampson, Elliott, Davies, & Flint, 1997; McCarthy, Faccini, & Plint, 1999). The palynofauna associated with one such interuve palaeosol (SB600) indicates regionally improved drainage or increased aridity (Simon Petroleum Technology Ltd, 1994). Thick (up to 30 m), tide-inuenced channel-ll sandstones (facies 1.4; Table 2) containing additional inuxes of coarse-grained, extrabasinal sand occur within the Tarbert Formation. The channel-ll sandstones erode deeply into upward-coarsening, regressive tide-inuenced sheet sandstones and are overlain by transgressive tide-inuenced sheet sandstone successions. The stratigraphic context of these channel-ll sandstones, at the transition from regression to transgression, combined with their deep basal erosion and extrabasinal provenance suggests that they occur within incised valleys that dene intra-Tarbert sequence boundaries (SB1100, SB1200). 4.4. Flooding surfaces and transgressive deposits A hierarchy of ooding surfaces is recognised in the Brent Group, with the order of a particular ooding surface reected in the magnitude of associated landward facies shift and the stacking pattern of units above and below the surface. Major, low-frequency ooding surfaces occur at the top of the Broom Formation (FS200) and the top of

the Tarbert Formation (FS1200). Both of these surfaces are associated with major landward facies shifts (. 50 km) that place offshore mudstones above shallow-marine sandstones. The top-Broom maximum ooding surface (FS200) records a change from a net-transgressive depositional system that consisted of weakly wave-inuenced shallow marine facies (facies 1.2 1.4; Table 2) to a net-regressive depositional system that consisted of wave-dominated shallow marine and marginal marine facies (facies 2.1 2.6; Table 2). The top-Tarbert ooding surface (FS1200) records the nal drowning of the Brent delta, and it caps a net-transgressive depositional system that consisted of weakly wave-inuenced shallow marine facies (facies 1.2 1.4; Table 2). Higher frequency ooding surfaces occur within the Etive, Ness and Tarbert Formations (e.g. FS300, FS350, FS400, FS500, FS800, FS850, FS900, FS1000, FS1050, FS1100, FS1150; Fig. 7). Within the Ness Formation, ooding surfaces occur at the base of lagoonal and lacustrine mudstones (facies 3.1; Table 2) that overlie coal seams (facies 3.5; Table 2). Such coal seams are prominent, reservoir-scale correlation markers within lagoonal successions (e.g. within the Brent Field, Livera, 1989) and may also form regionally extensive seismic markers (e.g. the coal below FS800 is picked as a base-Tarbert reector by Davies et al., 2000; McLeod et al., 2000). The more signicant of these ooding surfaces (FS300, FS350, FS400, FS500, FS800, FS850, FS900) are associated with development of deeper and more laterally extensive lagoons with marine palynofacies. Several of these lagoonal ooding surfaces overlie thin (, 5 m), upward-ning intervals above a coal

470

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

seam. These upward-ning intervals commonly contain sharp-based, well-sorted, medium-grained sandstones containing a limited assemblage of trace fossils (Ophiomorpha, Thalassinoides, Skolithos, Arenicolites, Palaeophycus; Fig. 9A C). The textural maturity of the sandstones implies that they have been reworked from a shoreface or barrier, while the trace fossil assemblage has a marginal marine aflinity. Several of the sandstones have unlined Thalassinoides burrows at their base, which are interpreted to constitute a rmground, Glossifungites ichnofacies (MacEachern, Raychaudhuri, & Pemberton, 1992; Fig. 9A C). This interpretation implies that the surfaces were exhumed by transgressive erosion. In combination, the features described above suggest that barrier sands were reworked during transgression, either in large washover fan systems or as a result of barrier breaching and destruction during its rapid retreat. The ning-upward intervals that contain these

sandstones represent transgressive deposition that culminated in development of a lagoonal ooding surface. The development of deep (up to 20 m) tidal channel lls (facies 2.5; Table 2) is associated with several ooding surfaces in the uppermost Ness and lowermost Tarbert Formations. These deep channels overlie transgressive erosion surfaces and erode deeply into underlying lagoonal deposits (Fig. 9E). Flooding surfaces within the Etive Formation are more difcult to identify, particularly in successions dominated by channel-ll sandstones that erode underlying strata. However, several ooding surfaces juxtapose the deposits of barrier systems (facies 2.4 2.6; Table 2) above coal seams (facies 3.5; Table 2) or rooted horizons that in turn overlie older barrier system deposits (facies 2.4 2.6; Table 2). In the northern part of the study area, the Etive Formation comprises several stacked barrier systems each bounded by such ooding

Fig. 9. Photographs showing sedimentological aspects of ooding surfaces in the Ness Formation. (A) Large unlined burrows (Thalassinoides?), constituting a Glossifungites ichnofacies, at a transgressive erosion (ravinement) surface (94310 in 3/4-9). The erosion surface truncates root-penetrated oodplain deposits and is overlain by lagoonal mudstones. Coarse-grained sand inlling the burrows may have been reworked along depositional strike from an underlying incised valley ll (overlying SB600). (B) A transgressive erosion surface, characterised by a Glossifungites ichnofacies, overlain by a thin (60 cm), well-sorted, medium-grained sandstone bed (96110 in 3/4-12; Fig. 7A). The surface overlies a thin interval of lagoonal heteroliths and a coal (base of photograph). (C) Burrows (Arenicolites, Ophiomorpha) in the sandstone bed overlying this transgressive erosion surface (B), implying marine inuence. (D) Thin (5 cm), bioturbated (Arenicolites, Thalassinoides), ne-grained sandstone bed enclosed within root-penetrated oodplain deposits (124340 in 211/27-11). The trace fossil assemblage implies marine inuence, and consequently the bed is interpreted to overlie a ooding surface. (E) Transgressive erosion surface juxtaposing a thick (15 m) well-sorted, medium-grained sandstone unit above lagoonal heteroliths (90220 in 211/16-6; Fig. 7C). Correlation demonstrates that the sandstone has signicant erosive relief at its base (Tarbert Formation of Jennette & Riley, 1996, in the Eider Field), and it is interpreted as a tidal channel ll. (F) The top of this sandstone is overlain by offshore mudstones (Heather Shales). The contact between these units is marked by an unlined, mud-lled burrow (89800 in 211/16-6; Fig. 7C), interpreted as part of a Glossifungites ichnofacies, and coarse, extrabasinal sand grains that have been reworked into the offshore mudstones. The coarse sand grains are interpreted as evidence of a sequence boundary (SB1000) which overlay the thick, well-sorted, medium-grained sandstone unit in the lower part of the photograph and which was subsequently reworked at a transgressive erosion surface associated with the unlined burrow.

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

471

surfaces (e.g. Fig. 7D; Brown & Richards, 1989; Reynolds, 1995). These successions imply that the Etive Ness system was largely aggradational. 4.5. Core to wireline-log calibration Widespread use of the facies scheme (Table 2) and sequence stratigraphic interpretations described above relies on the accurate calibration of core and wireline-log data. Several facies associations possess a distinctive wireline-log character, particularly the lagoonal and coastal plain facies associations of the Ness Formation, which generally display pronounced variations in lithology and consequent wirelinelog response (e.g. Bryant & Livera, 1991; Livera, 1989). However, differentiating uvial channel-ll sandstones (facies 3.3; Table 2) and estuarine channel-ll sandstones (facies 2.5; Table 2), which most likely occur within incised valleys, is not possible from wireline-log data alone, but requires core interpretation (e.g. Flint et al., 1998). The various sand-dominated, shallow marine facies of the Broom, Rannoch, Etive and Tarbert Formations possess less distinctive wireline-log characteristics. Concentrations of heavy minerals are diagnostic of wave-dominated foreshore deposits (facies 2.3; Table 2), and thus sandprone successions containing thin (, 2 m), high gamma-ray

spikes (e.g. at 107680 in well 211/19-6; Fig. 7D) are likely to comprise wave-dominated facies (facies 2.1 2.6; Table 2). Tidal inlet/estuarine and uvial channel-ll deposits (facies 2.5 and 3.3, respectively; Table 2) in the Etive Formation are difcult to distinguish, because each is characterised by abrupt increases in sand content, grain size and porosity across their bases. In our experience, detailed interpretation of the Broom, Rannoch and Tarbert Formations in uncored wells must rely heavily on calibration with nearby cored wells. Inter-well correlation using key sequence stratigraphic surfaces is essential to such calibration, because this methodology provides a framework in which core-based facies trends may be extrapolated using appropriate depositional models.

5. Regional high-resolution sequence stratigraphic framework The key surfaces described above, and the units that they bound, have been correlated to construct a regional, high-resolution sequence stratigraphic framework. Correlation between the wells selected for this study (Fig. 4) was constrained by a regional, palynologically based biostratigraphic scheme (Table 3, based on Simon

Table 3 Regional and sub-regional biostratigraphic markers used in this study (after Simon Petroleum Technology Ltd, 1994) Palynostratigraphic event 1: 2: 3: 4: Base of dinocyst Sirmiodinium grossii Acme occurrence of dinocyst Dissiliodinium willei Base acme occurrence of Dissiliodinium willei Base of common occurrence of dinocyst Sentusidinium granulatum Stratigraphic interpretation and condence level Reliable, regional marker in the lower part of the Heather Formation Reliable, regional marker Reliable, regional marker Sub-regional marker associated with the onset of marine sand deposition above two major ooding surfaces: FS1000 in the southern UK Brent Province and FS850 in the northern UK Brent Province Regional marker occurring above a major sequence boundary, SB1000, at the base of the Tarbert Formation. The true base of Cerebropollenites mesozoicus is not encountered due to the stratigraphic break across the sequence boundary Reliable, regional marker at a major ooding surface, FS800, marked by widespread development of lagoons Reliable, regional marker that is interpreted to record regionally improved drainage around a sequence boundary, SB600 A regional event usually associated with the Mid-Ness Shale (FS400 FS500). It is interpreted to reect an increase in salinity in the widespread Mid-Ness lagoon Reliable, regional marker interpreted to reect an increase in salinity (from freshwater/oligohaline to more mesohaline) across a major ooding surface, FS400, in the lower part of the Mid-Ness Shale Sub-regional marker interpreted to reect the development of lagoons in the southern UK Brent Province above two major ooding surfaces, FS350 and FS300 Reliable, regional marker that coincides with a major ooding surface, FS200, near the base of the Rannoch Formation A distinctive and diverse microplankton assemblage in the upper part of the Dunlin Group, consistent with deposition in a normal, shelfal marine environment

5: Base increase in sporomorph Cerebropollenites mesozoicus

6: Base increase in miospore Lycopodiumsporites spp. 7: Acme of Classopollis spp. 8: Signicant inux of marine microplankton, particularly Nannoceratopsis gracilis 9: Decrease in abundance of fresh to brackish water alga Botryococcus spp.

10: Signicant inux and top increase of marine microplankton, particularly Nannoceratopsis gracilis and N. senex 11: Top acme of Nannoceratopsis spp. 12: Top Parvocysta group, Nannoceratopsis triangulata and Pareodinia junior

These markers have been used to constrain our sequence stratigraphic correlations, supplemented by additional markers of local, intra-eld extent (Figs. 7, 10 and 11).

472

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

Fig. 10. Summary of the high-resolution sequence stratigraphic framework presented in this paper. Absolute ages and the regional North Sea stratigraphic scheme (3rd order cycles) are after Mitchener et al. (1992) and Rattey and Heyward (1993). Low-frequency sequence boundaries occur at the base of the Broom and Tarbert formations (SB100 and SB1000, respectively). Seven higher frequency sequence boundaries occur in the RannochEtiveNess interval (FS200SB1000), dening sequences with an average duration of ca. 1 Ma. Two higher frequency sequence boundaries occur in the Tarbert Formation (SB1000FS1200), dening sequences with an average duration of ca. 3 Ma.

Petroleum Technology Ltd, 1994). Published reservoir zonations in individual elds have also been used to guide correlations, where supported by sedimentological, stratigraphic and/or production data (Table 1). Our highresolution sequence stratigraphic framework is summarised in Fig. 10, and its key features illustrated in a series of cross-sections (Fig. 11), isopach maps (Fig. 12) and palaeogeographic reconstructions (Fig. 13). We have adopted and expanded the nomenclature of Jennette and Riley (1996) for the key sequence stratigraphic surfaces in our framework. 5.1. SB100 FS200 (Broom Formation) The base of the Broom Formation is a major sequence boundary (SB100; Figs. 10 and 11) marked by an abrupt basinward shift of facies from offshore shales (facies 1.1; Table 2) to tide-inuenced channel-ll and sheet sandstones (facies 1.2 1.4; Table 2). The top of the Broom Formation is a major ooding surface (FS200; Fig. 10) that is marked by an incursion of offshore shale in the northern part of the study area (e.g. wells 211/19-1, 211/19-5, 211/19-3 and 211/ 19-6 in Fig. 11F). In the southern part of the study area, this ooding surface occurs at the top of medium-grained tideinuenced sheet sandstones (facies 1.2 1.3; Table 2) that locally comprise the lower part of the Rannoch Formation (e.g. 107730 in well 2/5-17, Fig. 7B).

The Broom Formation has a complex internal stratigraphy, but its lower part generally consists of an erosively based blocky sandstone, interpreted as a tide-inuenced channel ll (facies 1.4; Table 2). In the eastern and northern part of the study area, these channels pass laterally into and/or incise through an upward-coarsening, tide-inuenced sheet sandstone succession (facies 1.2 1.3; Table 2; e.g. 102940 103200 in well 3/4-12, Fig. 7A). We interpret this stratigraphic architecture to record uvio-tidal incision through falling-stage and lowstand shoreline deposits (Fig. 13A). In the north-eastern part of the study area, at least one incised channel-ll sandstone is enclosed within offshore shales (e.g. well 211/19-6 in Figs. 11D and 13A). Thus, detached lowstand shoreline deposits may lie further down-dip, at the seaward termination of this incised channel. The upper part of the Broom Formation comprises one or more upward-coarsening, tide-inuenced sheet sandstone successions (facies 1.2 1.3; Table 2) that pass down-dip into offshore shales. These upward-coarsening successions are thicker and more common in the southwestern part of the study area (e.g. wells 2/5-17, 2/5-3, 3/11 and 3/2-3 in Fig. 11A) and in the immediate hangingwall of the Ninian Hutton Dunlin fault system (e.g. well 211/ 18-5 in Fig. 11B), as reected in isopach maps of this stratigraphic interval (Fig. 12A). We interpret these successions as retrogradationally stacked deposits that record net transgression of the Broom depositional system.

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

473

Fig. 12. Isopach maps of selected stratigraphic intervals: (A) SB100FS200, corresponding to the Broom Formation; (B) FS200FS300, corresponding to the Rannoch and Etive Formations; (C) FS300FS400, corresponding to the lower Ness Formation; (D) FS400FS500, corresponding to the middle Ness Formation; (E) FS500FS800; corresponding to the upper Ness Formation; and (F) SB1000FS1000, corresponding to the lower Tarbert Formation.

474

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

Their distribution and thickness variations imply that sediment was transported from the southwest of the study area and along the southern part of the Ninian Hutton Dunlin fault hangingwall (e.g. Fig. 13A and B). Isopach maps and palaeogeographic reconstructions also suggest

that there was a third sediment transport route along the northern margin of the Tern Eider Horst, which acted as an intra-basin high (Figs. 12A and 13A and B). An unusually thick succession is observed locally in the area of the Cormorant Block IV Field, along the eastern margin

Fig. 13. Palaeogeographic reconstructions of selected stratigraphic intervals: (A) SB100, base-Broom sequence boundary; (B) SB100FS200, upper-Broom transgression; (C) FS200SB300, lower-Rannoch progradation (normal regression); (D) SB300, base-Etive sequence boundary and upper-Rannoch progradation (forced regression); (E) FS300, top-Etive transgression; (F) SB350, lower-Ness sequence boundary; (G) FS350, lower-Ness transgression; (H) FS400, mid-Ness transgression; (I) FS500, upper-Ness transgression; (J) SB600, upper-Ness sequence boundary; (K) SB700, upper-Ness sequence boundary; (L) FS800, upper-Ness transgression; (M) FS850, upper-Ness transgression; (N) FS900, upper-Ness transgression; (O) SB1000-FS1050, lower-Tarbert transgression; and (P) SB1100, mid-Tarbert sequence boundary and progradation (forced regression).

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

475

Fig. 13 (continued )

of the Tern Eider Horst (e.g. wells 211/21-5 and 211/21-6 in Figs. 11C and 12A). 5.2. FS200 FS300 (Rannoch and Etive Formations) The base of the Rannoch Formation generally coincides with a major ooding surface (FS200; Fig. 10) that is marked by an incursion of offshore shale in the northern part

of the study area (e.g. wells 211/19-1, 211/19-5, 211/19-3 and 211/19-6 in Fig. 11F). A widespread and prominent coal seam marks the top of the overlying Etive Formation over most of the study area. This coal seam is overlain by a regionally extensive lagoon across a major ooding surface (FS300; Fig. 10). The internal stratigraphy of the Rannoch and Etive Formations is simple. The Rannoch Formation comprises

476

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

Fig. 13 (continued )

a single, upward-shallowing wave-dominated shoreface succession (facies 2.1 2.3; Table 2) overlain locally by barrier deposits (facies 2.4; Table 2). Although this succession is thick, no major ooding surfaces are identied (Figs. 7 and 11). The Rannoch shoreface succession is gradationally based over most of the study area, implying normal regression under conditions of rising relative sea-level (Fig. 13C). However, the shoreface succession is

sharp-based (sensu Plint, 1988) in much of the northern part of the study area (e.g. 108690 in well 211/19-6 in Fig. 7D), implying forced regression under conditions of falling relative sea-level. The Etive Formation comprises stacked uvial and tidal inlet/estuarine channel-ll (facies 2.5, 3.3; Table 2) and barrier sandstones (facies 2.4; Table 2). Over much of the study area, the base of the formation is associated with an extrabasinal lag that is overlain by

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

477

a multistorey channel-ll body with a lower, uvial component (facies 3.3; Table 2) and an upper, tideinuenced component (facies 2.5; Table 2). These features imply an incised valley-ll origin. Etive valleys are interpreted to incise into the deposits of an unconned aggradational barrier system, which lack a basal extrabasinal lag. We interpret the initiation of Etive uvial incision to have been coincident with progradation of the sharp-based Rannoch shoreface in the northern part of the study area, with Etive valleys acting as conduits for sediment bypass to the forced-regressive Rannoch shoreface (Fig. 13D). The widespread extent of Etive valley-ll systems implies a prolonged period of incision and valleywidening, and possibly several discrete episodes of incision. The lack of lateral facies variability in the Rannoch Etive succession (Figs. 11, 13C and D) makes it difcult to interpret sediment transport routes and shoreline progradation directions, but studies in individual elds imply a tributary system of west east-trending Etive valleys (Jennette & Riley, 1996; Livera, 1989). The Rannoch Etive succession also exhibits gradual thickness variations across the study area, thickening abruptly only in the hangingwall of the Ninian Hutton Dunlin fault system (Fig. 12B). 5.3. FS300 FS400 (Lower Ness Formation) The Ness Formation comprises predominantly coastal plain deposits (facies 3.1 3.5; Table 2). Its internal stratigraphy is dened most clearly by ooding surfaces at the base of widespread lagoonal deposits. These lagoonal deposits commonly overlie, and are capped by, extensive coal seams (e.g. in the Brent Field; Livera, 1989) and they contain palynostratigraphic markers (e.g. regional and subregional events 6 10 in Table 3). Incised valley networks, which dene sequence boundaries, are also present within the Ness Formation, but their complex lateral distribution makes them unreliable stratigraphic markers. The most widespread and palynostratigraphically distinctive of the lagoonal deposits is known informally as the Mid-Ness Shale. We use the ooding surfaces that bound the MidNess Shale to subdivide the Ness Formation into lower, middle and upper units. The Lower Ness Formation is bounded at its base by the top-Etive ooding surface (FS300; Fig. 10) and at its top by the base-Mid-Ness Shale ooding surface (FS400; Fig. 10). This interval is relatively thin over the western part of the study area, but thickens considerably in the hangingwall of the Ninian Hutton Dunlin fault system (Fig. 12C). Flooding surfaces FS300 and FS350 are marked by the southward retreat of a west east-trending shoreline (Fig. 13E and G), whereas ooding surface FS400 is marked by the westward retreat of a north south-trending shoreline (Figs. 13H). All three transgressive shorelines comprise a wave-dominated shoreface and barrier system (facies 2.1 2.6; Table 2) that was associated with signicant transgressive erosion (e.g. sub-FS400 transgressive erosion surface at 99280 in well

3/4-12, Fig. 7A). Two incised valley systems are interpreted within the Lower Ness Formation (SB350, SB400; Fig. 10), but it is difcult to reconstruct their palaeogeography. In both cases, west east-trending valleys are documented in the Tern Field, on the Tern Eider Horst (Jennette & Riley, 1996), and east of the Ninian Hutton Dunlin fault system (e.g. Livera, 1989; Fig. 13F). It appears likely that the dominant sediment transport routes were along the northern margin of the Tern Eider Horst and from southwest of the hangingwall of the Ninian Hutton Dunlin fault system (Fig. 13F). Southward-retreating transgressive shorelines (e.g. FS300 in Fig. 13E, FS350 in Fig. 13G) may result from the abandonment of the northerly sediment transport route, whereas westward-retreating transgressive shorelines (e.g. FS400 in Fig. 13H) may reect abandonment of both sediment transport routes. Differential subsidence across the Ninian Hutton Dunlin fault system appears to have no signicant inuence on any of the palaeogeographies described above (Fig. 13E H). 5.4. FS400 FS500 (Middle Ness Formation) The Middle Ness Formation is bounded at its base by the base-Mid-Ness Shale ooding surface (FS400; Fig. 10) and at its top by a prominent ooding surface above the MidNess Shale (FS500; Fig. 10). This interval has a relatively uniform thickness across the area, with minor thickening in the hangingwall of the Ninian Hutton Dunlin fault system and in the area of the Cormorant Block IV Field, along the southern margin of the Tern Eider Horst (Fig. 12D). Flooding surface FS500 is marked by the southwestward retreat of a northwest southeast-trending shoreline (Fig. 13I) that comprised a wave-dominated shoreface and barrier system (facies 2.1 2.6; Table 2). One incised valley system is interpreted (SB500; Fig. 10), trending west east across the southern part of the study area (Fig. 11E and F). As in the Lower Ness Formation, palaeogeographies were not inuenced by differential subsidence across the Ninian Hutton Dunlin fault system (Fig. 13H and I). 5.5. FS500 SB1000 (Upper Ness Formation) The Upper Ness Formation is bounded at its base by a prominent ooding surface above the Mid-Ness Shale (FS500; Fig. 10) and at its top by a major sequence boundary at, or near to, the base of the Tarbert Formation (SB1000; Fig. 10). This interval is strongly truncated by erosion at the overlying sequence boundary (SB1000), particularly in the southwestern part of the study area, but it thickens considerably in the hangingwall of the Ninian Hutton Dunlin fault system (Fig. 12E). In addition, erosion associated with an intra-Upper Ness surface (FS850) is common in the northern part of the study area, although the depth of erosion is highly variable (Fig. 12E). The lower part of the Upper Ness Formation (FS500 FS800; Fig. 10) is dominated by aggradational oodplain

478

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

and uvial channel-ll deposits (facies 3.3,3.4; Table 2) in the southern part of the study area. Consequently, we interpret these strata to be the most proximal part of the Brent Group succession. They contain a high abundance of uvial channel-ll sandstones (facies 3.3,3.4; Table 2) that can be subdivided into three stratigraphically discrete, laterally extensive, multistorey channel-belt deposits in the hangingwall of the Ninian Hutton Dunlin fault system. Each of these channel-belt deposits is tentatively interpreted to overlie a sequence boundary (SB550, SB600, SB700; Figs. 10 and 11E). It is not clear from our correlations whether these channel belts are conned to incised valleys, although each appears to locally erode out lagoonal mudstones and minor ooding surfaces (Fig. 11E). Each channel belt trends west east (e.g. Fig. 13J and K), implying that sediment was dominantly transported from the southwest of the study area and was not inuenced by differential subsidence across the Ninian Hutton Dunlin fault system. The upper part of the Upper Ness Formation (FS800 SB1000; Fig. 10) is dominated by lagoonal deposits (facies 3.1,3.2; Table 2) in the southern part of the study area and a southward-retreating, wave-dominated shoreface and barrier system (facies 2.1 2.6; Table 2) in the northern part of the study area. Flooding surface FS800 is associated with the development of a regionally extensive lagoon overlying a widespread coal (Figs. 11 and 13L). This coal seam is taken as the base of the Tarbert Formation in the Norwegian sector of the East Shetland Basin (Tampen Spur in Fig. 2; Davies et al., 2000; McLeod et al., 2000). The overlying ooding surfaces FS850 and FS900 are both marked by the southward retreat of a west east-trending, wave-dominated shoreline (Fig. 13M and N). Transgressive erosion is associated with each ooding surface, both of which overlie deep (up to 15 m), broad (up to 5 km) scours inlled by estuarine sandstones (facies 2.4; Table 2; e.g. below FS850 in Fig. 11D). It is not clear if these scours were cut by tidal processes during transgression or whether they represent uvial valleys incised prior to transgression. We favour the former interpretation, because there is no evidence of a uvial lag at the base of the scours (e.g. Fig. 9E). The wavedominated shoreline built out northwards after each transgression, resulting in deposition of a regressive, upward-shallowing, wave-dominated shoreface succession (facies 2.1 2.4; Table 2; e.g. above FS900 in wells 211/29BC20, 211/29-BC6 and 211/29-BD25, Fig. 11E). None of the gross palaeogeographies described above appear to have been inuenced by differential subsidence across the Ninian Hutton Dunlin fault system (Fig. 13L N). However, there is evidence for locally variable erosion and/or non-deposition at ooding surface FS850 and sequence boundary SB1000 (Figs. 12E, 13L N). Details of this erosion cannot be resolved using the widely spaced wells in the study dataset, but it appears to be associated with local angular stratigraphic relationships (e.g. between wells 211/ 16-6, 211/18-19 and 211/19-6 in Fig. 11D, and between

wells 211/23-2, 211/23-DA27, 211/19-1, 211/19-5, 211/193 and 211/19-6 in Fig. 11F). Using 3D seismic data, Davies et al. (2000) and McLeod et al. (2000) have interpreted similar stratigraphic relationships in the Upper Brent Group (above ooding surface FS800) as the result of small depocentres and intra-basinal highs created by the initiation of major rift fault arrays (e.g. the North Alwyn Brent Statfjord fault system in Fig. 2; McLeod et al., 2000). Our interpretations of local angular stratigraphic relationships are consistent with this model. 5.6. SB1000 FS1200 (Tarbert Formation) The base of the Tarbert Formation is interpreted as a major sequence boundary (SB1000; Figs. 10 and 11) marked by an abrupt inux of coarse-grained, extrabasinal material (Mitchener et al., 1992; Morton, 1992) and a signicant time gap (. 1 Ma). There is also an abrupt change in facies character across the surface, from lagoonal and wave-dominated shoreface and barrier deposits (facies 2.1 2.6; Table 2) to tide-inuenced channel-ll and sheet sandstones (facies 1.2 1.4; Table 2). The top of the formation is dened by the transition into overlying offshore Heather Shales (Fig. 10). This transition is diachronous, and within the study area the top of the Tarbert Formation coincides locally with four discrete ooding surfaces (FS1050, FS1100, FS1150, FS1200; Figs. 10 and 11). The base-Tarbert sequence boundary (SB1000; Figs. 10 and 11) is a subtle, regional angular unconformity that truncates progressively older strata towards the south and west of the study area (Figs. 12C E and 13F N). Enhanced erosion and/or non-deposition at this surface also occurred locally and may have been the result of actively growing, small-scale structures. For example, deep local erosion at this surface in well 3/10-1 (Fig. 11A) may be the result of uplift on the crest of a growth monocline above a blind extensional fault (the southern segment of the North Alwyn Brent Statfjord fault system; Fig. 2B), which later broke surface and resulted in deposition of an expanded Tarbert section (Figs. 11A and 12E F). However, basin margin uplift is the key regional characteristic of the SB1000 unconformity, and can perhaps be linked to the inux of large volumes of extrabasinal material above this surface. Within the Brent Group, only the base-Broom sequence boundary (SB100) displays similar morphological and sediment provenance characteristics. The Tarbert Formation has a complex internal stratigraphy, but its lower part generally consists of an erosively based blocky sandstone, interpreted as a tide-inuenced channel ll (facies 1.4; Table 2), or a correlative coarsegrained lag (Fig. 6A C). The channel-ll sandstone or lag is generally overlain by a series of retrogradationally stacked, thick (2 10 m) upward-coarsening, tide-inuenced sheet sandstone successions (facies 1.2 1.3; Table 2) and/or a thin (, 5 m), upward-ning, tide-inuenced sheet sandstone succession (facies 1.2 1.3; Table 2; e.g. wells 2/5-17 and

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

479

211/19-6 in Fig. 7B and D). Upward-coarsening successions are thicker and more common in the south-western part of the study area (e.g. wells 2/5-17, 2/5-3, 3/1-1, 3/2-3 and 3/24 in Figs. 7B and 11A) and in the immediate hangingwall of the Ninian Hutton Dunlin fault system (e.g. well 3/3-8 in Fig. 11A, well 211/18-5 in Fig. 11B, and well 211/24-5 in Fig. 11C), as reected in isopach maps of the lower part of the Tarbert Formation (Fig. 12F). These retrogradationally stacked successions record net transgression of the Tarbert depositional system, culminating in ooding surface FS1050 and the deposition of offshore shales over the western and northern part of the study area (Fig. 11). Their distribution and thickness variations imply that sediment was transported from the south west of the study area and along the southern part of the Ninian Hutton Dunlin fault hangingwall (Figs. 12F and 13O). Isopach maps and palaeogeographic reconstructions suggest that there was a third sediment transport route along the northern margin of the Tern Eider Horst (Fig. 13O) and a local depocentre in the area of the Cormorant Block IV Field, along the southern margin of the Tern Eider Horst (e.g. wells 211/215 in Figs. 11C and 12F). In the hangingwall of the southern part of the Ninian Hutton Dunlin fault system (e.g. wells 3/3-8, 3/4-8, 3/9-4, 3/10-1, 3/4-12 and 3/4-9 in Fig. 11A and E), ooding surface FS1050 is overlain by a thick (10 20 m) upwardcoarsening, tide-inuenced sheet sandstone succession (facies 1.2 1.3; Table 2) that is truncated by a tideinuenced channel ll (facies 1.4; Table 2). This facies architecture is interpreted to record shoreline regression, with the channel-ll sandstones overlying an intra-Tarbert sequence boundary (SB1100; Figs. 10 and 11). The limited distribution of these deposits implies that deposition was conned to the hangingwall of the Ninian Hutton Dunlin fault system (Fig. 13P), which at this time must been a major topographic feature. This interpretation is supported by the observation that coarse-grained sediment is largely conned to the immediate hangingwall of the Ninian Hutton Dunlin fault system (e.g. well 3/3-8 in Fig. 11A). Transgressive deposition above sequence boundary SB1100 culminated in ooding surface FS1150 and the deposition of offshore shales over much of the hangingwall area of the Ninian Hutton Dunlin fault system (e.g. wells 3/4-8, 3/9-4, 3/4-12 and 3/4-9 in Fig. 11A and E). We tentatively interpret a second intra-Tarbert sequence boundary in wells 3/3-8 and 3/10-1 (SB1200; Figs. 10 and 11A). Strata associated with this sequence boundary may record sediment transport and deposition along the hangingwalls of the Ninian Hutton Dunlin fault system and the evolving North Alwyn Brent Statfjord fault system. New biostratigraphic data demonstrate that oolitic ironstones developed at two discrete stratigraphic levels in the Heather Shales above the Tern Eider Horst and other intra-basinal highs are coeval with sequence boundary SB1100 and ooding surface FS1100 (e.g. wells 210/25-2, 211/21-5, 211/21-6 in Figs. 11C and 13P) and with sequence

boundary SB1200 and ooding surface FS1200 (e.g. well 210/20-1, 210/20-2 and 211/16-6 in Fig. 11D). Such ironstones require extended physical reworking in areas of clastic sediment starvation (Young, 1989) and have been documented at similar sediment-starved sequence boundaries that underwent subsequent transgressive reworking (e.g. Taylor, Simo, Yokum, & Leckie, 2002).

6. Discussion: the added value of an integrated regionalto reservoir-scale approach The high-resolution sequence stratigraphic framework summarised above integrates core-scale sedimentology, reservoir-scale facies architecture and regional stratigraphy. The extensive use of core data is important, because it allows sedimentological and sequence stratigraphic interpretations to be constructed from rst principles. The integrated regional- to reservoir-scale approach aids identication of subtle intra- and inter-reservoir features, which are not evident via the detailed study of individual reservoirs in isolation. This approach is particularly valuable in a mature hydrocarbon province, such as the UK Brent Province, because it generates new insights, concepts and models that will likely contribute to improved ultimate recovery and the identication of additional exploration potential. We highlight several key insights below. 6.1. Improved denition of depositional trends The framework outlined here resolves temporal and spatial trends in Brent Group deposition across the entire UK Brent Province at a very high (i.e. reservoir-scale) resolution. Our core-based sequence stratigraphic approach subdivides the Brent Group into genetically related stratal packages within which lateral facies relationships and sandbody distribution can be robustly elucidated (e.g. Figs. 10, 11 and 13). The stratal packages identied by the framework are dened by widespread facies discontinuities, and consequently they correspond closely to the existing reservoir zonation in many elds. By placing these reservoir zones within a regional context, it is possible to rene intra-zone facies trends both within and between elds. The framework also provides a context in which reservoir zonation may be tested and rened, particularly for those reservoirs in the early stages of development. 6.2. Predictive depositional models of the Broom and Tarbert Formations Current sedimentological models of Brent Group reservoirs focus primarily on the Rannoch, Etive and Ness Formations, and emphasise the genetic links between the deposits of these formations (e.g. Budding & Inglin, 1981; Fig. 5). Facies types and stratigraphic architecture in

480

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

the Broom and Tarbert Formations are not explained by these models. Instead, these two reservoir intervals have been described only within individual elds, resulting in apparently contrasting interpretations that have limited predictive value outside of a specic reservoir. For example, the Tarbert Formation has been variously interpreted as a tidal valley-ll sandbody (Jennette & Riley, 1996 in the Eider Field; Flint et al., 1998 in the Northwest Hutton Field), a series of retrogradationally stacked barrier sandstones and lagoonal mudstones (Rnning & Steel, 1987 in the North Alwyn, Alwyn and Hild Fields) that locally contains a valley-ll sandbody tevik, Jakobsen, & Helland-Hansen, 1999 in (Bruaset, Ba the Gullfaks Field; Davies et al., 2000 in the Snorre and Tordis Fields), and as a complex series of wave- and tidedominated sandstones (Reynolds, 1995 in the Thistle Field). Although each of these interpretations is essentially correct for a specic reservoir, our work indicates that they apply to depositional systems developed in different stratigraphic intervals. For example, the Tarbert Formation comprises an estuarine channel-ll sandbody developed within a wave-dominated barrier-shoreface system and deposited below ooding surface FS850 in the Eider Field (e.g. 89800 90230 in well 211/16-6, Figs. 7C and 11D), and a series of stacked tide-inuenced sheet sandstones developed within a weakly wave-inuenced shallowmarine system and deposited above sequence boundary SB1000 in the Strathspey Field (e.g. 94470 95420 in well 3/4-12, Figs. 7A and 11E). A model developed for either reservoir is wholly inappropriate for the other. Indeed, neither model can be robustly exported to the other Brent Group reservoirs outside the context of a regional sequence stratigraphic framework. Our interpretations present regionally consistent sedimentological and sequence stratigraphic models of the Broom and Tarbert Formations that allow robust prediction of facies architecture and reservoir character between elds. Both of the stratigraphic intervals SB100 FS200 (corresponding to the Broom Formation; Fig. 10) and SB1000 FS1200 (corresponding regionally to the Tarbert Formation; Fig. 10) comprise net-transgressive, tide-inuenced deposits that are genetically unrelated to the Rannoch, Eive and Ness Formations. Both intervals have sequence boundaries at their bases, characterised by an inux of coarse-grained extrabasinal sediment and a signicant time gap (. 1 Ma). If present, the lowstand deposits that lie at the down-dip termination of the sequence boundaries occur beyond the northern and eastern limits of the study area, but they may include detached shoreline deposits (e.g. Fig. 13A). The Broom Formation (SB100 FS200; Fig. 10) comprises a regionally extensive, tide-inuenced channelll sandstone complex at its base overlain by retrogradationally stacked, upward-coarsening tide-inuenced sheet sandstones (Figs. 11A and 13A and B). The lower part of the Tarbert Formation (SB1000 FS1050; Fig. 10) has a similar facies architecture (Figs. 11A and 13O), but

the locally restricted, upper part of the formation (FS1050 FS1200; Fig. 10) comprises two cycles of regression and subsequent transgression of the same depositional system (Figs. 11A, E and 13P). Retrogradationally stacked, wavedominated barrier and lagoonal deposits that are genetically afliated to the Ness Formation (FS800 SB1000; Fig. 10) and that underlie the regional base-Tarbert sequence boundary (SB1000; Fig. 10) are allocated to the Tarbert Formation in some elds on the basis of lithostratigraphy. Such a lithostratigraphic approach to regional correlation produces erroneous palaeogeographic reconstructions that are difcult to reconcile with current sedimentological models of wave- and tide-inuenced depositional systems. 6.3. Tectono-stratigraphic controls on reservoir architecture The tectono-stratigraphic evolution of the Brent Province exerts a signicant, but underappreciated, inuence on reservoir architecture. We interpret a threefold hierarchy of tectono-stratigraphic controls, outlined below. At the rst-order scale, the interplay between two variables exerted a fundamental control on palaeo-geomorphology, sediment transport routes and sedimentary process regimes; (1) net sediment-supply rate, and (2) the rate at which accommodation space was generated across key structures (Fig. 2B) by differential tectonic subsidence. Deposition was characterised by the net-transgessive Broom and Tarbert Formations (SB100 FS200 and SB1000 FS1200, respectively; Fig. 10) at times when sediment supply was low relative to differential accommodation generation. These units appear to comprise three palaeogeographically distinct depositional systems: (1) one system was conned to the hangingwall of the Ninian Hutton Dunlin fault system, (2) the second built out from the southwestern corner of the study area, near the Heather Field, and (3) the third built out along the northern margin of the Tern Eider Horst (Figs. 11 and 13A, B, O and P). Thus, palaeogeography and sediment transport routes strongly reect differential subsidence across the Ninian Hutton Dunlin fault system, with depositional systems oriented from south (proximal) to north (distal) in its hangingwall (Fig. 13A, B, O, and P), and the Tern Eider Horst basement structure (Fig. 12A and F). We infer that both the Ninian Hutton Dunlin fault system and the Tern Eider Horst had a strong geomorphological expression during times of low net sediment supply, and thus controlled palaeogeography and sediment transport. In addition, the sedimentology of the Broom and Tarbert Formations is characterised by the absence of wave processes and the presence of relatively strong tidal inuence. We attribute this process regime to the strong surface expression of the Ninian Hutton Dunlin fault system and the Tern Eider Horst, which may have compartmentalised the UK Brent Province into a series of conned basins, each with small wave fetch and, possibly, amplied tidal currents. In contrast, deposition at times of high sediment supply rate,

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

481

relative to differential accommodation generation, was characterised by the regressive to aggradational Rannoch, Etive and Ness Formations (FS200 SB1000; Fig. 10). Despite signicant differential subsidence across the Ninian Hutton Dunlin fault system and the Tern Eider Horst (Fig. 12B E), these features did not strongly inuence palaeogeographic trends and sediment supply routes during these times (Fig. 13C N). Thus, we infer that sediment supply was sufcient to continuously inll the differential accommodation created across these structures, so that they had no surface, geomorphological expression. The wavedominated character of Rannoch Etive shoreline systems reects deposition in an unconned, open basin with large wave fetch. This facies character is also consistent with the subdued surface expression of underlying structural features. The thick (. 100 m), widespread character of the Rannoch Etive Ness succession (FS200 SB1000) implies that regional tectonic subsidence across the entire Brent Province was more rapid than during deposition of the underlying Broom Formation (SB100 FS200), which represents a similar timespan (6 8 Ma, Fig. 10). At the second-order scale, the key sequence stratigraphic surfaces identied in our study (Fig. 10) dene unconformity-bounded sequences that extend across the UK Brent Province (Figs. 11 13). With the exception of the base-Broom and base-Tarbert unconformities (SB100 and SB1000, respectively; Fig. 10), sequence boundaries are not characterised by regional angular stratal relationships. This observation, combined with the regional extent of the sequences and their durations of between ca. 1.0 Ma (Rannoch Etive Ness interval, FS200 SB1000; Fig. 10) and ca. 3.0 Ma (Tarbert Formation, SB1000 FS1200; Fig. 10) implies that they were produced by a regional, and possibly eustatic, control on relative sea-level. The base-Broom and base-Tarbert sequence boundaries (SB100 and SB1000, respectively; Fig. 10) are both associated with regional angular truncation of underlying strata and are thus interpreted to be of a different origin. Underhill and Partington (1993) document angular truncation at the baseBroom sequence boundary (SB100) increasing towards the central North Sea triple junction, which lay approximately 300 km south of the study area, and interpret the unconformity to have been produced by thermal doming near the triple junction. The base-Tarbert sequence boundary has generally been attributed to the onset of late Jurassic rifting (e.g. Mitchener et al., 1992), but many of the late Jurassic rift fault systems were poorly developed at this time (e.g. Fig. 2; Dawers & Underhill, 2000; McLeod et al., 2000). Our work suggests that the sequence boundary is instead associated with regional angular truncation that increases towards the southwest of the study area, implying uplift here (Fig. 12D and E). While this uplift may have been due to the early development of late Jurassic rift faults near the western margin of the East Shetland Basin, it can also be attributed to other mechanisms. For example, the large time gap across

the base-Tarbert sequence boundary (up to 5 Ma; Fig. 10), its angular character and the inux of extrabasinal material above it are consistent with a phase of renewed thermal doming. Given the limited current knowledge of the regional distribution of this surface outside our study area, we regard its origin as enigmatic. At the third-order scale, we interpret a number of small (, 10 km wide), short-lived and localised depocentres. A series of such depocentres formed along the eastern margin of the Tern Eider Horst at various stages of Brent Group deposition (Fig. 12A, D and F). We tentatively interpret these depocentres to have resulted from episodic movement of the basement-involved Tern Eider Horst, perhaps in response to intra-plate stresses associated with thermal doming and relaxation near the central North Sea triple junction. The depocentres may have been fault bounded. We also interpret several local angular stratigraphic relationships in the upper Brent group (FS850 FS1200; Fig. 10; e.g. between wells 211/16-6, 211/18-19 and 211/19-6 in Fig. 11D, and between wells 211/23-2, 211/23-DA27, 211/19-1, 211/19-5, 211/19-3 and 211/19-6 in Fig. 11F) that are consistent with models of small (, 5 km wide) depocentres and highs developed in response to rift initiation (Davies et al., 2000; McLeod et al., 2000). Individual depocentres suggested by these models cannot be resolved in detail using the widely spaced wells in our study dataset, but our regional correlations suggest that the depocentre-ll successions do not comprise locally restricted facies, but instead regionally extensive facies units (e.g. wave-dominated shoreface and barrier successions in the uppermost Ness Formation, FS850 SB1000, in potential depocentres penetrated by wells 211/18-19, 211/19-1, 211/19-5 and 211/19-3 in Fig. 11D and E). 6.4. Applications to near-eld exploration potential All currently producing Brent Group reservoirs are dened by structural traps. Using the insights described above, we identify three potential stratigraphic trapping mechanisms that may offer near-eld exploration potential if combined with further studies. (1) Firstly, the base-Broom and base-Tarbert sequence boundaries (SB100 and SB1000, respectively; Fig. 10) may be associated with lowstand shoreline and/or deep-water deposits at their seaward limit, beyond the northern and eastern limits of the study area (e.g. Fig. 13A). This stratigraphic trapping mechanism is speculative, because the seaward pinch-out of the Broom and Tarbert depositional systems is poorly constrained by our interpretations. Also, any such lowstand shoreline deposits may be very deeply buried (. 5000 m) in the centre of the North Viking Graben or in the Magnus Basin, thus limiting their reservoir potential. (2) The seaward pinch-out of the Rannoch and Etive Formations (FS200 SB1000; Fig. 10) is poorly dened and may provide trap potential in combination with post-depositional structural dips. Again, the palaeogeographic orientation and extent of

482

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

this pinchout is poorly constrained by our interpretations. Delineation of these trends is likely to require detailed stratigraphic interpretation and/or attribute analysis of 3D seismic data volumes north of the Tern Eider Horst (Figs. 2 and 4). (3) Localised depocentres lled by the shallowmarine sandstones of the uppermost Ness and Tarbert Formations (FS850 FS1200; Fig. 10) may provide exploration targets. The location of such depocentres is not predictable from late Jurassic structure maps (e.g. Fig. 2C), but will require detailed and careful tectono-stratigraphic analysis of 3D seismic data volumes (e.g. Davies et al., 2000; McLeod et al., 2000). 6.5. Applications to improved in-eld recovery The insights discussed above also have applications to improved understanding of facies architecture in producing reservoirs, and thus, via input to reservoir models, to predicting the distribution of remaining oil in place. We highlight three approaches through which reservoir facies architecture may be improved. (1) The framework discussed above may provide a context for improved temporal and spatial resolution of depositional trends within some reservoirs, thus leading to renement of reservoir zonation and intra-zone facies trends. For example, regional and subregional facies trends may be extrapolated into reservoir zones to constrain intra-zone trends, particularly where data density is low. This approach may be used to increase the deterministic geological knowledge available during reservoir model construction, and thus reduce the range of uncertainty in model realisations. (2) The framework highlights similarities and differences in reservoir zonation and facies architecture between various elds, and may serve as a basis for exporting well-constrained subsurface datasets from one Brent Group reservoir to another. For example, differential subsidence across the Ninian Hutton Dunlin fault system may have resulted in different channel-ll stacking patterns in the Ness Formation across the fault, such that it may not be appropriate to export wellconstrained datasets on channel-ll connectivity from the hangingwall (e.g. Brent Field) to the footwall (e.g. Ninian Field). Regional interpretations can thus be used as a lter for the application of well-constrained analogue datasets from the same depositional system in reservoir modelling. (3) The framework highlights reservoirs and reservoir zones where subtle, and previously unconsidered, angular stratigraphic relationships may occur. For example, intrareservoir thinning of the Brent Group onto the crest of the Ninian Hutton Dunlin fault system may be accompanied by subtle angular discordances, onlap and/or truncation at key sequence stratigraphic surfaces (e.g. between wells 211/23-DA27, 211/23-DA36 and 211/23-DA12 in the Dunlin Field, Fig. 11D). Subtle angular stratigraphic relationships may also be present within the uppermost Ness and Tarbert Formations (FS850 FS1200), as a consequence of local depocentre generation during rift

initiation. These relationships may be tested using 3D seismic data, where sufcient vertical resolution is possible, and focused biostratigraphic analysis.

7. Conclusions Using an extensive core and wireline-log dataset, integrated with palynostratigraphy and published literature, we have constructed a high-resolution sequence stratigraphic framework for the UK Brent Province. This framework allows temporal and spatial trends in regional deposition to be interpreted at a higher resolution than previously possible. The resulting high-resolution interpretations of reservoir distribution are consistent both within and between established elds. We interpret a hierarchy of basinwide, unconformitybounded sequences within the Brent Group. Regionally extensive, low-frequency sequence boundaries occur at the base of the Broom Formation and at, or near to, the base of the Tarbert Formation. Both sequence boundaries are associated with regional angular truncation, signicant missing time (up to ca. 5 Ma) and an inux of extrabasinal material, and they are interpreted to be tectonically driven or enhanced. Higher frequency sequences of basinwide extent occur within the Rannoch, Etive, Ness and Tarbert Formations. Stratigraphic architecture, sediment dispersal patterns and sedimentary processes within different intervals of Brent Group stratigraphy are all governed by the interplay between sediment supply and tectonically generated accommodation space. During deposition of lowfrequency transgressive systems tracts (constituting the Broom and Tarbert Formations), differential subsidence across the active Ninian Hutton Dunlin fault system and the basement Tern Eider Horst trend outpaced sediment supply. Consequently, these structural trends accumulated signicant geomorphological expression and were the dominant control on palaeogeography and sediment transport routes. These structural features also compartmentalised the Brent Province, thus suppressing wave processes and enhancing tides during deposition of the Broom and Tarbert Formations. During deposition of a low-frequency highstand systems tract (constituting the Rannoch, Etive and Ness Formations), sediment supply outpaced differential tectonic subsidence, such that active faults and basement structural trends had no geomorphological expression, despite signicant thickness changes across them. These structural features therefore had little inuence on palaeogeography and sediment transport routes, while the wave-dominated character of shallowmarine deposition at these times testies to the open, noncompartmentalised character of the Brent Province. At a smaller scale, differential subsidence across the active Ninian Hutton Dunlin fault and various small, rift initiation faults is associated with pronounced thickness

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484

483

changes and angular stratal relationships within several higher frequency sequences. The insights gained from an integrated regional- to reservoir-scale approach may contribute to the identication of near-eld exploration potential and to improved in-eld recovery. The former is achieved via the recognition of various stratigraphic trapping mechanisms. The latter involves using the regional sedimentological and tectonostratigraphic context to constrain intra-reservoir depositional trends and the choice of analogue datasets in reservoir model construction, therefore reducing uncertainty in reservoir characterisation.

Acknowledgements This work has been funded by Shell UK Expro, and has beneted from numerous discussions with Shell and ExxonMobil geoscientists, in particular Frances Abbotts, Janet Almond, Duncan Erratt, Nick Hogg, Winfrield Leopoldt, John Marshall, David Taylor and Steve Taylor. Additional discussions with Huw Williams, Paul Davies (Reservoir Geology Consultants), Aileen McLeod, Mark Tomasso, John Underhill (University of Edinburgh), Sarah Davies (University of Leicester), John Howell (University of Bergen) and Steve Flint (University of Liverpool) were fruitful. We thank two un-named reviewers for their constructive comments.

References
Besly, B. M., & Fielding, C. R. (1989). Palaeosols in Westphalian coalbearing and red bed sequences, central and northern England. Palaeogeography, Palaeoclimatology and Palaeoecology , 70, 303330. Brown, S., & Richards, P. C. (1989). Facies and development of the Middle Jurassic Brent Delta near the northern limit of its progradation, UK North Sea. In M. K. G. Whateley, & K. T. Pickering (Eds.), Deltas: Sites and traps for fossil fuels (pp. 253 268). Geological Society of London, Special Publication 41. Brown, S., Richards, P. C., & Thomson, A. R. (1987). Patterns in the deposition of the Brent Group (Middle Jurassic), UK North Sea. In J. Brooks, & K. Glennie (Eds.), Petroleum geology of Northwest Europe (pp. 899913). London: Graham and Trotman. tevik, A., Jakobsen, K. G., & Helland-Hansen, W. (1999). Bruaset, V., Ba An anomalous thick, elongated sandbody, Tarbert Formation, Gullfaks Field: High accommodation uvial channel stacking, incised valley deposition or both? (abstract). Palaeozoic to Recent sedimentary environments, offshore Norway, Oslo: Norwegian Petroleum Directorate (NPF), pp. 101 104. Bryant, I. D., & Livera, S. E. (1991). Identication of unswept oil volumes in a mature eld by using integrated data analysis: Ness Formation, Brent Field, UK North Sea. In A. M. Spencer (Ed.), Generation, accumulation and production of Europes hydrocarbons (pp. 75 88). European Association of Petroleum Geologists, Special Publication 1. Budding, M. C., & Inglin, H. F. (1981). A reservoir geological model of the Brent Sands in southern Cormorant. In V. Illing, & G. D. Hobson (Eds.), Petroleum geology of the continental shelf of north-west Europe (pp. 326334). London: Institute of Petroleum.

Davies, S. J., Dawers, N. H., McLeod, A. E., & Underhill, J. R. (2000). The structural and sedimentological evolution of early syn-rift successions: The Middle Jurassic Tarbert Formation, North Sea. Basin Research, 12, 343 365. Davies, R. J., Turner, J. D., & Underhill, J. R. (2001). Sequential dip-slip fault movement during rifting: A new model for the evolution of the Jurassic trilete North Sea rift system. Petroleum Geoscience, 7, 371 388. Dawers, N. H., & Underhill, J. R. (2000). The role of fault interaction and linkage in controlling synrift stratigraphic sequences: Late Jurassic, Statfjord East area, northern North Sea. American Association of Petroleum Geologists Bulletin, 84, 4564. Deegan, C. E., & Scull, B. J. (1977). A standard lithostratigraphic nomenclature for the central and northern North Sea. Report of the Institute of Geological Sciences, 77/25. Frseth, R. B. (1996). Interaction of Permo-Triassic and Jurassic extensional fault-blocks during the development of the northern North Sea. Journal of the Geological Society of London, 153, 931944. Fjellganger, E., Olsen, T. R., & Rubino, J. L. (1996). Sequence stratigraphy and palaeogeography of the Middle Jurassic Brent and Vestland deltaic systems, northern North Sea. Norsk Geologisk Tidsskrift, 76, 75106. Flint, S. S., Knight, S., & Tilbrook, A. (1998). Application of highresolution sequence stratigraphy to the Northwest Hutton Field, northern North Sea: Implications for management of a mature Brent Group eld. Bulletin of the American Association of Petroleum Geologists, 82, 14161436. Goldring, R., Pollard, J. E., & Taylor, A. M. (1991). Anconichnus horizontalis: A pervasive ichnofabric-forming trace fossil in postPalaeozoic offshore siliciclastic facies. Palaios, 6, 250263. Graue, E., Helland-Hansen, W., Johnsen, J., Lmo, L., Nttvedt, A., Rnning, K., Ryseth, A., & Steel, R. J. (1987). Advance and retreat of the Brent Delta System, Norwegian North Sea. In J. Brooks, & K. Glennie (Eds.), Petroleum geology of northwest Europe (pp. 915 937). London: Graham and Trotman. Hamilton, P. J., Fallick, A. E., & MacIntyre, R. M. (1987). Isotopic tracing of the provenance and diagenesis of Lower Brent Group sands, North Sea. In J. Brooks, & K. Glennie (Eds.), Petroleum geology of northwest Europe (pp. 939949). London: Graham and Trotman. Hampson, G. J. (2000). Discontinuity surfaces, clinoforms, and facies architecture in a wave-dominated, shoreface-shelf parasequence. Journal of Sedimentary Research, 70, 325340. Hampson, G. J., Elliott, T., Davies, S. J., & Flint, S. S. (1997). The application of sequence stratigraphy to Upper Carboniferous uviodeltaic strata of the onshore UK and Ireland: Implications for the southern North Sea. Journal of the Geological Society of London, 154, 719 733. Helland-Hansen, W., Badley, M. E., Lmo, L., & Steel, R. J. (1992). Advance and retreat of the Brent Delta: Recent contributions to the depositional model. In A. C. Morton, R. S. Haszeldine, M. R. Giles, & S. Brown (Eds.), Geology of the Brent Group (pp. 109 127). Geological Society of London, Special Publication 61. Howell, J. A., Flint, S. S., & Hunt, C. (1996). Sedimentological aspects of the Humber Group (Upper Jurassic) of the southern Central Graben, UK North Sea. Sedimentology, 43, 89 114. Jennette, D. C., & Riley, C. O. (1996). Inuence of relative sea-level on facies and reservoir geometry of the Middle Jurassic lower Brent Group, UK North Viking Graben. In J. A. Howell, & J. F. Aitken (Eds.), High resolution sequence stratigraphy: Innovations and applications (pp. 87113). Geological Society of London, Special Publication 104. Johnson, R. J., & Dingwall, R. G. (1981). The Caledonides: Their inuence on the stratigraphy of the North-West European Continental Shelf. In V. Illing, & G. D. Hobson (Eds.), Petroleum geology of the continental shelf of north-west Europe (pp. 8598). London: Institute of Petroleum. Johnson, H. D., & Stewart, D. J. (1985). The role of clastic sedimentology in the exploration and production of oil and gas in the North Sea. In P. J. Brenchley, & B. P. J Williams (Eds.), Sedimentology: Recent

484

G.J. Hampson et al. / Marine and Petroleum Geology 21 (2004) 457484 Retallack, G. J. (1990). Soils of the past. London: Unwin Hyman, p. 520. Reynolds, A. D. (1995). Sedimentology and sequence stratigraphy of the Thistle Field, northern North Sea. In R. J. Steel, V. L. Felt, J. P. Johannessen, & C. Mathieu (Eds.), Sequence stratigraphy on the Northwest European margin (pp. 257271). Norwegian Petroleum Directorate (NPF), Special Publication 5. Richards, P. C., & Brown, S. (1986). Shoreface storm deposits in the Rannoch Formation (Middle Jurassic), North West Hutton oileld. Scottish Journal of Geology, 22, 367375. Roberts, A. M., Kusznir, N. J., Walker, I. M., & Dorn-Lopez, D. (1995). Quantitative analysis of Triassic extension in the northern Viking Graben. Journal of the Geological Society of London, 152, 1526. Rnning, K., & Steel, R. J. (1987). Depositional sequences within a transgressive Reservoir sandstone unit: The Middle Jurassic Tarbert Formation, Hild Area, northern North Sea. In J. Kleppe, E. W. Berg, A. T. Buller, O. Hjelmeland, & O. Torstaeter (Eds.), North Sea oil and gas reservoirs (pp. 169 176). London: Graham and Trotman. Sawyer, M. J., & Keegan, J. B. (1996). Use of palynofacies characterization in sand-dominated sequences, Brent Group, Ninian Field, UK North Sea. Petroleum Geoscience, 2, 289 297. Scott, E. S. (1992). The palaeoenvironments and dynamics of the Rannoch Etive nearshore and coastal succession, Brent Group, northern North Sea. In A. C. Morton, R. S. Haszeldine, M. R. Giles, & S. Brown (Eds.), Geology of the Brent Group (pp. 129 147). Geological Society of London, Special Publication 61. Simon Petroleum Technology Ltd (1994). The Brent DeltaAn integrated sedimentological and palynological study of the Brent Group between latitudes 598N and 628N, UK and Norwegian sectors, North Sea. Unpublished report for Shell UK Expro. Taylor, A. M., & Gawthorpe, R. L. (1993). Application of sequence stratigraphy and trace fossil analysis to reservoir description: Examples from the Jurassic of the North Sea. In J. R. Parker (Ed.), (pp. 317336). Petroleum Geology of Northwest Europe: Proceedings of the Fourth Conference, Geological Society of London. Taylor, A. M., & Goldring, R. (1993). Description and analysis of bioturbation and ichnofabric. Journal of the Geological Society of London, 150, 141148. Taylor, K. G., Simo, T., Yokum, D., & Leckie, D. A. (2002). Stratigraphic Signicance of ooidal ironstones from the Cretaceous Western Interior Seaway: The Castlegate Sandstone, Utah and the Peace River Formation, Alberta. Journal of Sedimentary Research, 72, 316327. Underhill, J. R., & Partington, M. A. (1993). Jurassic thermal doming and deation in the North Sea: Implications of the sequence stratigraphic evidence. In J. R. Parker (Ed.), (pp. 337 345). Petroleum Geology of Northwest Europe: Proceedings of the Fourth Conference, Geological Society of London. Underhill, J. R., Sawyer, M. J., Hodgson, P., Shallcross, M. D., & Gawthorpe, R. L. (1997). Implications of fault scarp degradation for Brent Group prospectivity, Ninian Field, northern North Sea. Bulletin of the American Association of Petroleum Geologists, 81, 999 1022. Young, T. P. (1989). Phanerozoic ironstones: An introduction and review. In T. P. Young, & W. E. G. Taylor (Eds.), Phanerozoic ironstones (pp. ix xxv). Geological Society of London, Special Publication 46. Zaitlin, B. A., Dalrymple, R. W., & Boyd, R. (1994). The stratigraphic organization of incised-valley systems associated with relative sealevel change. In R. W. Dalrymple, R. Boyd, & B. A. Zaitlin (Eds.), Incised-valley systems: Origin and sedimentary sequences (pp. 4560). Society of Economic Paleontologists and Mineralogists, Special Publication 51.

developments and applied aspects (pp. 249 310). Geological Society of London, Special Publication 18. Lee, M. J., & Hwang, Y. J. (1993). Tectonic evolution and structural styles of the East Shetland Basin. In J. R. Parker (Ed.), (pp. 1137 1149). Petroleum Geology of Northwest Europe: Proceedings of the fourth conference. Livera, S. E. (1989). Facies associations and sand-body geometries in the Ness Formation of the Brent Group, Brent Field. In M. K. G. Whateley, & K. T. Pickering (Eds.), Deltas: Sites and traps for fossil fuels (pp. 269 286). Geological Society of London, Special Publication 41. Livera, S. E., & Caline, B. (1990). The sedimentology of the Brent Group in the Cormorant Block IV oileld. Journal of Petroleum Geology, 13, 367 396. MacEachern, J. A., Raychaudhuri, I., & Pemberton, S. G. (1992). Stratigraphic applications of the Glossifungites ichnofacies: Delineating discontinuities in the rock record. In S. G. Pemberton (Ed.), Applications of ichnology to petroleum exploration (pp. 169 198). Society of Economic Palaeontologists and Mineralogists, Core Workshop 17. McCarthy, P. J., Faccini, U. F., & Plint, A. G. (1999). Evolution of an ancient coastal plain: Palaeosols, interuves and alluvial architecture in a sequence stratigraphic framework, Cenomanian Dunvegan Formation, NE British Columbia, Canada. Sedimentology, 46, 861 891. McLeod, A. E., Dawers, N. H., & Underhill, J. R. (2000). The propogation and linkage of normal faults: insights from the StrathspeyBrent Statfjord fault array, northern North Sea. Basin Research, 12, 263284. Mearns, E. W. (1992). Samariumneodymium isotopic constraints on the provenance of the Brent Group. In A. C. Morton, R. S. Haszeldine, M. R. Giles, & S. Brown (Eds.), Geology of the Brent Group (pp. 213 225). Geological Society of London, Special Publication 61. Mitchener, B. C., Lawrence, D. A., Partington, M. A., Bowman, M. B. J., & Gluyas, J. (1992). Brent Group: Sequence stratigraphy and regional implications. In A. C. Morton, R. S. Haszeldine, M. R. Giles, & S. Brown (Eds.), Geology of the Brent Group (pp. 4580). Geological Society of London, Special Publication 61. Morris, J. E., Hampson, G. J., & Maxwell, G. (2003). Controls on facies architecture in the Brent Group, Strathspey Field, UK North Sea: Implications for reservoir characterization. Petroleum Geoscience, in press. Morton, A. C. (1992). Provenance of Brent Group sandstones: Heavy mineral constraints. In A. C. Morton, R. S. Haszeldine, M. R. Giles, & S. Brown (Eds.), Geology of the Brent Group (pp. 227 244). Geological Society of London, Special Publication 61. Olsen, T. R., & Steel, R. (2000). The signicance of the Etive Formation in the development of the Brent system: A review of the likelihood of forced regression during progradation. In D. Hunt, & R. L. Gawthorpe (Eds.), Sedimentary responses to forced regression (pp. 91 112). Geological Society of London, Special Publication 172. Plint, A. G. (1988). Sharp-based shoreface sequences and offshore bars in the Cardium formation of Alberta: Their relationship to relative changes in sea level. In C. K. Wilgus, B. S. Hastings, C. G. St. C. Kendall, H. W. Posamentier, C. A. Ross, & J. C. Van Wagoner (Eds.), Sea-level changesAn integrated approach (pp. 357 370). Society of Economic Paleontologists and Mineralogists, Special Publication 42. Rattey, R. P., & Heyward, A. B. (1993). Sequence stratigraphy of a failed rift system: The Middle Jurassic to Early Cretaceous basin evolution of the central and northern North Sea. In J. R. Parker (Ed.), (pp. 215 249). Petroleum Geology of Northwest Europe: Proceedings of the Fourth Conference.

Fig. 11. Correlation panels across the East Shetland Basin (see Fig. 4 for location): (A) WestEast panel through the Heather, Lyell, Ninian and North Alwyn Fields; (B) WestEast panel through the South Cormorant, Northwest Hutton, Hutton and Brent Fields; (C) West East panel through the Tern, North Cormorant, Dunlin and Statfjord Fields; (D) WestEast panel through the Eider, Thistle and Murchison Fields; (E) NorthSouth panel through the Statfjord, Brent, Strathspey and North Alwyn Fields; and (F) NorthSouth panel through the Murchison, Thistle, Dunlin, Hutton, Ninian and Columba Fields.

Fig. 11B (continued )

Fig. 11C (continued )

Fig. 11D (continued )

Fig. 11E (continued )

Fig. 11F (continued )

You might also like