Skip to Main Content
Skip Nav Destination
This is an Open Access article distributed under the terms of the Creative Commons Attribution 4.0 License (http://creativecommons.org/licenses/by/4.0/).

We conducted a high-resolution multi-disciplinary analysis of two core sections in the borehole Ellesmere Port-1, Cheshire, UK. Biostratigraphic analysis indicates that the core sections are Kinderscoutian and late Arnsbergian–Chokierian in age, respectively. Both cores are assigned to the Bowland Shale Formation (Holywell Shale). Coupled core scan and discrete geochemical analysis enables interpretation of syngenetic processes at a high stratigraphic resolution. Both cores exhibit the classic cyclicity of limestones, calcareous to non-calcareous mudstones and siltstones, interpreted to represent sediment deposition during fourth-order sea-level fluctuation. Machine learning of the well log data coupled to the core scan data enabled prediction of the key lithofacies through the entire Bowland Shale interval in Ellesmere Port-1. The machine predictions show that the Bowland Shale is interfingered with three turbiditic leaves of the Cefn-y-Fedw Sandstone Formation and contains at least 12 complete fourth-order cycles. The Bowland Shale exhibits high radiogenic heat productivity in comparison with other sedimentary rocks, due primarily to relative U enrichment under intermittently euxinic conditions. Thermal modelling, however, shows that the radiogenic heat productivity of the Bowland Shale contributes a negligible source of additional heat at the scale of hundreds of metres.

Supplementary material: Discrete bulk geochemical datasets (appendix A) , details of the calculations used Q2 for the heat flow box modelling (appendix B) and original TOC (appendix C) are available at https://doi.org/10.6084/m9.figshare.c.6911105

Organic and metal-rich mudstones, herein termed black shales, represent an important sedimentary rock type as archives of environmental perturbations (e.g. Wignall 1994; Algeo 2004; Meyer and Kump 2008; Kidder and Worsley 2010); sources and resources of hydrocarbons (e.g. Tissot and Welte 1984; Jarvie 2012a, b), ligands (e.g. Kendrick et al. 2002) and metals (e.g. Gadd et al. 2020); seals for injected CO2 (e.g. Worden et al. 2020); and as sources of in situ sedimentary heat (e.g. Rybach 1976; Beardsmore et al. 2010). The UK Late Mississippian–Early Pennsylvanian Bowland Shale Formation represents a relatively thick black shale succession in the UK (e.g. Waters et al. 2007) with equivalent units in Europe (Kerschke and Schulz 2013; Zijp et al. 2017), and is a proven unconventional hydrocarbon resource in the UK (Andrews 2013; Clarke et al. 2018). Here we focus on the Bowland Shale Formation in the Ellesmere Port-1 well, drilled for unconventional hydrocarbons in 2014.

Palpable microseismicity (Clarke et al. 2019), uncertainty related to connectivity to shallow aquifers (e.g. Loveless et al. 2018) and recoverable hydrocarbons (Whitelaw et al. 2019) and overarching net-zero goals (e.g. IPCC 2014) suggest that the Bowland Shale is unlikely to be extensively developed as an unconventional hydrocarbon resource. Despite this, the Bowland Shale Formation remains an important conventional hydrocarbon source rock, for example in the North Sea (Monaghan et al. 2019). The Bowland Shale is a component of at least one geothermal play in the UK where it is a seal for interdigitating sandstone aquifers (Gluyas et al. 2018) and it caps and buries several granitic intrusions in the UK (Leeder 1982; Busby and Terrington 2017), potentially enhancing this geothermal play. The Bowland Shale is a candidate source of base metals (Pb–Zn, Cu) (Parnell and Swainbank 1990; Kendrick et al. 2002; Juerges et al. 2016; Emmings et al. 2020b) and is a borderline highly enriched metalliferous black shale (Emmings et al. 2020c). Weathering of the Bowland Shale may release redox-sensitive metals and semi-metals into the environment, potentially impacting groundwater, soil quality and the health of livestock (Brogan et al. 1973; Parnell et al. 2016). The relatively high U content in the Bowland Shale (Emmings et al. 2020c) implicates it as a source of Rn, derived primarily from the decay of 238U, which may contribute to the known Rn hazard in areas such as the Ribble Valley (Miles et al. 2007) or in deep groundwater aquifers. Finally, the Bowland Shale spans the mid-Carboniferous (Mississippian–Pennsylvanian) boundary and extinction event (Saunders and Ramsbottom 1986), representing a period of extreme climatic and oceanic conditions of interest to a range of deep time palaeoenvironmental and modern climate research questions (e.g. van de Velde et al. 2020).

Here we integrate sedimentological, geochemical and petrophysical observations through two cored sections of the Bowland Shale (historically recognized as the Holywell Shale in North Wales) in Ellesmere Port-1 to develop a high-resolution syngenetic and early diagenetic facies framework through the Bowland Shale. We utilize these observations to build on the idealized sea-level faunal/salinity and sedimentary stacking patterns recognized by Ramsbottom et al. (1962) and Holdsworth and Collinson (1988). Using machine learning of wireline data aligned to core scan X-ray fluorescence (XRF) data, we translate core-scale observations into predictions of sedimentological facies through the entirety of the Bowland Shale in Ellesmere Port-1, representing a c. 324 m thick succession spanning from the Arnsbergian (c. 329 Ma) to the Marsdenian or Yeadonian (c. 317 Ma). The facies framework and interpreted stacking patterns are relevant for any drilling activity into, or through, the Bowland Shale Formation. Finally, we provide new constraints on the radiogenic heat productivity (RHP) and heat flow through the Bowland Shale in Ellesmere Port-1. In an accompanying paper, Hennissen et al. (2024) explore the Bowland Shale in Ellesmere Port-1 in the context of hydrocarbon prospectivity.

Oblique collision between Gondwana and Laurussia during the Carboniferous (Warr 2000) prompted crustal back-arc extension across central and northern England and North Wales, within a seaway that extended from present-day North America to Poland (e.g. Davies et al. 1999). Extension generated a series of graben and half-graben separated by shelves and highs (Fig. 1), termed a ‘block-and-basin’ topography, aligned along inherited pre-Carboniferous structures and bounded by the Southern Uplands in the north and Wales–Brabant Massif (WBM) in the south (Leeder 1982; Fraser and Gawthorpe 1990, 2003; Waters and Davies 2006). This array of interconnected basins gradually merged owing to post-rift thermal subsidence through the Namurian, and are collectively defined as the Pennine Basin, representing a long-lived depocentre during most of the Carboniferous (Waters and Davies 2006).

Fig. 1.

Brigantian block-and-basin structures and key borehole and outcrop sections across central Britain, with emphasis on the vicinity of the Blacon Basin. BL, Bala Lineament; BLW-E, Blacon East-1 well; CLH, Central Lancashire High; CRO, Croxteth-1 well; EFZ, Edgerley–Waverton Fault Zone; ERB, Erbistock-1 well; GEF, Great Ewloe Fault; HH, Holme High; LRR, Llŷn–Rossendale Ridge; MGH, Milton Green High; MIL, Milton Green-1 well; NFZ, Nercwys–Nant-figillt Fault Zone; PF, Pendle fault; RB, Rossendale Basin. Source: after Andrews (2013), Fraser and Gawthorpe (1990, 2003), Fraser et al. (1990), Pharaoh et al. (2020), Smith et al. (2005), Waters and Davies (2006) and Waters et al. (2007, 2009, 2011, 2020, 2021).

Fig. 1.

Brigantian block-and-basin structures and key borehole and outcrop sections across central Britain, with emphasis on the vicinity of the Blacon Basin. BL, Bala Lineament; BLW-E, Blacon East-1 well; CLH, Central Lancashire High; CRO, Croxteth-1 well; EFZ, Edgerley–Waverton Fault Zone; ERB, Erbistock-1 well; GEF, Great Ewloe Fault; HH, Holme High; LRR, Llŷn–Rossendale Ridge; MGH, Milton Green High; MIL, Milton Green-1 well; NFZ, Nercwys–Nant-figillt Fault Zone; PF, Pendle fault; RB, Rossendale Basin. Source: after Andrews (2013), Fraser and Gawthorpe (1990, 2003), Fraser et al. (1990), Pharaoh et al. (2020), Smith et al. (2005), Waters and Davies (2006) and Waters et al. (2007, 2009, 2011, 2020, 2021).

Ellesmere Port-1 is situated on the NE margin of the Blacon Basin in North Wales and Cheshire, a sub-basin of the Pennine Basin (Andrews 2013), which was possibly linked to the Widnes Basin of Fraser et al. (1990) (Fig. 1). The Blacon Basin is bounded in the south by the Bryneglwys fault (the Bala Lineament) and WBM, in the west by the Nercwys–Nant-figillt Fault Zone, Vale of Clwyd Fault and North Wales Shelf, and in the NE by the Great Ewloe Fault, Edgerley–Waverton Fault Zone (EFZ) and ultimately the Central Lancashire High (Fig. 1) (Smith et al. 2005). The locus of deposition in the Blacon Basin was consistently centred immediately to the north of the Bala Lineament (on the downthrown hanging wall block), as evidenced by: (1) a thick Early Carboniferous carbonate succession (Smith et al. 2005); (2) thickening of Minera Formation (Brigantian) sandstones into the basin (Davies et al. 2004); (3) early Namurian (Pendleian) megasliding on the southwestern fringe of the basin (Kirkham 2021); (4) an expanded Namurian succession in the Blacon East-1 well in the Blacon Basin (Davies et al. 2004; Andrews 2013; Waters et al. 2021); (5) pinch-out of the Namurian succession on the margin of the North Wales Shelf in the west (Williams and Eaton 1993); and (6) a highly condensed Namurian succession in the Erbistock-1 well on the emergent landmass in the south (Andrews 2013) (Fig. 1). A condensed Namurian succession is also proven in the Milton Green-1 well (Andrews 2013) on the nearby Milton Green High (Smith et al. 2005).

During the Brigantian (regional substage in the latest Visean), widespread ramp-to-shelf carbonates of the Clwyd Limestone Group, including the Cefn Mawr Limestone Formation and Minera Formation, indicate development of shallow water carbonate platform and slope conditions (Somerville 1979; Davies et al. 2004, 2011; Waters et al. 2021). The Cefn Mawr Limestone Formation exhibits shoaling-upward sequences of wackestone, packstone and crinoidal limestone, often capped by karstic surfaces, soils and calcrete horizons. The sequences were primarily driven by eustatic sea-level fluctuation overprinted by late synrift tectonism, including subaerial platform exposure during lowstands (Somerville 1979). The Minera Formation, defined by cyclic sequences of mudstone, packstone, wackestone and calcareous sandstone, represents a fully subaqueous, proximal, sand-rich equivalent to the Cefn Mawr Limestone (Davies et al. 2004). Off-platform and distal to detrital sand input, the Brigantian succession grades laterally into the deep-water argillaceous carbonate slope apron facies of the Prestatyn Limestone Formation and the succeeding Teilia Formation (Warren et al. 1984; Davies et al. 2011). Initially shallow waters probably covered the Llŷn–Rossendale Ridge (Jackson et al. 1997; Smith et al. 2005). This high, representing the northern limit to the Blacon Basin, was probably connected to the Central Lancashire High (Fig. 1; sensu.Fraser and Gawthorpe 1990; DECC 2013) as the manifestation of the southwestern end of the Pendle fault system (Pharaoh et al. 2020). The presence of the Prestatyn Limestone and Teilia Formation near the NE margin the North Wales Shelf suggests that the Llŷn–Rossendale Ridge was a discontinuous feature, enabling weak connectivity between the Blacon Basin and distal Craven Basin (Davies et al. 2004; Waters et al. 2007; Andrews 2013; Juerges et al. 2016).

A transition from the Clwyd Limestone Group to the Craven Group at the Visean–Namurian boundary indicates an important regional shift from active rifting to thermal subsidence (Leeder 1982; Waters et al. 2009), and, as a consequence, a switch from a carbonate to siliciclastic depositional system (e.g. Fraser and Gawthorpe 2003). Cessation of active rifting and onset of thermal subsidence prompted progressive infill of the inherited Visean basin structures during the Namurian (Fig. 1). Assuming relatively sheltered conditions (Peters and Loss 2012), the inherited platforms and highs probably initially exhibited a water depth of less than 50 m. Taking the proximal Craven Basin as an analogue (Emmings et al. 2020a), the water depth in the Blacon Basin probably ranged from c. 100 to 200 m on the flanks and highs (Holdsworth and Collinson 1988; Davies 2008) to several hundred metres in the deepest part of the basin (Davies et al. 1993).

The Pentre Chert Formation represents the basal part of the Craven Group in the Blacon Basin (Fig. 2), which unconformably overlies the Clwyd Limestone Group (Davies et al. 2004). The Pentre Chert is defined as a sponge spicule- and radiolarian-bearing chert deposited in relatively deep water, below storm wave base (Davies et al. 2011). The Pentre Chert probably represents the condensed counterpart of Pendleian (E1) siliceous mudstones of the Bowland Shale Formation in the Craven Basin (Emmings et al. 2020a, b). During the Pendleian, most of the siliciclastic sediment was sourced from the Pendle delta system in the north of the Craven Basin (Fraser and Gawthorpe 1990, 2003; Waters and Condon 2012), ensuring that the Blacon Basin remained isolated from detrital sediment. Deposition of siliceous sediments in the Blacon Basin during the Pendleian is consistent with: (1) development of widely hypoxic or intermittently anoxic bottom water conditions, favouring colonization by sponges (e.g. Schuster et al. 2021); (2) water column conditions that favoured colonization by radiolarians (Emmings et al. 2020a, c), perhaps associated with elevated riverine nutrient-loading and/or silica input (Dunham and Wilson 1985); and (3) increased influx of detrital mud onto platforms (potentially a distal extension of the Pendle delta), prompting regional termination of carbonate production (Davies et al. 2004).

Fig. 2.

General stratigraphy, regional Arnsbergian lithofacies and selected well log data. (a) Generalized stratigraphy and (b) regional lithofacies from Waters et al. (2007), including absolute ages (from Davydov et al. 2012; Waters and Condon 2012) and the updated Early–Late Serpukhovian boundary (Cózar and Somerville 2021). (c) The Bowland Shale Formation exhibits elevated total gamma-ray (GR, in American Petroleum Institute units (API)) and U (ppm) concentrations. (d) Correlation between wireline gamma v. synthetic gamma-ray (estimated from measured U, Th and K contents; see ‘Methods’) verifies the alignment of reported corebox depths to the wireline data (see Fig. 3). Ellesmere Port-1 is highlighted (53° 17′ 43.3841″ N, 02° 54′ 21.13091″ W; 339629 mE 378016 mN, British National Grid). Y, Yeadonian; M, Marsdenian; K, Kinderscoutian; Al, Alportian; C, Chokierian; Ar, Arnsbergian; P, Pendleian; Brig., Brigantian.

Fig. 2.

General stratigraphy, regional Arnsbergian lithofacies and selected well log data. (a) Generalized stratigraphy and (b) regional lithofacies from Waters et al. (2007), including absolute ages (from Davydov et al. 2012; Waters and Condon 2012) and the updated Early–Late Serpukhovian boundary (Cózar and Somerville 2021). (c) The Bowland Shale Formation exhibits elevated total gamma-ray (GR, in American Petroleum Institute units (API)) and U (ppm) concentrations. (d) Correlation between wireline gamma v. synthetic gamma-ray (estimated from measured U, Th and K contents; see ‘Methods’) verifies the alignment of reported corebox depths to the wireline data (see Fig. 3). Ellesmere Port-1 is highlighted (53° 17′ 43.3841″ N, 02° 54′ 21.13091″ W; 339629 mE 378016 mN, British National Grid). Y, Yeadonian; M, Marsdenian; K, Kinderscoutian; Al, Alportian; C, Chokierian; Ar, Arnsbergian; P, Pendleian; Brig., Brigantian.

The Pentre Chert is overlain by the Bowland Shale Formation (herein Bowland Shale), a highly heterogeneous, mudstone-dominated succession spanning from the mid-late Pendleian (c. 329 Ma) to Marsdenian or Yeadonian in Ellesmere Port-1 (c. 317 Ma; Fig. 2) (Davies et al. 2011; Waters and Condon 2012). The Bowland Shale definition adopted (Waters et al. 2009) encompasses the Holywell Shale in North Wales (e.g. Davies et al. 2004; Newport et al. 2016). The onset of third- to fourth-order sea-level fluctuation (Mitchum and Van Wagoner 1991; Waters and Condon 2012) coupled to thermal subsidence prompted deposition of the Bowland Shale across a wide range of basinal, prodelta and periodically emergent shelfal environments (Davies et al. 2004). The Pentre Chert and Bowland Shale grade laterally southwards into the Cefn-y-Fedw Sandstone Formation of the Millstone Grit Group derived primarily from the WBM (Fig. 2) (Davies et al. 2004; Waters et al. 2009, 2021). The Cefn-y-Fedw Sandstone comprises coarsening–upward cycles of quartzitic sandstones interbedded with subordinate mudstone and coal beds, interpreted as fluviodeltaic progradational sequences (Davies et al. 2004; Waters et al. 2021). In the Blacon Basin, three turbidite sand packages within the Bowland Shale represent the distal extensions of the Lower, Middle and Upper Cefn-y-Fedw Sandstone units (Davies et al. 2004; Waters et al. 2021). The Cefn-y-Fedw Sandstone is analogous and equivalent to sands of the Morridge Formation in the Widmerpool Gulf and Edale Gulf (Waters et al. 2009; Morton et al. 2015; Hennissen et al. 2017); both are associated with delta systems shed from the WBM in the south (Fraser and Gawthorpe 1990; Waters et al. 2021). The Bowland Shale grades into, and is overlain by, the fluviodeltaic Gwespyr Sandstone Formation (also Millstone Grit Group). Waters et al. (2021) showed that the Gwespyr Sandstone Formation derives from a combination of provenances from the south, north and west.

Marine transgressions and associated maximum flooding surfaces in the Bowland Shale are manifested as carbonate-rich, typically macrofauna-bearing sedimentary packages termed ‘marine bands’ (e.g. Ramsbottom 1977). Sea-level cyclicity in the Namurian was probably driven by far-field glaciation on Gondwana (Veevers and Powell 1987). The amplitude of sea-level fluctuation is estimated at 42 m (Maynard and Leeder 1992) or 60–100 m (Church and Gawthorpe 1994; Rygel et al. 2008). Marine bands on shelves mark the base of cyclothemic ‘Yoredale’ limestone–shale–sandstone triplets commonly capped by coal (e.g. Hampson 1997; Davies et al. 2004; Waters et al. 2007; Dean et al. 2011). Basinal marine bands are typically 1–5 m thick and are often overlain by an asymmetrical sequence of: (1) non-bioclastic, hemipelagic (lenticular) mudstone; and (2) turbidites, hybrid beds and debrites (Martinsen et al. 1995; Emmings et al. 2020a). Lowstands were associated with localized to extensive basinal sand deposition depending on proximity to delta systems (Millstone Grit Group). The third- and fourth-order marine band cycles may be superimposed onto 1.1–1.35 Myr duration Namurian ‘mesothem’ cycles (Ramsbottom 1979), associated with 17 major base-level falls (Waters and Condon 2012) and shifts in the loci of sediment deposition (Martinsen et al. 1995). Salinity fluctuations from fully marine (transgressions) to brackish or freshwater conditions are also widely recognized in the Bowland Shale on the basis of volumetric calculations (Holdsworth and Collinson 1988), macrofauna and microfauna such as conodonts (Ramsbottom et al. 1962; Holdsworth and Collinson 1988), biomarkers (Gross et al. 2015; Sims et al. 2024), sedimentology (Holdsworth and Collinson 1988; Gross et al. 2015; Emmings et al. 2020a, c) and inorganic geochemistry (Leeder et al. 1990; Maynard et al. 1991; Emmings et al. 2020c).

Coupled to sea-level and salinity fluctuation, palaeoredox proxies show that the Bowland Shale was deposited under fluctuating oxic, anoxic and at least intermittently sulfidic (euxinic) conditions (Leeder et al. 1990; Gross et al. 2015; Riley et al. 2018; Emmings et al. 2019, 2020c), a phenomenon probably driven by euhaline phytoplankton productivity. Euxinia in the Bowland Shale (and younger Carboniferous rocks) is most strongly associated with transgressions and maximum flooding, whereas reduced basin accommodation is linked to increased bottom water ventilation (Bloxam and Thomas 1968; Leeder et al. 1990; Maynard et al. 1991; Fisher and Wignall 2001; Pearce et al. 2010; Gross et al. 2015; Emmings et al. 2020c). Considering the evidence for significant freshwater discharge from delta systems, shelfal anoxia (as opposed to basinal anoxia) was possibly caused by upwelling of marine anoxic bottom waters trapped beneath a freshwater lid (Algeo et al. 2008). By contrast trace-element quotas in the basinal shales favour the development of anoxia in weakly restricted, thermally stratified and highly productive water column conditions (Riley et al. 2018; Emmings et al. 2020c). Euxinia was potentially confined to mid-depth waters, potentially part of one or more linked regional-scale oxygen minimum zones that developed in the late Devonian (Kabanov and Jiang 2020; Emmings et al. 2022).

Macrofauna associated with the marine bands form the basis of a high-resolution biostratigraphic framework through the Namurian (Ramsbottom and Saunders 1985; Holdsworth and Collinson 1988). Marine bands and overlying mudstones exhibit the following idealized faunal sequence (phases 6 to 1): thick-shelled ammonoids (6); thin-shelled ammonoids (5); molluscan spat (4); Lingula (3); Planolites (2); fish remains (1); and barren zones (Ramsbottom et al. 1962; Baines 1977; Ramsbottom 1977). These faunal phases are thought to indicate cycling between euhaline (6) and freshwater conditions (1) (Ramsbottom et al. 1962; Holdsworth and Collinson 1988), although preservation or absence of bioclasts is also dependent on early diagenetic porewater conditions (Emmings et al. 2020c). The Bowland Shale in North Wales spans all the regional Namurian substages, from the Pendleian (E1) to Yeadonian (G1) (Davies et al. 2011). Marine bands exhibit an average periodicity of 111 kyr (Waters and Condon 2012), indicative of fourth-order cycles (Mitchum and Van Wagoner 1991). E1c1, E2a2, E2b3, H1a, H2b1, R1b, R1c4, R2a1, R2b1, R2c1, G1a1 and G1b1 peak flooding events are recognized in North Wales (Davies et al. 2011), probably related to third-order forcing (Mitchum and Van Wagoner 1991; Waters and Condon 2012). Multiple flooding surfaces are recognized for many of the marine bands, potentially linked to sub-100 kyr precession or obliquity cycles (Waters and Condon 2012).

Pb–Zn mineralization on the North Wales Shelf suggests thermal maturation and dewatering of adjacent basinal sediments in the latest Namurian (Juerges et al. 2016). During the late Pennsylvanian, closure of the Rheic–palaeo-Tethys seaway associated with the Variscan Orogeny (e.g. Leeder 1982) prompted a major depositional hiatus and the inversion of many basin structures across central England (e.g. Smith et al. 2005). In the Pennines, inversion combined with a relatively high basinal heat flow (in contrast to North Wales) enhanced basin dewatering and led to protracted Pb–Zn–F mineralization on blocks (Bouch et al. 2006; Hollis and Walkden 2012). Inverted structures in the vicinity of the Blacon Basin include the anticlines on the Bala Lineament and EFZ (Smith et al. 2005) and the Môn–Deemster–Ribblesdale fold belt in the north (Arthurton 1984; Pharaoh et al. 2020). The Cheshire Basin, partially superposed on the Blacon Basin, developed in response to renewed extension during the Permo-Triassic (Davies et al. 2004). The Permo-Triassic in the Cheshire Basin comprises aeolian and fluvial sediments (Sherwood Sandstone Group); Fe–Cu mineralization on the adjacent North Wales Platform may derive from charging by Permo-Triassic oxidized brines (Juerges et al. 2016) or from basinal Carboniferous fluids (Parnell and Swainbank 1990). Maximum burial of the Blacon Basin was probably attained in the Late Cretaceous, prior to Tertiary uplift (Williams and Eaton 1993; Green et al. 1997; Pharaoh et al. 2019).

Two 17.8 m cored sections through the borehole Ellesmere Port-1 (53° 17′ 43.3841″ N, 02° 54′ 21.13091″ W; 339629 mE 378016 mN, British National Grid), drilled by IGas Energy in 2014, were logged and sampled at the National Geoscience Data Centre (NGDC) core repository, British Geological Survey. Sedimentological structures and facies were logged prior to major- and minor-element XRF core scanning and sampling. The reported absolute core depths were aligned with the reported wireline absolute depth by matching features observed in core to the borehole image logs (Figs 2 & 3). For validation, we compared synthetic (derived from discrete XRF) and measured gamma-ray curves (Fig. 2). The wireline gamma ray tends to smooth across thin beds exhibiting particularly high or low true (synthetic) gamma-ray values, but overall, the synthetic and wireline gamma-ray datasets are strongly correlated (Fig. 2d). This assessment indicated an offset of 1.6 and 1.76 m between the downhole absolute depth and reported cored depth, for core 1 and core 2, respectively. The synthetic total gamma ray was estimated from bulk U, Th and K concentrations using the approximation of Ellis and Singer (2007) (equation 1):
(1)
Fig. 3.

Key static borehole image log and core observations used for the alignment of reported well and core depth. Depth alignment was achieved by matching the same linear features present in both datasets. MTD, mass transport deposit.

Fig. 3.

Key static borehole image log and core observations used for the alignment of reported well and core depth. Depth alignment was achieved by matching the same linear features present in both datasets. MTD, mass transport deposit.

Core scan XRF measurements were collected using an Itrax MC core scanner (Cox Analytical Systems) in the Core Scanning Facility at the British Geological Survey, Keyworth. This dataset is available under Open Government Licence (BGS Core Scanning Facility 2022). Before core scanning, fragments of the split core were aligned with a levelling tool to ensure optimal alignment of the detector. XRF was measured at a 1 cm step-size and 30 s integration time. For each core section, an initial surface scan was performed to compute the trajectory of the XRF detector. The Itrax MC core scanner automatically produces an XRF sum spectrum for each core section, which was peak-fitted using the processing software Q-Spec version 8.6 (Cox Analytical Systems, Sweden). A single fitting template was used to fit most of the core data. Core horizons that showed high Ca, Fe or Cl elemental abundances were refitted accordingly. These fitted XRF data were exported as elemental peak areas (counts) for processing in R. Data were aggregated into the two core runs and assigned a true core depth. The dataset was cleaned to remove erroneous data points, for example spanning cracks, missing core sections and intervals of generally poor core quality.

A total of 110 samples, split between core 1 (57 samples) and core 2 (53 samples), were selected for bulk major- and trace-element geochemistry, Rock-Eval pyrolysis, palynology (see Hennissen et al. 2024) and stable isotope analysis. Discrete XRF major- and trace-element data were acquired on fused beads (using ignited sample powders) and powder briquettes at the University of Leicester, using a PANalytical Axios Advanced XRF spectrometer and default PANalytical SuperQ conditions (Malvern PANalytical, Malvern, UK). Repeat analyses of reference materials and prior Bowland Shale samples indicate an XRF accuracy of ±5% for major elements and ±5% for trace elements, and a precision of up to ±5% (but commonly ±1%, 2σ) for major elements and ±10% (but commonly ±5%) for trace elements (Emmings 2018; Emmings et al. 2020c).

Total sulfur (TS) content was determined via a Flash 1112 elemental analyser coupled to an isotope ratio mass spectrometer at the National Environmental Isotope Facility (NEIF) laboratory at the British Geological Survey. TS accuracy and precision are estimated at ±15% and ±0.2 wt% S, respectively, based on repeat analyses and independent measurement via LECO elemental analysis (Emmings et al. 2020c). Trace-element enrichment factors (EFs; e.g. Tribovillard et al. 2006) were utilized using Post–Archean Average Shale (PAAS) (Taylor and McLennan 1985) (equation 2) and local mean concentrations (sensuEmmings et al. 2020c). Pyrolysis was conducted in standard mode using a Rock-Eval 6 apparatus (Vinci Technologies, Nanterre, France) at the British Geological Survey. Accuracy is typically 1–2% for most Rock-Eval parameters, based on repeat analyses of the IFP Rock-Eval standard. Repeat analyses of similar Bowland Shale samples indicate a precision (2σ) of ±0.11 mgHC g−1 (S1), ±0.44 mgHC g−1 (S2) and ±0.03 wt% total organic and inorganic carbon (total organic carbon; TOC, inorganic/mineral carbon; MINC). Derived Rock-Eval parameters include the oil saturation index (OSI) and hydrogen index (HI) (Lafargue et al. 1998).
(2)
The original TOC (TOCo) was calculated for palynological samples using the method of Jarvie (2012a) and values are reported in Hennissen et al. (2024). TOCo was extrapolated for all discrete samples (n = 110) via a linear regression between the calculated TOCo and present-day TOC (TOCpd, r2 = 0.99; see Supplementary material; equation 3), where:
(3)

On the basis that the core sections are oil to marginally gas-mature (Hennissen et al. 2024), TOCo was preferred over TOCpd for a more robust S/TOC palaeosalinity proxy (Raiswell and Berner 1987).

All stable isotope data are reported in standard delta (δ) notation as per mil (‰) variations. Bulk S isotope ratios were measured using a Thermo Fisher Delta Plus XL isotope ratio mass spectrometer interfaced with a Flash 1112 elemental analyser. δ34Sbulk values are reported on the Vienna-Canyon Diablo Troilite standard (VCDT) scale. S isotope ratios were calibrated using a linear fit through IAEA-S-1 (−0.3‰), IAEA-S-2 (+22.67‰) and IAEA-S-3 (−32.3‰). δ34Sbulk reproducibility is ±1.2‰ (2σ) based on repeat analyses.

Bulk C isotope analyses were performed by combustion in a Costech ECS4010 Elemental Analyser on-line to a VG TripleTrap and Optima dual-inlet mass spectrometer, with δ13C values calculated to the Vienna Peedee Belemnite (VPDB) scale using within-run laboratory standards calibrated against the NBS-19 and NBS-22 (VPDB) scale, using within-run laboratory standards (NBS19 +1.95‰ and NBS22 −30.03‰). δ13Corg reproducibility is ±<0.1‰ (1σ). C/N ratios are calibrated against a Broccoli standard (BROC2). Replicate analysis of well-mixed samples indicated a precision of ±<0.1‰.

All data manipulation and statistical analysis was coded in R using RStudio (R Core Team 2018). XRF core scan, discrete analyses (conventional XRF, stable isotopes, Rock-Eval pyrolysis) and wireline data were manipulated using the following R packages: ggplot2 (Wickham 2009), gridExtra (Auguie 2016), zoo (Zeileis and Grothendieck 2005), plyr and dplyr (Wickham and Francois 2016), data.table (Dowle and Srinivasan 2020), compositions (van den Boogaart et al. 2020), ggtern (Hamilton 2017) and caret (Kuhn 2020). Chemofacies were determined by hierarchical cluster analysis (HCA) of principal component analysis (PCA) scores (e.g. Emmings et al. 2022) using selected core scan XRF major and minor elements (Mg, Al, Si, P, K, Ca, V, Mn, Fe, Ni, As, Sr, Zr, Mo and Ba). These elements were selected on the basis of relatively high abundances, and as proxies for detrital phases (Al, Si, K, Zr; e.g. Hild and Brumsack 1998; Emmings et al. 2020a), Fe oxides (Fe, As; e.g. Tribovillard 2020), autochthonous input including biogenic and diagenetic carbonates (Si, Ca, Mn, Mg, Sr, Ba, P) and palaeoredox (Fe, Mn, Mo, V, Ni, As; e.g. Tribovillard et al. 2006). Processed count data were not converted to absolute concentrations; this approach circumvents closed data effects that require transformation prior to statistical analysis (e.g. centred log-ratio; Aitchison 1986). PCA was conducted on the correlation matrix (scaled and centred data). The first six principal components were selected on the basis of assessment of the hinge point on the scree plot. HCA of the sample scores for the first six principal components was conducted via the sum of squares (Ward clustering) and Euclidian distance matrix.

We explored the potential of machine learning to aid in extrapolating the core scan clusters (e.g. chemofacies) through the Bowland Shale Formation. Ahead of machine learning, the seven chemofacies were aggregated into four classes: limestones (A); weakly to moderately calcareous mudstones and siltstones (C–E); claystone density flow deposits (B, F); and siltstone to sandstone density flow deposits (G). Simplification into four classes for machine learning was justified on the basis of (1) experimentation to optimize the balance between the accuracy and complexity of predictions and (2) the thickness of beds and co-occurrences with other chemofacies: specifically, laminae of siderite-cemented claystone or siltstone are not directly resolvable in the well log data, but commonly co-occur with the claystone chemofacies (F).

The k-nearest neighbour (kNN) algorithm was selected for machine learning based on superior accuracy compared to linear discriminant analysis (LDA), classification and regression trees (CART), support vector machine (SVM) and random forest methods. The four simplified chemofacies classes were matched to the nearest, depth-aligned well log datum. The models were trained and deployed using the following well log data: compensated neutron porosity (CN); compressional wave slowness (DTC); shear wave slowness (DTS); spectral gamma-ray (U, Th, K); photo-electric (PE); and calculated formation bulk density (ZDEN). We trialled three approaches: (1) kNN using raw wireline data (87% accuracy at 0.05 P-value, optimal k = 7); (2) kNN using standardized (scaled and centred z-score) wireline data (89% accuracy at 0.05 P-value, optimal k = 5); and (3) kNN using standardized data and with null well log values (−9999) removed (87% accuracy at 0.05 P-value, optimal k = 5). Training, verification and accuracy estimation was determined on a 4:1 training-to-verification split. It is beyond the scope of this chapter to further refine the machine learning algorithms, but future work could impute null values (e.g. Dmitrijeva et al. 2020) as a compromise between methods (2) and (3). The trained models were deployed on the entire Bowland Shale section at Ellesmere Port-1 (1443 to 1767 m below surface) to delineate limestone, calcareous mudstone, claystone and siltstone to sandstone (density flow).

RHP (in μW m−3) was calculated using the equation of Rybach (1976) (equation 4):
(4)
where U, Th and K concentrations are in ppm, ppm and wt%, respectively, and p is the bulk density (g cm−3). We assume p = 2.6 g cm−3 for the Bowland Shale (Clarke et al. 2018). The temperature gradient through the Bowland Shale was modelled using the equation of Beardsmore et al. (2010), where T is the temperature at the base of the given sedimentary package, T0 is the temperature at the top of the package, Q0 is the surface heat flow (W m−2), Ks is the thermal conductivity (W m−1 K−1), As is RHP (in W m−3) and S is the thickness of the sedimentary package (m), and inversely proportional to the thermal conductivity (Ks in W m−1 K−1), as defined by equation (5):
(5)

Based on macrofaunal (ammonoid) biostratigraphy, cores 1 and 2 are assigned to the Late Kinderscoutian (R1c1) and late Arnsbergian–Chokierian (H1a1–H1a3) biozones, respectively (Figs 4 & 5). Thus, core 2 spans the onset of the mid-Carboniferous boundary (Saunders and Ramsbottom 1986), although the presence of intervening faults and a paucity of diagnostic Arnsbergian macrofauna does not preclude the potential for hiatuses beneath the H1a1–H1a3 interval. The core sections comprise packages of (1) massive, weakly to strongly calcareous, non-bioclastic to bioclastic mudstones, often interbedded with thin limestone beds, overlain by (2) thinly laminated or chaotic, non-calcareous claystone beds (Figs 4 & 5). The calcareous, bioclastic mudstones are often macrofauna-bearing and are collectively interpreted as basinal marine band facies, characteristic of the Bowland Shale, deposited during fourth-order marine transgressions and highstands (e.g. Martinsen et al. 1995; Waters and Condon 2012).

Fig. 4.

Sedimentology and selected geochemistry through the late Arnsbergian–Chokierian Bowland Shale in Ellesmere Port-1 (core 2). See Figure 3 for key to sedimentary structures. Palynofacies data are from Hennissen et al. (2024), using the fields of Tyson (1995). I, highly proximal shelf/basin; II, marginal dysoxic–anoxic basin; III, heterolithic (proximal) shelf; IVa, dysoxic–suboxic shelf edge; IVb, dysoxic–anoxic shelf edge; V, mud-dominated (distal) oxic shelf; VI, proximal suboxic–anoxic shelf; VII, distal dysoxic–suboxic shelf; VIII, distal dysoxic–anoxic shelf; IX, distal suboxic–anoxic basin. MTD, mass transport deposit.

Fig. 4.

Sedimentology and selected geochemistry through the late Arnsbergian–Chokierian Bowland Shale in Ellesmere Port-1 (core 2). See Figure 3 for key to sedimentary structures. Palynofacies data are from Hennissen et al. (2024), using the fields of Tyson (1995). I, highly proximal shelf/basin; II, marginal dysoxic–anoxic basin; III, heterolithic (proximal) shelf; IVa, dysoxic–suboxic shelf edge; IVb, dysoxic–anoxic shelf edge; V, mud-dominated (distal) oxic shelf; VI, proximal suboxic–anoxic shelf; VII, distal dysoxic–suboxic shelf; VIII, distal dysoxic–anoxic shelf; IX, distal suboxic–anoxic basin. MTD, mass transport deposit.

Fig. 5.

Sedimentology and selected geochemistry through the late Kinderscoutian Bowland Shale in Ellesmere Port-1 (core 1). See Figure 3 for key to sedimentary structures. Palynofacies data are from Hennissen et al. (2024), using the fields of Tyson (1995). I, highly proximal shelf/basin; II, marginal dysoxic–anoxic basin; III, heterolithic (proximal) shelf; Iva, dysoxic–suboxic shelf edge; IVb, dysoxic–anoxic shelf edge; V, mud-dominated (distal) oxic shelf; VI, proximal suboxic–anoxic shelf; VII, distal dysoxic–suboxic shelf; VIII, distal dysoxic–anoxic shelf; IX, distal suboxic–anoxic basin. MTD, mass transport deposit.

Fig. 5.

Sedimentology and selected geochemistry through the late Kinderscoutian Bowland Shale in Ellesmere Port-1 (core 1). See Figure 3 for key to sedimentary structures. Palynofacies data are from Hennissen et al. (2024), using the fields of Tyson (1995). I, highly proximal shelf/basin; II, marginal dysoxic–anoxic basin; III, heterolithic (proximal) shelf; Iva, dysoxic–suboxic shelf edge; IVb, dysoxic–anoxic shelf edge; V, mud-dominated (distal) oxic shelf; VI, proximal suboxic–anoxic shelf; VII, distal dysoxic–suboxic shelf; VIII, distal dysoxic–anoxic shelf; IX, distal suboxic–anoxic basin. MTD, mass transport deposit.

The overlying non-calcareous claystone packages represent deposition from a range of turbulent, hybrid and laminar flows (e.g. Talling et al. 2012) during falling basin accommodation (Figs 4 & 5). In some cases the non-calcareous claystone packages contain pyritized organic matter and pyrite nodules, similar to the Bowland Shale in the Craven Basin, suggestive of intermittently non-steady-state sulfidic pore water conditions (Emmings et al. 2020c). In many cases, consecutive, stacked chaotica suggest that the claystone packages were transported en masse, potentially as discrete, metre-scale blocks rafted within debris flows (e.g. mass transport deposits; MTDs). Towards to the top of both core sections, the non-calcareous claystones are capped by thinly to thickly laminated, interbedded, bioturbated siltstone and sandstone facies, probably deposited during lowstands. In core 1, the contact between the siltstones and the underlying claystones is faulted; thus, the sequence stratigraphic relationship is unclear.

Zr/Al aliases the heavy mineral (zircon) component and is interpreted as a local proxy for the detrital grain size in this system (e.g. silt- to sand-fraction; Dowey and Taylor 2020; Emmings et al. 2020a). This is consistent with the observed relationship between Zr/Al and sedimentary facies; the marine band, claystone and siltstone facies exhibit low, medium and high to very high Zr/Al, respectively (Figs 4 & 5). Mo is a redox-sensitive metal that is fixed under at least intermittently sulfidic pore water conditions (e.g. Tribovillard et al. 2006). Elevated Mo/Al within the marine band packages (in both core sections) suggests the development of anoxic and intermittently sulfidic conditions coupled to sea-level rise. This is consistent with a productivity-driven, weakly restricted model for anoxia during deposition of the Bowland Shale (Emmings et al. 2020c). Arsenic is also redox-sensitive in the marine environment, but interpretation in terms of palaeoredox is generally complex (Tribovillard 2020). In this system, As/Al tends to co-vary with Mo/Al (Figs 4 & 5); this suggests operation of a strong Fe–Mn (oxyhydr)oxide shuttle during deposition (Tribovillard 2020), analogous to the model for anoxia in the adjacent Craven Basin (Emmings et al. 2020c). As/Al enrichment tends to span a wider depth interval compared to Mo/Al, and As/Al sinuosity, particularly in core 1, suggests generally high sensitivity to intermediate redox conditions compared with Mo/Al. Thus, high As/Al coupled to low Mo/Al is interpreted to indicate ferruginous water column conditions overlying non-sulfidic pore waters (Figs 4 & 5).

Finally, the palaeosalinity measure Sr/Ba (Wei and Algeo 2020) shows that conditions were generally marine and intermittently brackish. Whilst the Sr/Ba salinity proxy is limited to non-carbonate-bearing lithologies, the presence of euhaline macrofauna within the marine band facies permits a comprehensive reconstruction of the palaeosalinity variation throughout both cores (Figs 4 & 5). Several of the interpreted falling stage to lowstand sedimentary packages exhibit Sr/Ba <0.5; this suggests deposition under at least intermittently brackish conditions during periods of reduced basin accommodation. Sr/Ba shows that reduced salinity was more prevalent in the late Kinderscoutian section (core 1) compared to the generally more saline Arnsbergian–Chokierian section (core 2); this is consistent with the increasing proximity (i.e. progradation) of the Cefn-y-Fedw delta system towards the top of the Bowland Shale.

Cluster analysis of selected XRF core scan major and minor elements delineates seven chemofacies (A–G), interpreted as: limestone deposited under sulfidic conditions (A); siderite-cemented claystone or siltstone deposited under mixed redox conditions (B); bioclastic, moderately to strongly calcareous mudstone deposited under sulfidic conditions (C); weakly to moderately calcareous phosphatic mudstone deposited under mixed redox conditions (D); weakly bioclastic, weakly calcareous mudstone deposited under weakly sulfidic to non-sulfidic conditions (E); claystone deposited under oxic or suboxic conditions (density flows) (F); and siltstone to sandstone deposited under oxic or suboxic conditions (density flows) (G). Thus, the chemofacies are an objective measure for lithological variation, supplementary to the conventional sedimentological descriptions (Figs 4 & 5). The XRF core scan chemofacies are generally consistent with the conventional sedimentary descriptions, although there are some discrepancies, particularly in the assignment of calcareous and non-calcareous sediments. This is best explained by the difficulty in the visual estimation of the fraction of carbonate (e.g. reaction with a small amount of HCl).

A pseudo-compositional ternary plot of the core scan XRF data indicates most intervals are argillaceous, and less commonly quartz or carbonate-rich (Fig. 6). Carbonate-bearing, siliceous mudstones are rare (Fig. 6); this is consistent with the absence of silica cementation in the two cored sections. The discrete quantitative XRF analyses support this observation; dual Si v. Al (Fig. 7a) and Si v. Zr (Fig. 7b) biplots of Emmings et al. (2020b) show that biogenic silica is negligible in the two cored intervals. Thus, it is clear the cored sections are not comparable to, or transitional from, the underlying Pentre Chert and do not contain reworked biogenic silica. By extension, the lack of pervasive cementation shows that the cored sections in Ellesmere Port-1 are not comparable to key intervals in the Craven Basin, such as: (1) the siliceous mudstones of the Upper Bowland Shale (Emmings et al. 2020b); and (2) the calcareous and siliceous mudstones and limestones of the Lower Bowland Shale (e.g. Fauchille et al. 2017, 2018; Newport et al. 2018). Instead, the observed pattern of interfingering argillaceous and carbonate marine band facies is more comparable to the Bowland Shale in other paralic basins, such as the Widmerpool Gulf (Könitzer et al. 2014; Hennissen et al. 2017).

Fig. 6.

Clay, carbonate and quartz bulk compositional ternary plot estimated using scaled core scan XRF count data (2Al, 0.05Ca and 10Zr, respectively). See text for discussion.

Fig. 6.

Clay, carbonate and quartz bulk compositional ternary plot estimated using scaled core scan XRF count data (2Al, 0.05Ca and 10Zr, respectively). See text for discussion.

Fig. 7.

Excess silica biplots, after Emmings et al. (2020b). Regression lines for the Ellesmere Port-1 data (red line, with 95% confidence interval) and Emmings et al. (2020b) are also plotted. The excess Si field delineates siliceous mudstones. The field is calibrated for Bowland Shale mudstones and compositionally immature siltstones and is not necessarily valid for other sedimentary rocks. See Figure 6 for key to facies.

Fig. 7.

Excess silica biplots, after Emmings et al. (2020b). Regression lines for the Ellesmere Port-1 data (red line, with 95% confidence interval) and Emmings et al. (2020b) are also plotted. The excess Si field delineates siliceous mudstones. The field is calibrated for Bowland Shale mudstones and compositionally immature siltstones and is not necessarily valid for other sedimentary rocks. See Figure 6 for key to facies.

Each discrete sample spans c. 2 cm stratigraphic thickness. Taking a local mean sediment accumulation rate (mSAR) of c. 22 mm kyr−1 (Emmings et al. 2020b, after Davies et al. 2004) and 55% compaction (Emmings et al. 2020a), each sample represents a c. 2 kyr duration. This is a first-order estimate since the compaction ratio and mSAR are dependent on sedimentary facies (e.g. Shanmugam and Moiola 1982; Droxler and Schlager 1985; Emmings et al. 2020a). Enrichment factors for discrete XRF analyses support the interpretations based on the core scan XRF dataset. The strongly to weakly carbonate-bearing chemofacies (A–E) are consistently enriched in a suite of redox-sensitive elements, including Se, Mo, U, Zn, V, Ni and Fe (Fig. 8). Se enrichment is a prominent and unusual characteristic of the Bowland Shale that is best explained by ashfall inputs from far-field active volcanic centres (e.g. Brandon et al. 1995; Spears et al. 1999) or derived from weathering of volcanic rocks in the hinterland, followed by trapping under sulfidic conditions (Parnell et al. 2016). Parnell et al. (2022) show that Se, Mo and, to a lesser degree, As are complexed with organic matter and contained within pyrite in the Bowland Shale. By contrast Cu and U, other important redox-sensitive elements, are primarily hosted within recalcitrant (e.g. silicate or refractory kerogen) phases.

Fig. 8.

Major- and trace-element enrichment factors for chemofacies A–G, normalized to the chemofacies F–G (oxic–suboxic) mean. Red lines for the enrichment factor plots are mean curves. Bottom right: δ34Sbulk, δ13Corg and C/N histograms and density curves for chemofacies A–G.

Fig. 8.

Major- and trace-element enrichment factors for chemofacies A–G, normalized to the chemofacies F–G (oxic–suboxic) mean. Red lines for the enrichment factor plots are mean curves. Bottom right: δ34Sbulk, δ13Corg and C/N histograms and density curves for chemofacies A–G.

Chemofacies A limestones are fine grained, lack bioclasts and exhibit high concentrations of elements that proxy for carbonate, including Ca, Mg and Ba (Fig. 8). The limestone beds also lack internal sedimentary structures, suggestive of an early diagenetic origin (e.g. spherulitic limestones; Emmings et al. 2020c). Enrichment in redox-sensitive elements suggests early diagenetic precipitation via anaerobic oxidation of methane (Mazzini et al. 2004), similar to Pendleian spherulitic limestones observed in the Craven Basin (Emmings et al. 2020a). Siderite-cemented beds (chemofacies B) are most commonly interbedded with the claystone facies and exhibit relatively high concentrations of Mg, Fe and Mn, and moderate Ca (Fig. 8), consistent with dolomite and/or siderite cementation. The marine band packages are dominated by bioclastic calcareous mudstones (chemofacies C and D), interpreted as hemipelagic muds interbedded with subordinate calci-turbidites. The phosphatic chemofacies D is rare in core 1 and more common in core 2, consistent with phosphate input or fixation associated with generally more marine and productive conditions during the Arnsbergian (e.g. fish debris, authigenic phosphate). The marine band packages are typically overlain by chemofacies E, F and/or G, interpreted as falling sea-level and/or lowstand deposits.

The low δ34Sbulk and large offset from δ34S of contemporaneous seawater sulfate of c. 13–15‰ (Kampschulte et al. 2001; Kampschulte and Strauss 2004) in chemofacies A–E suggests precipitation of solid sulfides under open system conditions (e.g. Mossmann et al. 1991). In contrast, chemofacies F and G lack enrichment in redox-sensitive elements and exhibit variable, relatively high δ34Sbulk (Fig. 8). High δ34Sbulk in chemofacies F and G is best explained by precipitation of pyrite under closed system (i.e. relatively late diagenetic) conditions and/or quantitative sulfate. Any sulfate in chemofacies F and G is likely to be finely disseminated in the sedimentary matrix. Leached zones along the margins of the conjugate faults at c. 1537.73 m in core 1 (Fig. 5) exhibit δ34Sbulk of +19.8 to +22.6‰; this is best explained by the presence of secondary sulfate that precipitated during the passage of late diagenetic oxidizing fluids.

Mississippian terrestrial OM typically exhibits −23.9‰ (Peters-Kottig et al. 2006) to −25.1‰ δ13Corg (Könitzer et al. 2016), whereas amorphous organic matter (AOM; of probable marine origin) is usually close to c. −29.1‰ (Lewan 1986). δ13Corg exhibits a consistent range between ̶ 27‰ and ̶ 24‰ in all sedimentary facies (Fig. 8), favouring a generally mixed to terrestrial origin for the organic matter (see Hennissen et al. 2024 for further discussion). The C/N ratio is relatively wide ranging (C/N 5–25). The marine, hemipelagic, anoxic chemofacies (A–D) have higher C/N than the oxic, density flow chemofacies (F–G). Considering C/N as a proxy for organic matter type, the C/N ratios are counter-intuitive because cellulose, abundant in C3 and C4 land plants, is N-deficient. Thus, fresh land plant material typically has a C/N >20 whereas marine algal and bacterial sources have a C/N <10 (Meyers 1994). Yet the reverse is observed in the two core sections: chemofacies F–G, comprising abundant Type III terrestrial organic matter (Hennissen et al. 2024), have relatively low C/N, whereas the marine facies have high C/N (Fig. 8). This discrepancy shows that C/N is a poor measure for organic matter type in these rocks (consistent with e.g. Mercuzot et al. 2021 and references therein). Low C/N in argillaceous sediments (chemofacies F–G) is best explained by a quantitative contribution from inorganic (clay–bound) NH4+ (Lindgreen 1994), whereas high C/N in the anoxic facies is consistent with the preferential degradation of nitrogen-rich amino acids under suboxic conditions (Van Mooy et al. 2002).

Finally, dual assessment of S/TOCo and Sr/Ba palaeosalinity proxies (Berner and Raiswell 1984; Wei and Algeo 2020), Mo v. U enrichment factors (Algeo and Tribovillard 2009; Tribovillard et al. 2012) and Mo v. TOCo (Algeo and Lyons 2006) (Fig. 9a–c) provides further support for a weakly to moderately restricted, productivity-driven model for anoxia (Fig. 9d–g). Two samples exhibit S/TOCo <0.1 at 1533.00 and 1536.36 m in core 1 (Figs 5 & 9a), which suggests the development of at least intermittently fully fresh bottom water conditions. The results of palynological analysis (Hennissen et al. 2024) support the sequence stratigraphic framework and model for anoxia. Marine transgressions often show a step-change from (1) proximal or marginal, (sub)oxic shelf edge palynofacies in the underlying lowstand packages into (2) anoxic shelf to basin palynofacies in each marine band package (Figs 4 & 5). The bioturbation analysis of Hennissen et al. (2024) is also consistent with the interpreted palaeoredox cyclicity.

Fig. 9.

Geochemical proxies for bottom water salinity and redox. (a) S/TOC (Berner and Raiswell 1984) and Sr/Ba (Wei and Algeo 2020) palaeosalinity proxies. Fields are bounded by the thresholds for salinity as defined in Wei and Algeo (2020), favouring the more widely established S/TOC proxy. Conflict zones could indicate decoupling between the S/TOC and/or Sr/Ba proxies, likely due to diagenetic effects. Circled data points exhibit inorganic C >0.3 wt%, a local threshold for Sr–bound and/or Ba-associated carbonate. In these cases, the Sr/Ba palaeosalinity proxy is invalid (Wei and Algeo 2020). (b) Mo v. U enrichment factors (EFs) (Algeo and Tribovillard 2009; Tribovillard et al. 2012) suggest fluctuation between oxygenated and weakly to moderately anoxic, intermittently sulfidic particulate shuttle conditions. (c) Mo v. TOC suggests deposition under weakly to moderately restricted bottom water conditions (Algeo and Lyons 2006). See ‘Methods’ for calculation of EFs and derivation of original TOC (TOCo). (d–g) Interpreted redox and salinity response to fourth-order sea-level fluctuation, assuming productivity-driven anoxia in a weakly to moderately restricted basin (sensuEmmings et al. 2020c). The fourth-order cyclicity is partially overprinted by progradation of the Cefn-y-Fedw delta system. AMO, anaerobic oxidation of methane; TOC, total organic carbon.

Fig. 9.

Geochemical proxies for bottom water salinity and redox. (a) S/TOC (Berner and Raiswell 1984) and Sr/Ba (Wei and Algeo 2020) palaeosalinity proxies. Fields are bounded by the thresholds for salinity as defined in Wei and Algeo (2020), favouring the more widely established S/TOC proxy. Conflict zones could indicate decoupling between the S/TOC and/or Sr/Ba proxies, likely due to diagenetic effects. Circled data points exhibit inorganic C >0.3 wt%, a local threshold for Sr–bound and/or Ba-associated carbonate. In these cases, the Sr/Ba palaeosalinity proxy is invalid (Wei and Algeo 2020). (b) Mo v. U enrichment factors (EFs) (Algeo and Tribovillard 2009; Tribovillard et al. 2012) suggest fluctuation between oxygenated and weakly to moderately anoxic, intermittently sulfidic particulate shuttle conditions. (c) Mo v. TOC suggests deposition under weakly to moderately restricted bottom water conditions (Algeo and Lyons 2006). See ‘Methods’ for calculation of EFs and derivation of original TOC (TOCo). (d–g) Interpreted redox and salinity response to fourth-order sea-level fluctuation, assuming productivity-driven anoxia in a weakly to moderately restricted basin (sensuEmmings et al. 2020c). The fourth-order cyclicity is partially overprinted by progradation of the Cefn-y-Fedw delta system. AMO, anaerobic oxidation of methane; TOC, total organic carbon.

Overall, the sedimentological observations, patterns of redox-sensitive metals and semi-metals both in core scan (Figs 4 & 5) and discrete (Fig. 8) samples, S and C stable isotopes, C/N and palynology (Fig. 8; Hennissen et al. 2024) support a model of productivity-driven marine anoxia (Emmings et al. 2020c). Anoxia was probably driven by productive marine phytoplankton and the export of highly labile organic matter into bottom waters, in a weakly to moderately restricted basin (Emmings et al. 2020c). Falling sea-level induced progradation of a freshwater lid into the Blacon Basin, a process that led to the shut–down in photic zone marine productivity and consequent basin ventilation, while generally sustaining normal marine or moderate salinities in bottom waters. Thus, the palaeosalinity record often lags the palaeoredox signal (Fig. 9). Based on S/TOCo, bottom waters were fully desalinated only once, during a well-developed lowstand in the upper part of the Kinderscoutian section (core 1) (Fig. 5).

Three kNN machine learning algorithms were deployed to predict siltstone and sandstone (chemofacies G), claystone and siderite (chemofacies F, B), calcareous mudstone (chemofacies C, D and E) and limestone (chemofacies A) through the Bowland Shale Formation in Ellesmere Port-1 (Fig. 10). Importantly, differentiation of non-calcareous and calcareous mudstone facies helps to interpret the succession in the context of syngenetic processes (Fig. 9d–g) and, ultimately, sequence stratigraphy. Based on these predictions, the Bowland Shale Formation comprises c. 18% siltstone and sandstone, 37% non-calcareous mudstone, 42% calcareous mudstone and 3% limestone. The unit comprises 12 complete fourth-order cycles, and is interfingered with the three distal Lower, Middle and Upper units (leaves) of the Cefn-y-Fedw Sandstone Formation (Davies et al. 2004; Waters et al. 2021).

Fig. 10.

The results of three KNN predictive chemofacies models. (a) Chemofacies were derived from principal component analysis and hierarchical cluster analysis of selected XRF core scan elements, simplified (aggregated) and matched to the nearest wireline datum. The models were trained and deployed by matching the XRF-derived chemofacies to selected wireline logs. kNN (1–3) represent different approaches (see ‘Methods’). (b) The trained models were deployed on the observed Bowland Shale section at Ellesmere Port-1 (1443 to 1767 m below surface) (kNN method 2 is displayed). Prominent sand packages probably correspond to lower, middle and upper extensions of the Cefn-y-Fedw Sandstone (CSF) (Waters et al. 2021) coincident with three periods of regional channel incision (Waters and Condon 2012). Biostratigraphic ammonoid zones observed in core (bold), other inferred major flooding events and biozones (italicized, from Davies et al. 2011) and interpreted substages and extrapolated ages (Waters and Condon 2012) are also indicated.

Fig. 10.

The results of three KNN predictive chemofacies models. (a) Chemofacies were derived from principal component analysis and hierarchical cluster analysis of selected XRF core scan elements, simplified (aggregated) and matched to the nearest wireline datum. The models were trained and deployed by matching the XRF-derived chemofacies to selected wireline logs. kNN (1–3) represent different approaches (see ‘Methods’). (b) The trained models were deployed on the observed Bowland Shale section at Ellesmere Port-1 (1443 to 1767 m below surface) (kNN method 2 is displayed). Prominent sand packages probably correspond to lower, middle and upper extensions of the Cefn-y-Fedw Sandstone (CSF) (Waters et al. 2021) coincident with three periods of regional channel incision (Waters and Condon 2012). Biostratigraphic ammonoid zones observed in core (bold), other inferred major flooding events and biozones (italicized, from Davies et al. 2011) and interpreted substages and extrapolated ages (Waters and Condon 2012) are also indicated.

Using the ammonoid biozonation in the two core sections together with the known age ranges of each of the distal turbidite leaves of the Cefn-y-Fedw Sandstone (Davies et al. 2004), it is possible to infer the key index regional marine bands (Davies et al. 2011) in the Blacon Basin. Like the observations in the core sections, generally sharp transitions from non-calcareous to calcareous mudstones are interpreted to indicate deposition during marine transgressions. Eight of the 12 cycles exhibit resolvable thin limestone beds, which probably coincide with the maximum rate of transgression and/or early part of the sea-level highstand. Thus, the Blacon Basin apparently preserves a largely complete Namurian succession, analogous to the Craven Basin (Saunders and Ramsbottom 1986). In all cycles, the calcareous mudstones are progressively or abruptly replaced with non-calcareous mudstones and/or siltstone to sandstone facies. We interpret these shifts from calcareous to siliciclastic facies to represent deposition during falling sea-level, culminating in lowstands. The abrupt transitions from calcareous to non-calcareous mudstones are analogous to observations in core (Figs 4 & 5): metre-scale cohesive debris flows shed from the basin margins. The Lower, Middle and Upper units of the Cefn-y-Fedw Sandstone Formation plausibly occupy the lowstands of multiple cycles and are potentially associated with erosional basal surfaces that cut into the underlying strata. We interpret deposition of the Middle Cefn-y-Fedw Sandstone unit to span and/or to have incised through the entirety of the Alportian in the Blacon Basin. Thus, the mass transport deposit underlying the H1a1 marine band in core 2 (Fig. 4) probably represents a precursor or the first pulse of the Middle Cefn-y-Fedw Sandstone unit. Similarly, the siltstone facies at the top of the Kinderscoutian section (core 1) is interpreted as the base of the Upper Cefn-y-Fedw Sandstone unit (Fig. 5).

We integrated the sequence stratigraphic interpretations, based on the core scan observations and machine learning, with observations from other key wells in the Blacon Basin (Fig. 11). The results constrain the sedimentary architecture in the Blacon Basin. Ellesmere Port-1 is not located in the deepest part of the basin; it is located on the NE flank of the basin. Taking a basin-wide perspective shows the importance and influence of the Cefn-y-Fedw delta system; except for the relatively thin marine band facies, the Bowland Shale in the Blacon Basin is relatively expanded and totally dominated by prodeltaic mud (Fig. 11). This is consistent with the dominance of Type III organic matter (including large plant fragments) in both core sections (Hennissen et al. 2024). This contrasts with the Bowland Shale in the Craven Basin; whilst also supplied by a large delta system, the succession in the Craven Basin represents deposition from a more balanced mixture of competing hemipelagic (biogenic) and allochthonous components (Clarke et al. 2018; Newport et al. 2018; Emmings et al. 2020a, b, c). The relatively expanded, clay-rich and poorly cemented mudstone succession (Fig. 6) is also in contrast to the economically important, broadly time-equivalent Barnett Shale Formation (Loucks and Ruppel 2007). Ultimately, the structurally complex, collisional intercontinental setting explains why many of the US gas shales are poor analogues for the Bowland Shale. Structural complexity offers a direct challenge to unconventional hydrocarbon extraction from the Bowland Shale (Anderson and Underhill 2020), but is also implicated in the compartmentalization of key cemented lithofacies and the encroachment of delta systems linked to an uplifting hinterland.

Fig. 11.

Summary of interpreted Namurian sedimentary architecture in the Blacon Basin, with major flooding events of Davies et al. (2011) and general stratigraphy after Magraw and Ramsbottom (1957), Davies et al. (2004), Smith et al. (2005), Waters et al. (2009, 2021) and Andrews (2013). The shelf-edge delta wedge of the Late Pleistocene Po River (Adriatic Sea, offshore central Italy) may represent a partial (albeit short-lived) modern analogue (Pellegrini et al. 2017; Patruno and Helland-Hansen 2018). Gamma-ray well logs for Croxteth-1, Ellesmere Port-1 (this study), Blacon East-1 and Milton Green-1 are shown for reference. The line of section is flattened and truncated on a Yeadonian marine flooding surface (e.g., G1 marine band) within the Gwespyr Sandstone. See Figure 1 for line of section. MB, marine band.

Fig. 11.

Summary of interpreted Namurian sedimentary architecture in the Blacon Basin, with major flooding events of Davies et al. (2011) and general stratigraphy after Magraw and Ramsbottom (1957), Davies et al. (2004), Smith et al. (2005), Waters et al. (2009, 2021) and Andrews (2013). The shelf-edge delta wedge of the Late Pleistocene Po River (Adriatic Sea, offshore central Italy) may represent a partial (albeit short-lived) modern analogue (Pellegrini et al. 2017; Patruno and Helland-Hansen 2018). Gamma-ray well logs for Croxteth-1, Ellesmere Port-1 (this study), Blacon East-1 and Milton Green-1 are shown for reference. The line of section is flattened and truncated on a Yeadonian marine flooding surface (e.g., G1 marine band) within the Gwespyr Sandstone. See Figure 1 for line of section. MB, marine band.

The stratigraphic analysis delineates an inflection point immediately above the Middle Cefn-y-Fedw Sandstone turbidite package at c. 1615 m below surface, at or near the base of the Kinderscoutian (Fig. 10). The Pendleian and Arnsbergian cycles exhibit relatively low temporal frequency and a relatively low mSAR (e.g. condensed) and are more strongly asymmetrical, which is suggestive of deposition under the combined influence of eustatic sea-level fluctuation and basin subsidence (e.g. Martinsen et al. 1995). Waters et al. (2021) showed that, during the Pendleian–Arnsbergian, detrital sediments entered the Blacon Basin via the Cefn-y-Fedw delta systems in the SW (Fig. 11). Sediment flows were potentially aligned parallel to the long axis of the combined Blacon and Widnes basins (Fig. 1), analogous to the Gainsborough Trough (Palci et al. 2020). Whilst the age of the Middle Cefn-y-Fedw Sandstone turbidite package is poorly constrained (Waters et al. 2021), it: (1) is potentially Chokierian, Alportian and/or earliest Kinderscoutian in age, and therefore must span or juxtapose one or more significant hiatuses; (2) overlies the H1a1 biozone, and therefore probably occurs within the mid-Carboniferous (Mississippian–Pennsylvanian) eustatic event (Saunders and Ramsbottom 1986); and (3) is succeeded by a return to mud-dominated deposition in the early Kinderscoutian. Deposition of sand and the presence of hiatuses is consistent with the development of protracted low eustatic sea-level during the mid-Carboniferous event, which potentially shifted the main depocentres into deeper basins (e.g. the Craven Basin in the Irish Sea). Thus, the Middle Cefn-y-Fedw Sandstone could represent a turbidite lobe deposited during third-order sea-level regression associated with the mid-Carboniferous eustatic event, although the role of local controls such as fault reactivation is unresolved.

The lithostratigraphic analysis shows that the Blacon Basin became re-established as a major depocentre for mud during the Kinderscoutian. The well correlation together with heavy mineral analysis of Waters et al. (2021) suggests the Blacon Basin continued to receive detrital sediment from a single source via the Cefn-y-Fedw delta system, located in the south or SW. The Kinderscoutian to Yeadonian cycles apparently exhibit higher temporal frequency. Additionally, the Kinderscoutian–Yeadonian succession was deposited under a relatively high mSAR and the cycles are slightly more symmetrical, which probably reflects increasing proximity to the Cefn-y-Fedw delta front (Fig. 11). By the Yeadonian, sediment supply into the Blacon Basin became more complex, with provenances from the SW, SE, NW and NE (Waters et al. 2021).

RHP calculations for the Bowland Shale in borehole Ellesmere Port-1 indicate bulk RHPs of 2.6 ± 1.3 μW m−3 (Fig. 12). Similarly high RHPs are also observed in the Bowland Shale in the Craven Basin. The Bowland Shale RHPs are significantly higher than the global shale mean of c. 1.5 μW m−3 reported by Vilà et al. (2010), and approach the RHPs of granitic intrusions, targeted for geothermal energy in the UK (Gluyas et al. 2018). RHP derives from the decay of U, K and Th, in descending order of importance (Rybach 1976). The U, K and Th data (Fig. 12c) indicate that the high RHPs are primarily due to relatively high U concentrations, whereas Th and K concentrations are broadly in line with the global average. The U contents are relatively but not absolutely high; for example, the Alum Shale (Sweden) exhibits U concentrations between 100 and 300 ppm (Lecomte et al. 2017), an order of magnitude greater than those found in the Bowland Shale.

Fig. 12.

Comparison of RHPs and U, Th and K concentrations for a range of rock types. (a) Calculated RHPs through the Bowland Shale based on the spectral gamma-ray well log data superimposed on the machine predictions. (b, c) RHPs and U, Th and K concentrations for limestone, claystone, calcareous mudstone and siltstone–sandstone facies in the Bowland Shale in Ellesmere Port-1, based on spectral gamma-ray and discrete XRF analysis (the latter indicated by asterisks). The UK RHP sedimentary mean (Rollin 2002; Busby and Terrington 2017) and RHPs for the Bowland Shale in the Craven Basin (Newport 2019; Emmings et al. 2020c), the Kimmeridge Clay (Elm Tree Farm, Yorkshire, J.F.E. unpublished borehole data/analyses), Holywell Shale (Newport et al. 2016) and italicized baseline data (Vilà et al. 2010) are also shown.

Fig. 12.

Comparison of RHPs and U, Th and K concentrations for a range of rock types. (a) Calculated RHPs through the Bowland Shale based on the spectral gamma-ray well log data superimposed on the machine predictions. (b, c) RHPs and U, Th and K concentrations for limestone, claystone, calcareous mudstone and siltstone–sandstone facies in the Bowland Shale in Ellesmere Port-1, based on spectral gamma-ray and discrete XRF analysis (the latter indicated by asterisks). The UK RHP sedimentary mean (Rollin 2002; Busby and Terrington 2017) and RHPs for the Bowland Shale in the Craven Basin (Newport 2019; Emmings et al. 2020c), the Kimmeridge Clay (Elm Tree Farm, Yorkshire, J.F.E. unpublished borehole data/analyses), Holywell Shale (Newport et al. 2016) and italicized baseline data (Vilà et al. 2010) are also shown.

The relatively high U in Carboniferous sedimentary rocks is additionally important because U can become mobile in groundwater or produced waters (Smedley et al. 2006), and the decay of U contributes to Rn risk areas where the rocks crop out (Miles et al. 2007). However, it should be noted the vast majority of UK groundwaters exhibit U concentrations far below the World Health Organization provisional guideline value for U in drinking water (15 μg l−1) (Smedley et al. 2006). Additionally, the relatively high U contents in UK groundwaters commonly derive from sandstone red-beds, rather than Carboniferous shales, and U mobility is dependent on the mineral host and speciation (e.g. Smedley et al. 2006; Phan et al. 2015). From a regional perspective, U in the Bowland Shale is primarily hosted in recalcitrant mineral or organic phases (Parnell et al. 2022, see below) and is therefore likely to be relatively immobile.

The relatively high U contents observed in the Bowland Shale are potentially explained by three factors. Firstly, high U concentrations are a general feature of Devonian–Carboniferous shales globally (Mehra et al. 2021), suggesting a first-order, fundamental change in the availability and/or fixation pathway of U in marine sediments during this time interval. Secondly, a productivity-driven model for anoxia and high mSARs related to proximity to multiple delta systems probably promoted under-consolidation of seabed sediments during deposition of the Bowland Shale. Since U is primarily fixed within porewaters, rather than bottom waters (Tribovillard et al. 2006), under-consolidation could plausibly promote U drawdown due to generally enhanced shallow sediment porosity and diffusivity with bottom waters. Finally, the siltstone and sandstone facies in the Bowland Shale also exhibit high RHPs above the global sandstone average (Fig. 12). This strongly suggests an important role of provenance; granitic rocks were present in the hinterland (Waters et al. 2021) and the siltstones in Ellesmere Port-1 are compositionally immature, consistent with a granitic provenance. Specifically, silt and sand in the Bowland Shale is enriched in a suite of detrital minerals that are important hosts of U in granite (Nash 1979), including heavy minerals, mica and especially detrital magnetite (Emmings et al. 2020c).

We modelled a range of scenarios using the heat flow box model equations of Beardsmore et al. (2010), assuming complete isotropy and negligible convective heat transport. Conductive heat flow is controlled by two parameters (Beardsmore et al. 2010): in situ RHP (Rybach 1976) and the thermal conductivity of constituent phases. Carbonates, clay minerals and unsaturated or saturated pores are thermal insulators, whereas pyrite and quartz are important thermal conductors (Midttømme et al. 1998; Midttømme and Roaldset 1999). The distribution of these phases in black shales is synergistically linked to redox conditions during deposition and diagenesis and is therefore predictable (Fig. 9) (Emmings et al. 2020b, c).

The models show that heat flow is highly sensitive to thermal conductivity and is minimally affected by RHP (supplementary material B, see following discussion). Thus, black (‘hot’) shales are unlikely to represent a credible, direct target for geothermal energy. However, the Bowland Shale could act as an effective thermal blanket to underlying geothermal reservoirs (e.g. Nunn et al. 1999). This effect is potentially manifested as the Bowland Forest heat anomaly described by Lee et al. (1987), although the anomaly is not identified in more recent mapping (e.g. Busby 2014). Maintaining the same arbitrary surface heat flow (50 W m−2) and RHP (3 μW m−3), box modelling shows that the base of a 1 km-thick claystone unit is c. 17°C warmer than at the base of a 1 km-thick siliceous, pyritic mudstone unit. This is because clay minerals are insulators whereas quartz and pyrite are conductors of heat. In these models, we assumed the same in situ RHP of 3 μW m−3 (Fig. 12b), which results in a relatively small excess temperature gradient of 0.75°C km−1. However, adopting an in situ RHP of 10 μW m−3 that is more comparable to high-RHP granite (Vilà et al. 2010) imparts an excess temperature gradient of 2°C km−1. Thus, RHP becomes quantitively important where it is consistently high in a large rock volume (e.g. granite batholiths), a feature that is unlikely to be observed in sedimentary basins because it requires the presence of extremely thick metalliferous black shale packages. Whilst these modelling exercises are not entirely realistic, this approach helps to demonstrate the importance of understanding mudstone lithofacies as a control on subsurface heat flow.

In another exercise, we fitted two different, similarly accurate (but not necessarily realistic) temperature gradient models through the Bowland Shale in Ellesmere Port-1 (Fig. 13; supplementary material B). Whilst the measured borehole temperatures in Ellesmere Port-1 are unlikely to be equilibrated with the surrounding rock, the data delineate the key inflection at c. 1615 m interpreted as related to the mid-Carboniferous eustatic event and delta progradation (see prior discussion). This inflection differentiates geothermal gradients of c. 25°C km−1 and 18°C km−1, respectively. Model 1 is based on a single, homogeneous, low-conductivity claystone package with low RHP (1.5 μW m−3), 5% total porosity (Clarke et al. 2018), 80% clay minerals, 10% quartz and 5% carbonate minerals. Model 1 requires relatively low heat flows of 62 and 42 mW m−2, above and below 1615 m below surface, respectively, to match the observed borehole temperature profile. Model 2 is based on the calculated RHPs and machine predicted thicknesses and composition of each discrete package within the Bowland Shale (Figs 10 & 13), and is therefore considered more realistic. In model 2, thermal conductivities are higher in all facies, which necessitate higher heat flows of 75 and 55 mW m−2, above and below 1615 m below surface, respectively. Thus, based on the observations coupled to modelling, the Bowland Shale in Ellesmere Port-1 is not necessarily an effective thermal ‘blanket’ since it comprises c. 18% siltstone and sandstone, and instead may conduct heat at a faster rate (i.e. high thermal diffusivity) than pure claystones or condensed organic-rich shales. Again, these models are likely to be overly simplistic but demonstrate the importance of understanding the fractions of different mudstone facies within a given ‘shale’ package. The models do not consider the effect of organic matter (particularly coal) and are limited to the classes (lithologies) observed in the two cored sections. In particular, coals were not observed in either core section, but coal exhibits a very low thermal conductivity (Nunn and Lin 2002) and coal seams are present in the Bowland Shale in North Wales (Davies et al. 2004).

Fig. 13.

Measured v. modelled temperature gradients through the Bowland Shale Formation in Ellesmere Port-1. It should be noted the borehole temperatures are unlikely to represent true, equilibrated values and therefore probably represent minimum estimates.

Fig. 13.

Measured v. modelled temperature gradients through the Bowland Shale Formation in Ellesmere Port-1. It should be noted the borehole temperatures are unlikely to represent true, equilibrated values and therefore probably represent minimum estimates.

The Blacon Basin, in addition to many other sedimentary basins, is filled by hundreds of metres of Carboniferous mudstone. Important UK geothermal plays include granites (Gluyas et al. 2018) and karstified Carboniferous limestones (Busby 2014). Carboniferous limestone–mudstone cyclothems cap granites on several blocks (e.g. proven in the Raydale and Rookhope boreholes; Dunham and Wilson 1985) and cap lower Carboniferous limestones in the UK (e.g. Fig. 10; Dunham and Wilson 1985; Busby 2014). Additionally, basinal Carboniferous mudstones may cap granites beneath local intrabasin highs (Bott 1988; Donato 2020). Yet, our ability to explore for deep geothermal energy in sedimentary basins is limited by a lack of understanding of how different types of shales control heat flow. Efficient heat diffusion is also important for the safe geological storage of radioactive waste (Rutqvist 2020), and the design of heat pumps is dependent on the thermal properties of the storage medium (Busby et al. 2009; Parkes et al. 2021). Simple box modelling together with previous work (e.g. Midttømme et al. 1998) shows that the efficiency of the sedimentary thermal blanket is highly dependent on mudstone mineralogy and, by extension, lithofacies. Therefore, future work should seek to: (1) measure the key thermal properties for a range of mudstone microfacies; (2) better understand the role of depositional processes, basin type and age as controls on mudstone facies coupled to thermal conductivity; (3) constrain the effect of diagenesis on mudstone thermal properties; (4) link the thermal properties of mudstones to one or more holistic, predictive frameworks (e.g. sequence stratigraphy); and (5) investigate the potential for convective fluid flow through fracture networks, which are more likely to be developed within relatively brittle mudstone microfacies (e.g. siliceous and carbonate-cemented marine band facies).

We conducted a high-resolution multi-disciplinary analysis of two core sections in the borehole Ellesmere Port-1, Cheshire, UK. Biostratigraphic analysis indicates the core sections are Kinderscoutian and late Arnsbergian–Chokierian in age, respectively. Both cores are assigned to the Bowland Shale Formation. Coupled core scan and discrete geochemical analysis enables interpretation of syngenetic processes at a relatively high temporal resolution (c. 1–2 kyr). Both cores comprise cycles of limestone, calcareous to non-calcareous mudstone and siltstone that are characteristic of the Bowland Shale in the UK. The cycles are interpreted to represent sediment deposition during fourth-order, glacioeustatic sea-level fluctuation. Palaeoredox and palaeosalinity geochemical proxies support a model for productivity–driven anoxia in a weakly to moderately restricted basin. Coupled Mo/Al and As/Al core scan ratios support a model for intermittently euxinic bottom waters overlain by relatively thick ferruginous intermediate waters. Falling sea-level led to waning marine productivity; euxinic bottom waters were progressively replaced with ferruginous waters, followed by the development of fully suboxic or oxic conditions during the late falling stage and lowstands.

Unlike the Pendleian Bowland Shale in the Craven Basin, the cored Bowland Shale successions in Ellesmere Port-1 are not siliceous and are instead dominated by argillaceous mudstones. Marine band facies are interpreted to represent deposition during marine transgressions and highstands. Falling sea-level is linked to the progressive or abrupt replacement of calcareous mudstones with siliciclastic (non–carbonate-bearing) claystones. The siliciclastic mudstones are commonly laminated and often exhibit chaotic sedimentary structures, suggestive of deposition from a range of density flow types, including potential for en masse transport within debris flows. Laminated siltstone and sandstone packages occasionally cap the siliciclastic claystones and are interpreted to represent deposition during lowstands, where bottom water conditions were unambiguously oxygenated, brackish and at least intermittently freshwater.

Cluster analysis of the XRF core scan data coupled to machine learning of the well log data enabled prediction of limestone, calcareous mudstone, siliciclastic (non-carbonate) claystone, and siltstone and sandstone facies through the entire Bowland Shale interval in Ellesmere Port-1. The machine predictions show that the Bowland Shale is interfingered with the three turbiditic leaves of the Cefn-y-Fedw Sandstone and contains 12 complete fourth-order cycles resolvable in the well log data. The Bowland Shale comprises c. 18% siltstone and sandstone, 37% non-calcareous mudstone, 42% calcareous mudstone and 3% limestone. A step-change in the symmetry and thickness of cycles at c. 1615 m below surface probably corresponds to an important regional shift in the rate of sediment supply from the Cefn-y-Fedw delta system, above the mid-Carboniferous boundary.

The Bowland Shale exhibits high RHP in comparison to other sedimentary rocks, due primarily to relative enrichment in U. Despite this, box model heat modelling calibrated to borehole temperature measurements show that heat flow is minimally influenced by the high RHP in the Bowland Shale. Modelling suggests that it is not necessarily appropriate to assume that the Bowland Shale acts an insulating blanket over subsurface heat flow. Simple box modelling shows that the efficiency of the sedimentary thermal blanket is highly dependent on mudstone mineralogy and facies. The Bowland Shale in the Blacon Basin is relatively expanded and does not represent a highly condensed black shale unit. Based on the compositional (XRF) analysis, the Bowland Shale microfacies in the UK are likely to be more conductive than a simple claystone lithology. Considering the potential importance of Carboniferous mudstones as caprocks to buried granite and limestone geothermal plays in the UK, further work should: (1) measure the thermal properties of a range of mudstone microfacies; and (2) link the findings to established depositional and lithostratigraphic models to better understand the extent to which mudstone units act as thermal blankets.

We acknowledge the National Geoscience Data Centre (NGDC), British Geological Survey for enabling access to the Ellesmere Port-1 core materials. We acknowledge Mike Stephenson (formerly BGS) for helping to facilitate the research project. We thank Simon Harris (BGS) and Sarah Davies (University of Leicester) for facilitating access to laboratory facilities at the NGDC and the University of Leicester XRF laboratory, respectively. Contains data supplied by permission of the Natural Environment Research Council [2022]. Contains information provided by the Oil and Gas Authority and/or other third parties. We are grateful to the reviewers for helping to improve this manuscript through their valuable comments.

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

JFE: conceptualization (lead), data curation (lead), formal analysis (lead), methodology (lead), visualization (lead), writing – original draft (lead), writing – review & editing (lead); JAIH: data curation (lead), formal analysis (lead), methodology (lead), visualization (supporting), writing – original draft (supporting); CHV: formal analysis (supporting), resources (supporting), supervision (supporting); MD: data curation (lead), formal analysis (lead), resources (supporting), writing – original draft (supporting); LM: formal analysis (lead); VM-H: formal analysis (supporting); AL: formal analysis (lead), writing – original draft (supporting); JL: formal analysis (lead), writing – original draft (supporting); ML: funding acquisition (supporting), resources (supporting), supervision (supporting); NJR: formal analysis (supporting).

This research was primarily funded by a British Geological Survey (UKRI) National Capability grant NEE6640S (to JFE). JAIH is supported by the Natural Environment Research Council through the Unconventional Hydrocarbons in the UK Energy System Programme (UKUH) Challenge 2 (NE/R017964/1). Stable C isotope analyses were funded by the NEIF Steering Committee [grant no. IP-1918-0619] (to MJL).

All discrete sample geochemical datasets (Rock-Eval pyrolysis, XRF elemental analysis, stable C and S isotopes) are reported in the Supplementary material. The XRF core scan dataset is archived and openly available via the NDGC (BGS Core Scanning Facility 2022). The Ellesmere Port-1 cores are openly available via the National Geoscience Data Centre (NGDC), British Geological Survey. Well log data are available via the National Data Repository (https://www.nstauthority.co.uk/data-centre/access-to-information-and-samples/). The palynology and vitrinite reflectance datasets are reported by Hennissen et al. (2024).

1.
Aitchison
J.
1986
.
The Statistical Analysis of Compositional Data
 .
Chapman & Hall
,
London
.
2.
Algeo
T.J.
2004
.
Can marine anoxic events draw down the trace element inventory of seawater?
Geology
 ,
32
,
1057
1060
, https://doi.org/10.1130/G20896.1
3.
Algeo
T.J.
and
Lyons
T.W.
2006
.
Mo–total organic carbon covariation in modern anoxic marine environments: implications for analysis of paleoredox and paleohydrographic conditions
.
Paleoceanography and Paleoclimatology
 ,
21
, https://doi.org/10.1029/2004PA001112
4.
Algeo
T.J.
and
Tribovillard
N.
2009
.
Environmental analysis of paleoceanographic systems based on molybdenum–uranium covariation
.
Chemical Geology
 ,
268
,
211
225
, https://doi.org/10.1016/j.chemgeo.2009.09.001
5.
Algeo
T.J.
,
Heckel
P.H.
,
Maynard
J.B.
,
Blakey
R.
and
Rowe
H.
2008
.
Modern and ancient epeiric seas and the super-estuarine circulation model of marine anoxia
.
GSA Special Papers
 ,
48
,
7
38
.
6.
Anderson
I.
and
Underhill
J.R.
2020
.
Structural constraints on Lower Carboniferous shale gas exploration in the Craven Basin, NW England
.
Petroleum Geoscience
 ,
26
,
303
324
, https://doi.org/10.1144/petgeo2019-125
7.
Andrews
I.J.
2013
.
The Carboniferous Bowland Shale Gas Study: Geology and Resource Estimation
 .
British Geological Survey for Department of Energy and Climate Change
.
8.
Arthurton
R.S.
1984
.
The Ribblesdale fold belt, NW England – a Dinantian–early Namurian dextral shear zone
.
Geological Society, London, Special Publications
 ,
14
,
131
138
, https://doi.org/10.1144/GSL.SP.1984.014.01.13
9.
Auguie
B.
2016
.
gridExtra R package version 2.2.1
, https://CRAN.R-project.org/package=gridExtra
10.
Baines
J.G.
1977
.
The Stratigraphy and Sedimentology of the Skipton Moor Grits (Namurian E1c) and their Lateral Equivalents
 . PhD thesis,
University of Keele
.
11.
Beardsmore
G.
,
Rybach
L.
,
Blackwell
D.
and
Baron
C.
2010
.
A protocol for estimating and mapping global EGS potential
.
Geothermal Resources Council Transactions
 ,
34
,
301
312
.
12.
Berner
R.A.
and
Raiswell
R.
1984
.
C/S method for distinguishing freshwater from marine sedimentary rocks
.
Geology
 ,
12
,
365
368
, https://doi.org/10.1130/0091-7613(1984)12<365:CMFDFF>2.0.CO;2
13.
BGS Core Scanning Facility
2022
. Ellesmere Port-1 Core Scanning Dataset.
NERC EDS National Geoscience Data Centre
.
14.
Bloxam
T.W.
and
Thomas
R.L.
1968
.
Palaeontological and geochemical facies in the Gastrioceras subcrenatum marine-band and associated rocks from the North Crop of the South Wales Coalfield
.
Quarterly Journal of the Geological Society
 ,
124
,
239
277
, https://doi.org/10.1144/gsjgs.124.1.0239
15.
Bott
M.H.P.
1988
.
The Market Weighton gravity anomaly; granite or graben?
Proceedings of the Yorkshire Geological Society
 ,
47
,
47
53
, https://doi.org/10.1144/pygs.47.1.47
16.
Bouch
J.E.
,
Naden
J.
et al
2006
.
Direct evidence of fluid mixing in the formation of stratabound Pb–Zn–Ba–F mineralisation in the Alston Block, North Pennine Orefield (England)
.
Mineralium Deposita
 ,
41
,
821
835
, https://doi.org/10.1007/s00126-006-0093-3
17.
Brandon
A.
,
Riley
N.J.
,
Wilson
A.A.
and
Ellison
R.A.
1995
.
Three new early Namurian (E1c-E2a) marine bands in central and northern England, UK, and their bearing on correlations with the Askrigg Block
.
Proceedings of the Yorkshire Geological and Polytechnic Society
 ,
50
,
333
355
, https://doi.org/10.1144/pygs.50.4.333
18.
Brogan
J.C.
,
Fleming
G.A.
and
Byrne
J.E.
1973
.
Molybdenum and copper in Irish pasture soils
.
Irish Journal of Agricultural Research
 ,
12
,
71
81
.
19.
Busby
J.
2014
.
Geothermal energy in sedimentary basins in the UK
.
Hydrogeology Journal
 ,
22
,
129
141
, https://doi.org/10.1007/s10040-013-1054-4
20.
Busby
J.
and
Terrington
R.
2017
.
Assessment of the resource base for engineered geothermal systems in Great Britain
.
Geothermal Energy
 ,
5
,
7
, https://doi.org/10.1186/s40517-017-0066-z
21.
Busby
J.
,
Lewis
M.
,
Reeves
H.
and
Lawley
R.
2009
.
Initial geological considerations before installing ground source heat pump systems
.
Quarterly Journal of Engineering Geology and Hydrogeology
 ,
42
,
295
306
, https://doi.org/10.1144/1470-9236/08-092
22.
Church
K.D.
and
Gawthorpe
R.L.
1994
.
High resolution sequence stratigraphy of the late Namurian in the Widmerpool Gulf (East Midlands, UK)
.
Marine and Petroleum Geology
 ,
11
,
528
544
, https://doi.org/10.1016/0264-8172(94)90066-3
23.
Clarke
H.
,
Turner
P.
,
Bustin
R.M.
,
Riley
N.
and
Besly
B.
2018
.
Shale gas resources of the Bowland Basin, NW England: a holistic study
.
Petroleum Geoscience
 ,
24
,
287
322
, https://doi.org/10.1144/petgeo2017-066
24.
Clarke
H.
,
Verdon
J.P.
,
Kettlety
T.
,
Baird
A.F.
and
Kendall
J.M.
2019
.
Real-time imaging, forecasting, and management of human-induced seismicity at Preston New Road, Lancashire, England
.
Seismological Research Letters
 ,
90
,
1902
1915
, https://doi.org/10.1785/0220190110
25.
Cózar
P.
and
Somerville
I.D.
2021
.
The Serpukhovian in Britain: use of foraminiferal assemblages for dating and correlating
.
Journal of the Geological Society, London
 ,
178
, https://doi.org/10.1144/jgs2020-170
26.
Davies
J.R.
,
Wilson
D.
and
Williamson
I.T.
2004
.
Geology of the Country around Flint
 .
Memoirs of the British Geological Survey, Sheet 108 (England and Wales)
.
27.
Davies
J.R.
,
Somerville
I.D.
,
Waters
C.N.
and
Jones
N.S.
2011
. North Wales. In:
Waters
C.N.
,
Somerville
I.D.
et al
(eds)
A Revised Correlation of Carboniferous Rocks in the British Isles
 .
Geological Society, London
.
28.
Davies
R.J.
,
Austin
R.
and
Moore
D.
1993
.
Environmental controls of Brigantian conodont-distribution: evidence from the Gayle limestone of the Yoredale Group in Northern England
.
Annales de la Société géologique de Belgique
 ,
116
,
221
241
.
29.
Davies
S.J.
2008
.
The record of Carboniferous sea-level change in low-latitude sedimentary successions from Britain and Ireland during the onset of the late Paleozoic ice age
.
GSA Special Papers
 ,
441
,
187
204
, https://doi.org/10.1130/2008.2441(13)
30.
Davies
S.
,
Hampson
G.
,
Flint
S.
and
Elliott
T.
1999
.
Continental-scale sequence stratigraphy of the Namurian, Upper Carboniferous and its applications to reservoir prediction
.
Geological Society, London, Petroleum Geology Conference Series
 ,
5
,
757
770
, https://doi.org/10.1144/0050757
31.
Davydov
V.I.
,
Korn
D.
and
Schmitz
M.
2012
. The Carboniferous Period. In:
Gradstein
F.M.
,
Ogg
J.G.
,
Schmitz
M.D.
and
Ogg
G.M.
(eds)
The Geologic Time Scale 2012
 .
Elsevier
,
603
651
, https://doi.org/10.1016/B978-0-444-59425-9.00023-8
32.
Dean
M.T.
,
Browne
M.A.E.
,
Waters
C.N.
and
Powell
J.H.
2011
.
A lithostratigraphical framework for the Carboniferous successions of northern Great Britain (Onshore)
.
British Geological Survey Research Report RR/10/07
.
33.
DECC
2013
.
The Hydrocarbon Prospectivity of Britain's Onshore Basins
 .
Department for Energy and Climate Change
.
34.
Dmitrijeva
M.
,
Cook
N.J.
et al
2020
.
Multivariate statistical analysis of trace elements in pyrite: prediction, bias and artefacts in defining mineral signatures
.
Minerals
 ,
10
,
61
, https://doi.org/10.3390/min10010061
35.
Donato
J.
2020
.
Gravity Modelling Across Two Postulated Granite Batholiths within the UK Onshore East Midlands Shelf
 .
University of Oxford/UKOGL Beneath Britain
.
36.
Dowey
P.J.
and
Taylor
K.G.
2020
.
Diagenetic mineral development within the Upper Jurassic Haynesville-Bossier Shale, USA
.
Sedimentology
 ,
67
,
47
77
, https://doi.org/10.1111/sed.12624
37.
Dowle
M.
and
Srinivasan
A.
2020
.
data.table R package version 1.13.2
, https://CRAN.R-project.org/package=data.table
38.
Droxler
A.W.
and
Schlager
W.
1985
.
Glacial v. interglacial sedimentation rates and turbidite frequency in the Bahamas
.
Geology
 ,
13
,
799
802
, https://doi.org/10.1130/0091-7613(1985)13<799:GVISRA>2.0.CO;2
39.
Dunham
A.C.
and
Wilson
A.A.
1985
.
Geology of the Northern Pennine Orefield: Volume 2, Stainmore to Craven
 .
Economic Memoir of the British Geological Survey
.
40.
Ellis
D.
and
Singer
J.
2007
.
Well Logging for Earth Scientists
 .
Springer
,
Dordrecht
.
41.
Emmings
J.F.
2018
.
Controls on UK Lower Namurian Shale Gas Prospectivity: Understanding the Spatial and Temporal Distribution of Organic Matter in Siliciclastic Mudstones
 . PhD thesis,
University of Leicester
.
42.
Emmings
J.F.
,
Hennissen
J.A.I.
et al
2019
.
Controls on amorphous organic matter type and sulphurization in a Mississippian black shale
.
Review of Palaeobotany and Palynology
 ,
268
,
1
18
, https://doi.org/10.1016/j.revpalbo.2019.04.004
43.
Emmings
J.F.
,
Davies
S.
,
Vane
C.H.
,
Moss-Hayes
V.
and
Stephenson
M.
2020a
.
From marine bands to hybrid flows: sedimentology of a Mississippian black shale
.
Sedimentology
 ,
67
,
261
304
, https://doi.org/10.1111/sed.12642
44.
Emmings
J.F.
,
Dowey
P.I.
et al
2020b
.
Origin and implications of early diagenetic quartz in the Mississippian Bowland Shale Formation, Craven Basin, UK
.
Marine and Petroleum Geology
 ,
120
, https://doi.org/10.1016/j.marpetgeo.2020.104567
45.
Emmings
J.F.
,
Poulton
S.W.
et al
2020c
.
A Mississippian black shale record of redox oscillation in the Craven Basin, UK
.
Palaeogeography, Palaeoclimatology, Palaeoecology
 ,
538
, https://doi.org/10.1016/j.palaeo.2019.109423
46.
Emmings
J.F.
,
Poulton
S.W.
et al
2022
.
Pyrite mega-analysis reveals modes of anoxia through geological time
.
Science Advances
 ,
8
,
eabj5687
, https://doi.org/10.1126/sciadv.abj5687
47.
Fauchille
A.L.
,
Ma
L.
et al
2017
.
An enhanced understanding of the Basinal Bowland shale in Lancashire (UK), through microtextural and mineralogical observations
.
Marine and Petroleum Geology
 ,
86
,
1374
1390
, https://doi.org/10.1016/j.marpetgeo.2017.07.030
48.
Fauchille
A.L.
,
van den Eijnden
A.P.
et al
2018
.
Variability in spatial distribution of mineral phases in the Lower Bowland Shale, UK, from the mm- to µm-scale: quantitative characterization and modelling
.
Marine and Petroleum Geology
 ,
92
,
109
127
, https://doi.org/10.1016/j.marpetgeo.2018.02.029
49.
Fisher
Q.J.
and
Wignall
P.B.
2001
.
Palaeoenvironmental controls on the uranium distribution in an Upper Carboniferous black shale (Gastrioceras listeri Marine Band) and associated strata; England
.
Chemical Geology
 ,
175
,
605
621
, https://doi.org/10.1016/S0009-2541(00)00376-4
50.
Fraser
A.J.
and
Gawthorpe
R.L.
1990
.
Tectono-stratigraphic development and hydrocarbon habitat of the Carboniferous in northern England
.
Geological Society, London, Special Publications
 ,
55
,
49
86
, https://doi.org/10.1144/GSL.SP.1990.055.01.03
51.
Fraser
A.J.
and
Gawthorpe
R.
2003
.
An Atlas of Carboniferous basin evolution in Northern England
.
Geological Society, London, Memoirs
 ,
28
, https://doi.org/10.1144/GSL.MEM.2003.028
52.
Fraser
A.J.
,
Nash
D.F.
,
Steele
R.P.
,
Ebdon
C.C.
and
Fraser
A.J.
1990
.
A regional assessment of the intra-Carboniferous play of Northern England
.
Geological Society, London, Special Publications
 ,
50
,
417
440
, https://doi.org/10.1144/GSL.SP.1990.050.01.26
53.
Gadd
M.G.
,
Peter
J.M.
et al
2020
.
A Middle Devonian basin-scale precious metal enrichment event across northern Yukon (Canada)
.
Geology
 ,
48
,
242
246
, https://doi.org/10.1130/G46874.1
54.
Gluyas
J.
,
Adams
C.A.
et al
2018
.
Keeping warm: a review of deep geothermal potential of the UK
.
Proceedings of the Institution of Mechanical Engineers, Part A: Journal of Power and Energy
 ,
232
,
115
126
, https://doi.org/10.1177/0957650917749693
55.
Green
P.F.
,
Duddy
I.R.
and
Bray
R.J.
1997
.
Variation in thermal history styles around the Irish Sea and adjacent areas: implications for hydrocarbon occurrence and tectonic evolution
.
Geological Society, London, Special Publications
 ,
124
,
73
93
, https://doi.org/10.1144/GSL.SP.1997.124.01.06
56.
Gross
D.
,
Sachsenhofer
R.F.
et al
2015
.
Organic geochemistry of Mississippian shales (Bowland Shale Formation) in central Britain: implications for depositional environment, source rock and gas shale potential
.
Marine and Petroleum Geology
 ,
59
,
1
21
, https://doi.org/10.1016/j.marpetgeo.2014.07.022
57.
Hamilton
N.
2017
.
ggtern R package version 2.2.1
, https://CRAN.R-project.org/package=ggtern
58.
Hampson
G.J.
1997
.
A sequence stratigraphic model for deposition of the Lower Kinderscout Delta, an Upper Carboniferous turbidite-fronted delta
.
Proceedings of the Yorkshire Geological and Polytechnic Society
 ,
51
,
273
296
, https://doi.org/10.1144/pygs.51.4.273
59.
Hennissen
J.A.I.
,
Hough
E.
et al
2017
.
The prospectivity of a potential shale gas play: an example from the southern Pennine Basin (central England, UK)
.
Marine and Petroleum Geology
 ,
86
,
1047
1066
, https://doi.org/10.1016/j.marpetgeo.2017.06.033
60.
Hennissen
J.A.I.
,
Emmings
J.
et al
2024
.
The Bowland Shale Formation in the Blacon Basin: palaeoecology and resource potential
.
Geological Society, London, Special Publications
 ,
534
, https://doi.org/10.1144/SP534-2022-211
61.
Hild
E.
and
Brumsack
H.J.
1998
.
Major and minor element geochemistry of Lower Aptian sediments from the NW German Basin (core Hohenegglesen KB 40)
.
Cretaceous Research
 ,
19
,
615
633
, https://doi.org/10.1006/cres.1998.0122
62.
Holdsworth
B.
and
Collinson
J.D.
1988
. Millstone Grit cyclicity revisited. In:
Besly
B.M.
and
Kelling
G.
(eds)
Sedimentation in a Synorogenic Basin Complex: The Upper Carboniferous of Northwest Europe
 .
Blackie
,
Glasgow
,
132
152
.
63.
Hollis
C.
and
Walkden
G.
2012
.
Burial diagenetic evolution of the Lower Carboniferous (Dinantian) of the southern margin of the Askrigg Platform and a comparison with the Derbyshire Platform
.
Petroleum Geoscience
 ,
18
,
83
95
, https://doi.org/10.1144/1354-079311-049
64.
IPCC
.
2014
.
Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change
 .
IPCC
,
Geneva, Switzerland
.
65.
Jackson
D.I.
,
Johnson
H.
and
Smith
N.J.P.
1997
.
Stratigraphical relationships and a revised lithostratigraphical nomenclature for the Carboniferous, Permian and Triassic rocks of the offshore East Irish Sea Basin
.
Geological Society, London, Special Publications
 ,
124
,
11
32
, https://doi.org/10.1144/GSL.SP.1997.124.01.02
66.
Jarvie
D.M.
2012a
.
Shale resource systems for oil and gas: part 1 – shale-gas resource systems
.
AAPG Memoirs
 ,
97
, https://doi.org/10.1306/13321446M973489
67.
Jarvie
D.M.
2012b
.
Shale resource systems for oil and gas: part 2 – shale-oil resource systems
.
AAPG Memoirs
 ,
97
, https://doi.org/10.1306/13321447M973489
68.
Juerges
A.
,
Hollis
C.E.
,
Marshall
J.
and
Crowley
S.
2016
.
The control of basin evolution on patterns of sedimentation and diagenesis: an example from the Mississippian Great Orme, North Wales
.
Journal of the Geological Society, London
 ,
173
,
438
456
, https://doi.org/10.1144/jgs2014-149
69.
Kabanov
P.
and
Jiang
C.
2020
.
Photic-zone euxinia and anoxic events in a Middle-Late Devonian shelfal sea of Panthalassan continental margin, NW Canada: changing paradigm of Devonian ocean and sea level fluctuations
.
Global and Planetary Change
 ,
188
,
103153
, https://doi.org/10.1016/j.gloplacha.2020.103153
70.
Kampschulte
A.
and
Strauss
H.
2004
.
The sulfur isotopic evolution of Phanerozoic seawater based on the analysis of structurally substituted sulfate in carbonates
.
Chemical Geology
 ,
204
,
255
286
, https://doi.org/10.1016/j.chemgeo.2003.11.013
71.
Kampschulte
A.
,
Bruckschen
P.
and
Strauss
H.
2001
.
The sulphur isotopic composition of trace sulphates in Carboniferous brachiopods: implications for coeval seawater, correlation with other geochemical cycles and isotope stratigraphy
.
Chemical Geology
 ,
175
,
149
173
, https://doi.org/10.1016/S0009-2541(00)00367-3
72.
Kendrick
M.A.
,
Burgess
R.
,
Pattrick
R.A.D.
and
Turner
G.
2002
.
Hydrothermal fluid origins in a fluorite-rich Mississippi Valley-Type District: combined noble gas (He, Ar, Kr) and halogen (Cl, Br, I) analysis of fluid inclusions from the South Pennine Ore Field, United Kingdom
.
Economic Geology
 ,
97
,
435
451
, https://doi.org/10.2113/gsecongeo.97.3.435
73.
Kerschke
D.
and
Schulz
H.-M.
2013
.
The shale gas potential of Tournaisian, Visean, and Namurian black shales in North Germany: baseline parameters in a geological context
.
Environmental Earth Sciences
 ,
70
,
3817
3837
, https://doi.org/10.1007/s12665-013-2745-9
74.
Kidder
D.L.
and
Worsley
T.R.
2010
.
Phanerozoic Large Igneous Provinces (LIPs), HEATT (Haline Euxinic Acidic Thermal Transgression) episodes, and mass extinctions
.
Palaeogeography, Palaeoclimatology, Palaeoecology
 ,
295
,
162
191
, https://doi.org/10.1016/j.palaeo.2010.05.036
75.
Kirkham
A.
2021
.
Mass collapse and resedimentation on a Brigantian–Early Namurian platform margin, Halkyn–Mold area, North Wales, UK
.
Geological Journal
 ,
56
,
3672
3686
, https://doi.org/10.1002/gj.4127
76.
Könitzer
S.F.
,
Davies
S.
,
Stephenson
M.
and
Leng
M.
2014
.
Depositional controls on mudstone lithofacies in a basinal setting: implications for the delivery of sedimentary organic matter
.
Journal of Sedimentary Research
 ,
84
,
198
214
, https://doi.org/10.2110/jsr.2014.18
77.
Könitzer
S.F.
,
Stephenson
M.H.
,
Davies
S.J.
,
Vane
C.H.
and
Leng
M.J.
2016
.
Significance of sedimentary organic matter input for shale gas generation potential of Mississippian Mudstones, Widmerpool Gulf, UK
.
Review of Palaeobotany and Palynology
 ,
224
,
146
168
, https://doi.org/10.1016/j.revpalbo.2015.10.003
78.
Kuhn
M.
2020
.
Caret R package version 6.0-86
, https://CRAN.R-project.org/package=caret
79.
Lafargue
E.
,
Marquis
F.
and
Pillot
D.
1998
.
Rock-Eval 6 applications in hydrocarbon exploration, production, and soil contamination studies
.
Revue de l'Institut Francais du Petrole
 ,
53
,
421
437
, https://doi.org/10.2516/ogst:1998036
80.
Lecomte
A.
,
Cathelineau
M.
,
Michels
R.
,
Peiffert
C.
and
Brouand
M.
2017
.
Uranium mineralization in the Alum Shale Formation (Sweden): evolution of a U-rich marine black shale from sedimentation to metamorphism
.
Ore Geology Reviews
 ,
88
,
71
98
, https://doi.org/10.1016/j.oregeorev.2017.04.021
81.
Lee
M.K.
,
Brown
G.C.
,
Webb
P.C.
,
Whieldon
J.
and
Rollin
K.E.
1987
.
Heat flow, heat production and thermo-tectonic setting in mainland UK
.
Journal of the Geological Society, London
 ,
144
,
35
42
, https://doi.org/10.1144/gsjgs.144.1.0035
82.
Leeder
M.R.
1982
.
Upper Palaeozoic basins of the British Isles – Caledonide inheritance versus Hercynian plate margin processes
.
Journal of the Geological Society, London
 ,
139
,
479
491
, https://doi.org/10.1144/gsjgs.139.4.0479
83.
Leeder
M.
,
Raiswell
R.
,
Al-Biatty
H.
,
Mcmahon
A.
and
Hardman
M.
1990
.
Carboniferous stratigraphy, sedimentation and correlation of well 48/3-3 in the southern North Sea Basin: integrated use of palynology, natural gamma/sonic logs and carbon/sulphur geochemistry
.
Journal of the Geological Society, London
 ,
147
,
287
300
, https://doi.org/10.1144/gsjgs.147.2.0287
84.
Lewan
M.D.
1986
.
Stable carbon isotopes of amorphous kerogens from Phanerozoic sedimentary rocks
.
Geochimica et Cosmochimica Acta
 ,
50
,
1583
1591
, https://doi.org/10.1016/0016-7037(86)90121-3
85.
Lindgreen
H.
1994
.
Ammonium fixation during illite-smectite diagenesis in Upper Jurassic shale, North Sea
.
Clay Minerals
 ,
29
,
527
537
, https://doi.org/10.1180/claymin.1994.029.4.10
86.
Loucks
R.
and
Ruppel
S.
2007
.
Mississippian Barnett Shale: lithofacies and depositional setting of a deep-water shale-gas succession in the Fort Worth Basin, Texas
.
AAPG Bulletin
 ,
91
,
579
601
, https://doi.org/10.1306/11020606059
87.
Loveless
S.E.
,
Bloomfield
J.P.
et al
2018
.
Characterising the vertical separation of shale-gas source rocks and aquifers across England and Wales (UK)
.
Hydrogeology Journal
 ,
26
,
1975
1987
, https://doi.org/10.1007/s10040-018-1737-y
88.
Magraw
D.
and
Ramsbottom
W.H.C.
1957
.
A deep borehole for oil at Croxteth Park, near Liverpool
.
Geological Journal
 ,
1
,
512
535
, https://doi.org/10.1002/gj.3350010605
89.
Martinsen
O.J.
,
Collinson
J.D.
and
Holdsworth
B.K.
1995
. Millstone Grit cyclicity revisited, II: sequence stratigraphy and sedimentary responses to changes of relative sea-level. In:
Plint
G.A.
(ed.)
Sedimentary Facies Analysis
 .
Blackwell Publishing
,
305
327
.
90.
Maynard
J.R.
and
Leeder
M.R.
1992
.
On the periodicity and magnitude of Late Carboniferous glacio-eustatic sea-level changes
.
Journal of the Geological Society, London
 ,
149
,
303
311
, https://doi.org/10.1144/gsjgs.149.3.0303
91.
Maynard
J.R.
,
Wignall
P.B.
and
Varker
W.J.
1991
.
A ‘hot’ new shale facies from the Upper Carboniferous of Northern England
.
Journal of the Geological Society, London
 ,
148
,
805
808
, https://doi.org/10.1144/gsjgs.148.5.0805
92.
Mazzini
A.
,
Ivanov
M.K.
et al
2004
.
Methane-related authigenic carbonates from the Black Sea: geochemical characterisation and relation to seeping fluids
.
Marine Geology
 ,
212
,
153
181
, https://doi.org/10.1016/j.margeo.2004.08.001
93.
Mehra
A.
,
Keller
B.
et al
2021
.
Curation and analysis of global sedimentary geochemical data to inform Earth history
.
GSA Today
 ,
31
, https://doi.org/10.1130/GSATG484A.1
94.
Mercuzot
M.
,
Thomazo
C.
et al
2021
.
Carbon and nitrogen cycle dynamic in continental late-Carboniferous to early Permian basins of Eastern Pangea (Northeastern Massif Central, France)
.
Frontiers in Earth Science
 ,
9
, https://doi.org/10.3389/feart.2021.705351
95.
Meyer
K.M.
and
Kump
L.R.
2008
.
Oceanic euxinia in earth history: causes and consequences
.
Annual Review of Earth and Planetary Sciences
 ,
36
,
251
288
, https://doi.org/10.1146/annurev.earth.36.031207.124256
96.
Meyers
P.A.
1994
.
Preservation of elemental and isotopic source identification of sedimentary organic matter
.
Chemical Geology
 ,
114
,
289
302
, https://doi.org/10.1016/0009-2541(94)90059-0
97.
Midttømme
K.
and
Roaldset
E.
1999
.
Thermal conductivity of sedimentary rocks: uncertainties in measurement and modelling
.
Geological Society, London, Special Publications
 ,
158
,
45
60
, https://doi.org/10.1144/GSL.SP.1999.158.01.04
98.
Midttømme
K.
,
Roaldset
E.
and
Aagaard
P.
1998
.
Thermal conductivity of selected claystones and mudstones from England
.
Clay Minerals
 ,
33
,
131
145
, https://doi.org/10.1180/000985598545327
99.
Miles
J.C.
,
Appleton
J.D.
et al
2007
.
Indicative Atlas of Radon in England and Wales
 .
Health Protection Agency and the British Geological Survey Report HPA-RPD-033
.
100.
Mitchum
R.M.
and
Van Wagoner
J.C.
1991
.
High-frequency sequences and their stacking patterns: sequence-stratigraphic evidence of high-frequency eustatic cycles
.
Sedimentary Geology
 ,
70
,
131
160
, https://doi.org/10.1016/0037-0738(91)90139-5
101.
Monaghan
A.A.
,
Underhill
J.R.
,
Marshall
J.E.A.
and
Hewett
A.J.
2019
.
Paleozoic plays of NW Europe: an introduction
.
Geological Society, London, Special Publications
 ,
471
,
1
15
, https://doi.org/10.1144/SP471.13
102.
Morton
A.
,
Waters
C.
,
Fanning
M.
,
Chisholm
I.
and
Brettle
M.
2015
.
Origin of Carboniferous sandstones fringing the northern margin of the Wales-Brabant Massif: insights from detrital zircon ages
.
Geological Journal
 ,
50
,
553
574
, https://doi.org/10.1002/gj.2572
103.
Mossmann
J.-R.
,
Aplin
A.C.
,
Curtis
C.D.
and
Coleman
M.L.
1991
.
Geochemistry of inorganic and organic sulphur in organic-rich sediments from the Peru Margin
.
Geochimica et Cosmochimica Acta
 ,
55
,
3581
3595
, https://doi.org/10.1016/0016-7037(91)90057-C
104.
Nash
T.
1979
.
Uranium and Thorium in Granitic Rocks of Northeastern Washington and Northern Idaho, with Comments on Uranium Resource Potential
 .
USGS Open-File Report 79-233
.
105.
Newport
L.P.
,
Aplin
A.C.
,
Gluyas
J.G.
,
Greenwell
H.C.
and
Gröcke
D.R.
2016
.
Geochemical and lithological controls on a potential shale reservoir: Carboniferous Holywell Shale, Wales
.
Marine and Petroleum Geology
 ,
71
,
198
210
, https://doi.org/10.1016/j.marpetgeo.2015.11.026
106.
Newport
S.M.
2019
.
The Bowland Shale (UK): Development of New Sedimentological and Diagenetic Models for a Palaeozoic Fine-Grained Succession
 .
PhD thesis
,
University of Manchester
.
107.
Newport
S.M.
,
Jerrett
R.M.
,
Taylor
K.G.
,
Hough
E.
and
Worden
R.H.
2018
.
Sedimentology and microfacies of a mud-rich slope succession: in the Carboniferous Bowland Basin, NW England (UK)
.
Journal of the Geological Society, London
 ,
175
,
247
262
, https://doi.org/10.1144/jgs2017-036
108.
Nunn
J.A.
and
Lin
G.
2002
.
Insulating effect of coals and organic rich shales: implications for topography-driven fluid flow, heat transport, and genesis of ore deposits in the Arkoma Basin and Ozark Plateau
.
Basin Research
 ,
14
,
129
145
, https://doi.org/10.1046/j.1365-2117.2002.00172.x
109.
Nunn
J.A.
,
Lin
G.
and
Zhang
L.
1999
. Thermal insulation by low thermal conductivity shales: implications for basin-scale fluid flow and heat transport. In:
Förster
A.
and
Merriam
D.F.
(eds)
Geothermics in Basin Analysis
 .
Springer US
,
Boston, MA
,
117
130
.
110.
Palci
F.
,
Fraser
A.J.
et al
2020
.
Shale oil and gas resource evaluation through 3D basin and petroleum systems modelling: a case study from the East Midlands, onshore UK
.
Petroleum Geoscience
 ,
26
,
525
543
, https://doi.org/10.1144/petgeo2019-069
111.
Parkes
D.
,
Busby
J.
,
Kemp
S.J.
,
Petitclerc
E.
and
Mounteney
I.
2021
.
The thermal properties of the Mercia Mudstone Group
.
Quarterly Journal of Engineering Geology and Hydrogeology
 ,
54
, https://doi.org/10.1144/qjegh2020-098
112.
Parnell
J.
and
Swainbank
I.
1990
.
Pb-Pb dating of hydrocarbon migration into a bitumen-bearing ore deposit, North Wales
.
Geology
 ,
18
,
1028
1030
, https://doi.org/10.1130/0091-7613(1990)018<1028:PPDOHM>2.3.CO;2
113.
Parnell
J.
,
Brolly
C.
,
Spinks
S.
and
Bowden
S.
2016
.
Selenium enrichment in Carboniferous Shales, Britain and Ireland: problem or opportunity for shale gas extraction?
Applied Geochemistry
 ,
66
,
82
87
, https://doi.org/10.1016/j.apgeochem.2015.12.008
114.
Parnell
J.
,
Nkoko
K.
,
Feldmann
J.
and
Armstrong
J.
2022
.
Speciation of trace elements in the Bowland Shale
.
Geological Society, London, Special Publications
 ,
534
, https://doi.org/10.1144/SP534-2020-195
115.
Patruno
S.
and
Helland-Hansen
W.
2018
.
Clinoforms and clinoform systems: review and dynamic classification scheme for shorelines, subaqueous deltas, shelf edges and continental margins
.
Earth-Science Reviews
 ,
185
,
202
233
, https://doi.org/10.1016/j.earscirev.2018.05.016
116.
Pearce
T.J.
,
McLean
D.
,
Martin
J.H.
,
Ratcliffe
K.
and
Wray
D.
2010
.
A whole-rock geochemical approach to the recognition and correlation of ‘marine bands
.
SEPM Special Publications
 ,
94
,
221
238
.
117.
Pellegrini
C.
,
Maselli
V.
et al
2017
.
How to make a 350-m-thick lowstand systems tract in 17 000 years: the Late Pleistocene Po River (Italy) lowstand wedge
.
Geology
 ,
45
,
327
330
, https://doi.org/10.1130/G38848.1
118.
Peters-Kottig
W.
,
Strauss
H.
and
Kerp
H.
2006
.
The land plant δ13C record and plant evolution in the Late Palaeozoic
.
Palaeogeography, Palaeoclimatology, Palaeoecology
 ,
240
,
237
252
, https://doi.org/10.1016/j.palaeo.2006.03.051
119.
Peters
S.
and
Loss
D.
2012
.
Storm and fair-weather wave base: a relevant distinction?
Geology
 ,
40
,
511
514
, https://doi.org/10.1130/G32791.1
120.
Phan
T.T.
,
Capo
R.C.
et al
2015
.
Trace metal distribution and mobility in drill cuttings and produced waters from Marcellus Shale gas extraction: uranium, arsenic, barium
.
Applied Geochemistry
 ,
60
,
89
103
, https://doi.org/10.1016/j.apgeochem.2015.01.013
121.
Pharaoh
T.C.
,
Gent
C.M.A.
et al
2019
.
An overlooked play? Structure, stratigraphy and hydrocarbon prospectivity of the Carboniferous in the East Irish Sea–North Channel basin complex
.
Geological Society, London, Special Publications
 ,
471
,
281
316
, https://doi.org/10.1144/SP471.7
122.
Pharaoh
T.
,
Haslam
R.
et al
2020
.
The Môn-Deemster-Ribblesdale Fold-Thrust Belt, Central UK: a concealed Variscan inversion belt located on weak Caledonian crust
.
Geological Society, London, Special Publications
 ,
490
,
153
176
, https://doi.org/10.1144/SP490-2018-109
123.
R, Core Team
2018
.
R: a language and environment for statistical computing
 .
R Foundation for Statistical Computing
,
Vienna, Austria
, https://www.R-project.org
124.
Raiswell
R.
and
Berner
R.A.
1987
.
Organic carbon losses during burial and thermal maturation of normal marine shales
.
Geology
 ,
15
,
853
856
, https://doi.org/10.1130/0091-7613(1987)15<853:OCLDBA>2.0.CO;2
125.
Ramsbottom
W.H.C.
1977
.
Major cycles of transgression and regression (mesothems) in the Namurian
.
Proceedings of the Yorkshire Geological and Polytechnic Society
 ,
41
,
261
291
, https://doi.org/10.1144/pygs.41.3.261
126.
Ramsbottom
W.H.C.
1979
.
Rates of transgression and regression in the Carboniferous of NW Europe
.
Journal of the Geological Society, London
 ,
136
,
147
153
, https://doi.org/10.1144/gsjgs.136.2.0147
127.
Ramsbottom
W.H.C.
and
Saunders
W.B.
1985
.
Evolution and evolutionary biostratigraphy of Carboniferous ammonoids
.
Journal of Paleontology
 ,
59
,
123
139
.
128.
Ramsbottom
W.H.C.
,
Rhys
G.J.
and
Smith
E.G.
1962
.
Boreholes in the Carboniferous rocks of the Ashover district, Derbyshire
.
Bulletin of the Geological Survey of Great Britain
 ,
19
,
75
168
.
129.
Riley
D.A.
,
Pearce
T.J.
,
Mathia
E.
,
Ratcliffe
K.
and
Martin
J.
2018
.
The application of elemental geochemistry to UK onshore unconventional plays
.
Geological Society, London, Petroleum Geology Conference Series
 ,
8
,
585
594
, https://doi.org/10.1144/PGC8.8
130.
Rollin
K.
2002
.
Assessment of BGS Data for Ground Source Heat Pump Installations in the UK
 .
British Geological Survey Internal Report IR/02/196
.
131.
Rutqvist
J.
2020
.
Thermal management associated with geologic disposal of large spent nuclear fuel canisters in tunnels with thermally engineered backfill
.
Tunnelling and Underground Space Technology
 ,
102
,
103454
, https://doi.org/10.1016/j.tust.2020.103454
132.
Rybach
L.
1976
. Radioactive heat production: a physical property determined by the chemistry of rocks. In:
Strens
R.G.J.
(ed.)
The Physics and Chemistry of Minerals and Rocks
 .
Wiley
,
London
,
309
318
.
133.
Rygel
M.C.
,
Fielding
C.R.
,
Frank
T.D.
and
Birgenheier
L.P.
2008
.
The magnitude of late Paleozoic glacioeustatic fluctuations: a synthesis
.
Journal of Sedimentary Research
 ,
78
,
500
511
, https://doi.org/10.2110/jsr.2008.058
134.
Saunders
W.B.
and
Ramsbottom
W.H.C.
1986
.
The mid-Carboniferous eustatic event
.
Geology
 ,
14
,
208
212
, https://doi.org/10.1130/0091-7613(1986)14<208:TMEE>2.0.CO;2
135.
Schuster
A.
,
Strehlow
B.W.
,
Eckford-Soper
L.
,
McAllen
R.
and
Canfield
D.E.
2021
.
Effects of seasonal anoxia on the microbial community structure in demosponges in a Marine Lake in Lough Hyne, Ireland
.
mSphere
 ,
6
,
e00991
e009920
, https://doi.org/10.1128/mSphere.00991-20
136.
Shanmugam
G.
and
Moiola
R.J.
1982
.
Eustatic control of turbidites and winnowed turbidites
.
Geology
 ,
10
,
231
235
, https://doi.org/10.1130/0091-7613(1982)10<231:ECOTAW>2.0.CO;2
137.
Sims
M.J.E.
,
Sephton
M.A.
,
Watson
J.S.
,
Fraser
A.J.
and
Vane
C.H.
2024
.
Biomarker evidence for the depositional environment of basinal UK Mississippian mudstones
.
Geological Society, London, Special Publications
 ,
534
, https://doi.org/10.1144/SP534-2022-225
138.
Smedley
P.L.
,
Smith
B.
,
Abesser
C.
and
Laptworth
D.
2006
.
Uranium Occurrence and Behaviour in British Groundwater
 .
British Geological Survey Commissioned Report CR/06/050N
.
139.
Smith
N.
,
Kirby
G.A.
,
Pharaoh
T.C.
,
Hulbert
A.G.
and
Shaw
K.
2005
.
Structure and Evolution of the South-West Pennine Basin and Adjacent Area
 .
British Geological Survey
.
140.
Somerville
I.D.
1979
.
A cyclicity in the early Brigantian (D2) limestones east of the Clwydian Range, North Wales and its use in correlation
.
Geological Journal
 ,
14
,
69
86
, https://doi.org/10.1002/gj.3350140105
141.
Spears
D.A.
,
Kanaris-Sotiriou
R.
,
Riley
N.
and
Krause
P.
1999
.
Namurian bentonites in the Pennine Basin, UK – origin and magmatic affinities
.
Sedimentology
 ,
46
,
385
401
, https://doi.org/10.1046/j.1365-3091.1999.00220.x
142.
Talling
P.
,
Masson
D.
,
Sumner
E.
and
Malgesini
G.
2012
.
Subaqueous sediment density flows: depositional processes and deposit types
.
Sedimentology
 ,
59
,
1937
2003
, https://doi.org/10.1111/j.1365-3091.2012.01353.x
143.
Taylor
S.
and
McLennan
S.
1985
.
The Continental Crust: Its Composition and Evolution
 .
Blackwell Scientific
.
144.
Tissot
B.
and
Welte
D.
1984
.
Petroleum Formation and Occurrence
 .
Springer
.
145.
Tribovillard
N.
2020
.
Arsenic in marine sediments: how robust a redox proxy?
Palaeogeography, Palaeoclimatology, Palaeoecology
 ,
550
,
109745
, https://doi.org/10.1016/j.palaeo.2020.109745
146.
Tribovillard
N.
,
Algeo
T.
,
Lyons
T.
and
Ruboulleau
A.
2006
.
Trace metals as paleoredox and paleoproductivity proxies: an update
.
Chemical Geology
 ,
232
,
12
32
, https://doi.org/10.1016/j.chemgeo.2006.02.012
147.
Tribovillard
N.
,
Algeo
T.J.
,
Baudin
F.
and
Riboulleau
A.
2012
.
Analysis of marine environmental conditions based on molybdenum–uranium covariation – applications to Mesozoic paleoceanography
.
Chemical Geology
 ,
324–325
,
46
58
, https://doi.org/10.1016/j.chemgeo.2011.09.009
148.
Tyson
R.
(ed.)
1995
.
Sedimentary Organic Matter: Organic Facies and Palynofacies
 .
Chapman and Hall
,
London
.
149.
van de Velde
S.J.
,
Reinhard
C.T.
,
Ridgwell
A.
and
Meysman
F.J.R.
2020
.
Bistability in the redox chemistry of sediments and oceans
.
Proceedings of the National Academy of Sciences
 ,
117
,
33043
33050
, https://doi.org/10.1073/pnas.2008235117
150.
van den Boogaart
G.
,
Tolosana-Delgado
R.
and
Bren
M.
2020
.
Compositions R package version 2.0-0
, https://CRAN.R-project.org/package=compositions
151.
Van Mooy
B.A.S.
,
Keil
R.G.
and
Devol
A.H.
2002
.
Impact of suboxia on sinking particulate organic carbon: enhanced carbon flux and preferential degradation of amino acids via denitrification
.
Geochimica et Cosmochimica Acta
 ,
66
,
457
465
, https://doi.org/10.1016/S0016-7037(01)00787-6
152.
Veevers
J.J.
and
Powell
C.M.
1987
.
Late Paleozoic glacial episodes in Gondwanaland reflected in transgressive-regressive depositional sequences in Euramerica
.
GSA Bulletin
 ,
98
,
475
487
, https://doi.org/10.1130/0016-7606(1987)98<475:LPGEIG>2.0.CO;2
153.
Vilà
M.
,
Fernández
M.
and
Jiménez-Munt
I.
2010
.
Radiogenic heat production variability of some common lithological groups and its significance to lithospheric thermal modeling
.
Tectonophysics
 ,
490
,
152
164
, https://doi.org/10.1016/j.tecto.2010.05.003
154.
Warr
L.N.
2000
. The Variscan Orogeny: the welding of Pangaea. In:
Woodcock
N.H.
and
Strachan
R.
(eds)
Geological History of Britain and Ireland
 .
Wiley-Blackwell
,
Hoboken, USA
,
271
294
.
155.
Warren
P.T.
,
Price
D.
et al
1984
.
Geology of the Country Around Rhyl and Denbigh, Memoir for 1:50 000 Geological Sheets 95 and 107 and Parts of Sheets 94 and 106
 .
Memoirs of the Geological Survey of Great Britain, England and Wales (Sheet – New Series)
.
HMSO
,
London
.
156.
Waters
C.N.
and
Condon
D.J.
2012
.
Nature and timing of Late Mississippian to Mid-Pennsylvanian glacio-eustatic sea-level changes of the Pennine Basin, UK
.
Journal of the Geological Society, London
 ,
169
,
37
51
, https://doi.org/10.1144/0016-76492011-047
157.
Waters
C.N.
and
Davies
S.J.
2006
. Carboniferous: extensional basins, advancing deltas and coal swamps. In:
Brenchley
P.J.
(ed.)
The Geology of England and Wales
 .
Geological Society, London
,
173
223
.
158.
Waters
C.N.
,
Browne
M.A.E.
,
Dean
M.T.
and
Powell
J.H.
2007
.
Lithostratigraphical Framework for Carboniferous Successions of Great Britain (Onshore)
 .
British Geological Survey Research Report RR/07/01
.
159.
Waters
C.N.
,
Waters
R.A.
,
Barclay
W.J.
and
Davies
J.R.
2009
.
A Lithostratigraphical Framework for the Carboniferous Successions of Southern Great Britain (Onshore)
 .
British Geological Survey Research Report RR/09/01
.
160.
Waters
C.N.
,
Somerville
I.D.
et al
2011
.
A Revised Correlation of Carboniferous Rocks in the British Isles
 .
Geological Society, London
.
161.
Waters
C.N.
,
Vane
C.H.
et al
2020
.
Lithological and chemostratigraphic discrimination of facies within the Bowland Shale Formation within the Craven and Edale basins, UK
.
Petroleum Geoscience
 ,
26
,
325
345
, https://doi.org/10.1144/petgeo2018-039
162.
Waters
C.N.
,
Morton
A.
and
Frei
D.
2021
.
Interplay of southern, western, and northern sources during deposition of North Wales Carboniferous sandstones, determined from heavy minerals, mineral chemistry, and detrital zircon ages
.
Geological Journal
 ,
56
,
2699
2719
, https://doi.org/10.1002/gj.4063
163.
Wei
W.
and
Algeo
T.J.
2020
.
Elemental proxies for paleosalinity analysis of ancient shales and mudrocks
.
Geochimica et Cosmochimica Acta
 ,
287
,
341
366
, https://doi.org/10.1016/j.gca.2019.06.034
164.
Whitelaw
P.
,
Uguna
C.N.
et al
2019
.
Shale gas reserve evaluation by laboratory pyrolysis and gas holding capacity consistent with field data
.
Nature Communications
 ,
10
,
3659
, https://doi.org/10.1038/s41467-019-11653-4
165.
Wickham
H.
2009
.
ggplot2: Elegant Graphics for Data Analysis
 .
Springer-Verlag
,
New York
.
166.
Wickham
H.
and
Francois
R.
2016
.
dplyr R package version 0.5.0
, https://CRAN.R-project.org/package=dplyr
167.
Wignall
P.B.
1994
.
Black Shales
 .
Clarendon Press; Oxford University Press
,
Oxford
.
168.
Williams
G.D.
and
Eaton
G.P.
1993
.
Stratigraphic and structural analysis of the Late Palaeozoic–Mesozoic of NE Wales and Liverpool Bay: implications for hydrocarbon prospectivity
.
Journal of the Geological Society, London
 ,
150
,
489
499
, https://doi.org/10.1144/gsjgs.150.3.0489
169.
Worden
R.H.
,
Allen
M.J.
et al
2020
.
Lower Cretaceous Rodby and Palaeocene Lista Shales: characterisation and comparison of top-seal mudstones at two planned CCS sites, Offshore UK
.
Minerals
 ,
10
,
691
, https://doi.org/10.3390/min10080691
170.
Zeileis
A.
and
Grothendieck
G.
2005
.
Zoo: S3 infrastructure for regular and irregular time series
.
Journal of Statistical Software
 ,
14
,
1
27
, https://doi.org/10.18637/jss.v014.i06
171.
Zijp
M.
,
Nelskamp
S.
and
Doornenbal
H.
2017
.
Resource Estimation of Shale Gas and Shale Oil in Europe
 .
Report T7b of the EUOGA study (EU Unconventional Oil and Gas Assessment) commissioned by European Commission Joint Research Centre to GEUS
.

Figures & Tables

Fig. 1.

Brigantian block-and-basin structures and key borehole and outcrop sections across central Britain, with emphasis on the vicinity of the Blacon Basin. BL, Bala Lineament; BLW-E, Blacon East-1 well; CLH, Central Lancashire High; CRO, Croxteth-1 well; EFZ, Edgerley–Waverton Fault Zone; ERB, Erbistock-1 well; GEF, Great Ewloe Fault; HH, Holme High; LRR, Llŷn–Rossendale Ridge; MGH, Milton Green High; MIL, Milton Green-1 well; NFZ, Nercwys–Nant-figillt Fault Zone; PF, Pendle fault; RB, Rossendale Basin. Source: after Andrews (2013), Fraser and Gawthorpe (1990, 2003), Fraser et al. (1990), Pharaoh et al. (2020), Smith et al. (2005), Waters and Davies (2006) and Waters et al. (2007, 2009, 2011, 2020, 2021).

Fig. 1.

Brigantian block-and-basin structures and key borehole and outcrop sections across central Britain, with emphasis on the vicinity of the Blacon Basin. BL, Bala Lineament; BLW-E, Blacon East-1 well; CLH, Central Lancashire High; CRO, Croxteth-1 well; EFZ, Edgerley–Waverton Fault Zone; ERB, Erbistock-1 well; GEF, Great Ewloe Fault; HH, Holme High; LRR, Llŷn–Rossendale Ridge; MGH, Milton Green High; MIL, Milton Green-1 well; NFZ, Nercwys–Nant-figillt Fault Zone; PF, Pendle fault; RB, Rossendale Basin. Source: after Andrews (2013), Fraser and Gawthorpe (1990, 2003), Fraser et al. (1990), Pharaoh et al. (2020), Smith et al. (2005), Waters and Davies (2006) and Waters et al. (2007, 2009, 2011, 2020, 2021).

Fig. 2.

General stratigraphy, regional Arnsbergian lithofacies and selected well log data. (a) Generalized stratigraphy and (b) regional lithofacies from Waters et al. (2007), including absolute ages (from Davydov et al. 2012; Waters and Condon 2012) and the updated Early–Late Serpukhovian boundary (Cózar and Somerville 2021). (c) The Bowland Shale Formation exhibits elevated total gamma-ray (GR, in American Petroleum Institute units (API)) and U (ppm) concentrations. (d) Correlation between wireline gamma v. synthetic gamma-ray (estimated from measured U, Th and K contents; see ‘Methods’) verifies the alignment of reported corebox depths to the wireline data (see Fig. 3). Ellesmere Port-1 is highlighted (53° 17′ 43.3841″ N, 02° 54′ 21.13091″ W; 339629 mE 378016 mN, British National Grid). Y, Yeadonian; M, Marsdenian; K, Kinderscoutian; Al, Alportian; C, Chokierian; Ar, Arnsbergian; P, Pendleian; Brig., Brigantian.

Fig. 2.

General stratigraphy, regional Arnsbergian lithofacies and selected well log data. (a) Generalized stratigraphy and (b) regional lithofacies from Waters et al. (2007), including absolute ages (from Davydov et al. 2012; Waters and Condon 2012) and the updated Early–Late Serpukhovian boundary (Cózar and Somerville 2021). (c) The Bowland Shale Formation exhibits elevated total gamma-ray (GR, in American Petroleum Institute units (API)) and U (ppm) concentrations. (d) Correlation between wireline gamma v. synthetic gamma-ray (estimated from measured U, Th and K contents; see ‘Methods’) verifies the alignment of reported corebox depths to the wireline data (see Fig. 3). Ellesmere Port-1 is highlighted (53° 17′ 43.3841″ N, 02° 54′ 21.13091″ W; 339629 mE 378016 mN, British National Grid). Y, Yeadonian; M, Marsdenian; K, Kinderscoutian; Al, Alportian; C, Chokierian; Ar, Arnsbergian; P, Pendleian; Brig., Brigantian.

Fig. 3.

Key static borehole image log and core observations used for the alignment of reported well and core depth. Depth alignment was achieved by matching the same linear features present in both datasets. MTD, mass transport deposit.

Fig. 3.

Key static borehole image log and core observations used for the alignment of reported well and core depth. Depth alignment was achieved by matching the same linear features present in both datasets. MTD, mass transport deposit.

Fig. 4.

Sedimentology and selected geochemistry through the late Arnsbergian–Chokierian Bowland Shale in Ellesmere Port-1 (core 2). See Figure 3 for key to sedimentary structures. Palynofacies data are from Hennissen et al. (2024), using the fields of Tyson (1995). I, highly proximal shelf/basin; II, marginal dysoxic–anoxic basin; III, heterolithic (proximal) shelf; IVa, dysoxic–suboxic shelf edge; IVb, dysoxic–anoxic shelf edge; V, mud-dominated (distal) oxic shelf; VI, proximal suboxic–anoxic shelf; VII, distal dysoxic–suboxic shelf; VIII, distal dysoxic–anoxic shelf; IX, distal suboxic–anoxic basin. MTD, mass transport deposit.

Fig. 4.

Sedimentology and selected geochemistry through the late Arnsbergian–Chokierian Bowland Shale in Ellesmere Port-1 (core 2). See Figure 3 for key to sedimentary structures. Palynofacies data are from Hennissen et al. (2024), using the fields of Tyson (1995). I, highly proximal shelf/basin; II, marginal dysoxic–anoxic basin; III, heterolithic (proximal) shelf; IVa, dysoxic–suboxic shelf edge; IVb, dysoxic–anoxic shelf edge; V, mud-dominated (distal) oxic shelf; VI, proximal suboxic–anoxic shelf; VII, distal dysoxic–suboxic shelf; VIII, distal dysoxic–anoxic shelf; IX, distal suboxic–anoxic basin. MTD, mass transport deposit.

Fig. 5.

Sedimentology and selected geochemistry through the late Kinderscoutian Bowland Shale in Ellesmere Port-1 (core 1). See Figure 3 for key to sedimentary structures. Palynofacies data are from Hennissen et al. (2024), using the fields of Tyson (1995). I, highly proximal shelf/basin; II, marginal dysoxic–anoxic basin; III, heterolithic (proximal) shelf; Iva, dysoxic–suboxic shelf edge; IVb, dysoxic–anoxic shelf edge; V, mud-dominated (distal) oxic shelf; VI, proximal suboxic–anoxic shelf; VII, distal dysoxic–suboxic shelf; VIII, distal dysoxic–anoxic shelf; IX, distal suboxic–anoxic basin. MTD, mass transport deposit.

Fig. 5.

Sedimentology and selected geochemistry through the late Kinderscoutian Bowland Shale in Ellesmere Port-1 (core 1). See Figure 3 for key to sedimentary structures. Palynofacies data are from Hennissen et al. (2024), using the fields of Tyson (1995). I, highly proximal shelf/basin; II, marginal dysoxic–anoxic basin; III, heterolithic (proximal) shelf; Iva, dysoxic–suboxic shelf edge; IVb, dysoxic–anoxic shelf edge; V, mud-dominated (distal) oxic shelf; VI, proximal suboxic–anoxic shelf; VII, distal dysoxic–suboxic shelf; VIII, distal dysoxic–anoxic shelf; IX, distal suboxic–anoxic basin. MTD, mass transport deposit.

Fig. 6.

Clay, carbonate and quartz bulk compositional ternary plot estimated using scaled core scan XRF count data (2Al, 0.05Ca and 10Zr, respectively). See text for discussion.

Fig. 6.

Clay, carbonate and quartz bulk compositional ternary plot estimated using scaled core scan XRF count data (2Al, 0.05Ca and 10Zr, respectively). See text for discussion.

Fig. 7.

Excess silica biplots, after Emmings et al. (2020b). Regression lines for the Ellesmere Port-1 data (red line, with 95% confidence interval) and Emmings et al. (2020b) are also plotted. The excess Si field delineates siliceous mudstones. The field is calibrated for Bowland Shale mudstones and compositionally immature siltstones and is not necessarily valid for other sedimentary rocks. See Figure 6 for key to facies.

Fig. 7.

Excess silica biplots, after Emmings et al. (2020b). Regression lines for the Ellesmere Port-1 data (red line, with 95% confidence interval) and Emmings et al. (2020b) are also plotted. The excess Si field delineates siliceous mudstones. The field is calibrated for Bowland Shale mudstones and compositionally immature siltstones and is not necessarily valid for other sedimentary rocks. See Figure 6 for key to facies.

Fig. 8.

Major- and trace-element enrichment factors for chemofacies A–G, normalized to the chemofacies F–G (oxic–suboxic) mean. Red lines for the enrichment factor plots are mean curves. Bottom right: δ34Sbulk, δ13Corg and C/N histograms and density curves for chemofacies A–G.

Fig. 8.

Major- and trace-element enrichment factors for chemofacies A–G, normalized to the chemofacies F–G (oxic–suboxic) mean. Red lines for the enrichment factor plots are mean curves. Bottom right: δ34Sbulk, δ13Corg and C/N histograms and density curves for chemofacies A–G.

Fig. 9.

Geochemical proxies for bottom water salinity and redox. (a) S/TOC (Berner and Raiswell 1984) and Sr/Ba (Wei and Algeo 2020) palaeosalinity proxies. Fields are bounded by the thresholds for salinity as defined in Wei and Algeo (2020), favouring the more widely established S/TOC proxy. Conflict zones could indicate decoupling between the S/TOC and/or Sr/Ba proxies, likely due to diagenetic effects. Circled data points exhibit inorganic C >0.3 wt%, a local threshold for Sr–bound and/or Ba-associated carbonate. In these cases, the Sr/Ba palaeosalinity proxy is invalid (Wei and Algeo 2020). (b) Mo v. U enrichment factors (EFs) (Algeo and Tribovillard 2009; Tribovillard et al. 2012) suggest fluctuation between oxygenated and weakly to moderately anoxic, intermittently sulfidic particulate shuttle conditions. (c) Mo v. TOC suggests deposition under weakly to moderately restricted bottom water conditions (Algeo and Lyons 2006). See ‘Methods’ for calculation of EFs and derivation of original TOC (TOCo). (d–g) Interpreted redox and salinity response to fourth-order sea-level fluctuation, assuming productivity-driven anoxia in a weakly to moderately restricted basin (sensuEmmings et al. 2020c). The fourth-order cyclicity is partially overprinted by progradation of the Cefn-y-Fedw delta system. AMO, anaerobic oxidation of methane; TOC, total organic carbon.

Fig. 9.

Geochemical proxies for bottom water salinity and redox. (a) S/TOC (Berner and Raiswell 1984) and Sr/Ba (Wei and Algeo 2020) palaeosalinity proxies. Fields are bounded by the thresholds for salinity as defined in Wei and Algeo (2020), favouring the more widely established S/TOC proxy. Conflict zones could indicate decoupling between the S/TOC and/or Sr/Ba proxies, likely due to diagenetic effects. Circled data points exhibit inorganic C >0.3 wt%, a local threshold for Sr–bound and/or Ba-associated carbonate. In these cases, the Sr/Ba palaeosalinity proxy is invalid (Wei and Algeo 2020). (b) Mo v. U enrichment factors (EFs) (Algeo and Tribovillard 2009; Tribovillard et al. 2012) suggest fluctuation between oxygenated and weakly to moderately anoxic, intermittently sulfidic particulate shuttle conditions. (c) Mo v. TOC suggests deposition under weakly to moderately restricted bottom water conditions (Algeo and Lyons 2006). See ‘Methods’ for calculation of EFs and derivation of original TOC (TOCo). (d–g) Interpreted redox and salinity response to fourth-order sea-level fluctuation, assuming productivity-driven anoxia in a weakly to moderately restricted basin (sensuEmmings et al. 2020c). The fourth-order cyclicity is partially overprinted by progradation of the Cefn-y-Fedw delta system. AMO, anaerobic oxidation of methane; TOC, total organic carbon.

Fig. 10.

The results of three KNN predictive chemofacies models. (a) Chemofacies were derived from principal component analysis and hierarchical cluster analysis of selected XRF core scan elements, simplified (aggregated) and matched to the nearest wireline datum. The models were trained and deployed by matching the XRF-derived chemofacies to selected wireline logs. kNN (1–3) represent different approaches (see ‘Methods’). (b) The trained models were deployed on the observed Bowland Shale section at Ellesmere Port-1 (1443 to 1767 m below surface) (kNN method 2 is displayed). Prominent sand packages probably correspond to lower, middle and upper extensions of the Cefn-y-Fedw Sandstone (CSF) (Waters et al. 2021) coincident with three periods of regional channel incision (Waters and Condon 2012). Biostratigraphic ammonoid zones observed in core (bold), other inferred major flooding events and biozones (italicized, from Davies et al. 2011) and interpreted substages and extrapolated ages (Waters and Condon 2012) are also indicated.

Fig. 10.

The results of three KNN predictive chemofacies models. (a) Chemofacies were derived from principal component analysis and hierarchical cluster analysis of selected XRF core scan elements, simplified (aggregated) and matched to the nearest wireline datum. The models were trained and deployed by matching the XRF-derived chemofacies to selected wireline logs. kNN (1–3) represent different approaches (see ‘Methods’). (b) The trained models were deployed on the observed Bowland Shale section at Ellesmere Port-1 (1443 to 1767 m below surface) (kNN method 2 is displayed). Prominent sand packages probably correspond to lower, middle and upper extensions of the Cefn-y-Fedw Sandstone (CSF) (Waters et al. 2021) coincident with three periods of regional channel incision (Waters and Condon 2012). Biostratigraphic ammonoid zones observed in core (bold), other inferred major flooding events and biozones (italicized, from Davies et al. 2011) and interpreted substages and extrapolated ages (Waters and Condon 2012) are also indicated.

Fig. 11.

Summary of interpreted Namurian sedimentary architecture in the Blacon Basin, with major flooding events of Davies et al. (2011) and general stratigraphy after Magraw and Ramsbottom (1957), Davies et al. (2004), Smith et al. (2005), Waters et al. (2009, 2021) and Andrews (2013). The shelf-edge delta wedge of the Late Pleistocene Po River (Adriatic Sea, offshore central Italy) may represent a partial (albeit short-lived) modern analogue (Pellegrini et al. 2017; Patruno and Helland-Hansen 2018). Gamma-ray well logs for Croxteth-1, Ellesmere Port-1 (this study), Blacon East-1 and Milton Green-1 are shown for reference. The line of section is flattened and truncated on a Yeadonian marine flooding surface (e.g., G1 marine band) within the Gwespyr Sandstone. See Figure 1 for line of section. MB, marine band.

Fig. 11.

Summary of interpreted Namurian sedimentary architecture in the Blacon Basin, with major flooding events of Davies et al. (2011) and general stratigraphy after Magraw and Ramsbottom (1957), Davies et al. (2004), Smith et al. (2005), Waters et al. (2009, 2021) and Andrews (2013). The shelf-edge delta wedge of the Late Pleistocene Po River (Adriatic Sea, offshore central Italy) may represent a partial (albeit short-lived) modern analogue (Pellegrini et al. 2017; Patruno and Helland-Hansen 2018). Gamma-ray well logs for Croxteth-1, Ellesmere Port-1 (this study), Blacon East-1 and Milton Green-1 are shown for reference. The line of section is flattened and truncated on a Yeadonian marine flooding surface (e.g., G1 marine band) within the Gwespyr Sandstone. See Figure 1 for line of section. MB, marine band.

Fig. 12.

Comparison of RHPs and U, Th and K concentrations for a range of rock types. (a) Calculated RHPs through the Bowland Shale based on the spectral gamma-ray well log data superimposed on the machine predictions. (b, c) RHPs and U, Th and K concentrations for limestone, claystone, calcareous mudstone and siltstone–sandstone facies in the Bowland Shale in Ellesmere Port-1, based on spectral gamma-ray and discrete XRF analysis (the latter indicated by asterisks). The UK RHP sedimentary mean (Rollin 2002; Busby and Terrington 2017) and RHPs for the Bowland Shale in the Craven Basin (Newport 2019; Emmings et al. 2020c), the Kimmeridge Clay (Elm Tree Farm, Yorkshire, J.F.E. unpublished borehole data/analyses), Holywell Shale (Newport et al. 2016) and italicized baseline data (Vilà et al. 2010) are also shown.

Fig. 12.

Comparison of RHPs and U, Th and K concentrations for a range of rock types. (a) Calculated RHPs through the Bowland Shale based on the spectral gamma-ray well log data superimposed on the machine predictions. (b, c) RHPs and U, Th and K concentrations for limestone, claystone, calcareous mudstone and siltstone–sandstone facies in the Bowland Shale in Ellesmere Port-1, based on spectral gamma-ray and discrete XRF analysis (the latter indicated by asterisks). The UK RHP sedimentary mean (Rollin 2002; Busby and Terrington 2017) and RHPs for the Bowland Shale in the Craven Basin (Newport 2019; Emmings et al. 2020c), the Kimmeridge Clay (Elm Tree Farm, Yorkshire, J.F.E. unpublished borehole data/analyses), Holywell Shale (Newport et al. 2016) and italicized baseline data (Vilà et al. 2010) are also shown.

Fig. 13.

Measured v. modelled temperature gradients through the Bowland Shale Formation in Ellesmere Port-1. It should be noted the borehole temperatures are unlikely to represent true, equilibrated values and therefore probably represent minimum estimates.

Fig. 13.

Measured v. modelled temperature gradients through the Bowland Shale Formation in Ellesmere Port-1. It should be noted the borehole temperatures are unlikely to represent true, equilibrated values and therefore probably represent minimum estimates.

or Create an Account

Close Modal
Close Modal