Hostname: page-component-78c5997874-s2hrs Total loading time: 0 Render date: 2024-10-28T02:23:16.551Z Has data issue: false hasContentIssue false

A global reference for black shale geochemistry and the T-OAE revisited: upper Pliensbachian – middle Toarcian (Lower Jurassic) chemostratigraphy in the Cleveland Basin, England

Published online by Cambridge University Press:  25 October 2024

Ian Jarvis*
Affiliation:
Department of Geography, Geology and the Environment, Kingston University London, Kingston upon Thames, UK
Elizabeth Atar
Affiliation:
Department of Earth Sciences, Durham University, Durham, UK
Darren R. Gröcke
Affiliation:
Department of Earth Sciences, Durham University, Durham, UK
Liam G. Herringshaw
Affiliation:
Centre for Lifelong Learning, University of York, York, UK
João P. Trabucho-Alexandre
Affiliation:
Department of Earth Sciences, Durham University, Durham, UK Department of Earth Sciences, Universiteit Utrecht, Utrecht, The Netherlands
*
Corresponding author: Ian Jarvis; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

The Pliensbachian–Toarcian succession of North Yorkshire provides a global reference for the interval incorporating the Toarcian Oceanic Anoxic Event (T-OAE, ∼183 Ma). Major and trace element, carbon stable-isotope (δ13Corg) and total organic carbon (TOC) data for the Dove’s Nest core, drilled close to the classic outcrop sections of the Yorkshire coast, demonstrate geochemical, mineralogical and grain-size trends linked to sea level and climate change in the Cleveland Basin. High-resolution correlation between the core and outcrop enables the integration of data to generate a comprehensive chemostratigraphic record. Palaeoredox proxies (Mo, U, V, TOC/P, DOP and Fe speciation) show a progressive shift from oxic bottom waters in the late Pliensbachian through dysoxic–anoxic conditions in the earliest Toarcian to euxinia during the T-OAE. Anoxia–dysoxia persisted into the middle Toarcian. Elemental and isotope data (Re, Re/Mo, δ34SCAS, δ98Mo and ε205Tl) from the coastal sections evidence global expansion of anoxic and euxinic seafloor area driving drawdown of redox-sensitive metals and sulfate from seawater leading to severe depletion in early Toarcian ocean water. The record of anoxia–euxinia in the Cleveland Basin largely reflects global-scale changes in ocean oxygenation, although metal depletion was temporarily enhanced by periods of local basin restriction. Osmium and Sr isotopes demonstrate a pulse of accelerated weathering accompanying the early Toarcian hyperthermal, coincident with the T-OAE. The combined core and outcrop records evidence local and global environmental change accompanying one of the largest perturbations in the global carbon cycle during the last 200 Ma and a period of major biotic turnover.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1. Introduction

The early Toarcian (183.73 – 181.17 Ma; Gradstein et al., Reference Gradstein, Ogg, Schmitz and Ogg2020; Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022) was a period of dramatic global palaeoenvironmental change and biotic turnover associated with one of the largest perturbations in the global carbon cycle of the last 200 Ma, evidenced by a large (∼3 – 7‰ Vienna Peedee Belemnite, VPDB) negative carbon isotope excursion (CIE) in marine carbonate (δ13Ccarb) and marine and terrestrial organic carbon (δ13Corg) records (Küspert, Reference Küspert, Einsele and Seilacher1982; Jenkyns & Clayton, Reference Jenkyns and Clayton1997; Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000, Reference Hesselbo, Jenkyns, Duarte and Oliveira2007; Suan et al., Reference Suan, van de Schootbrugge, Adatte, Fiebig and Oschmann2015; Them et al., Reference Them, Gill, Caruthers, Gröcke, Tulsky, Martindale, Poulton and Smith2017a; Fantasia et al., Reference Fantasia, Föllmi, Adatte, Bernárdez, Spangenberg and Mattioli2018a; Cramer & Jarvis, Reference Cramer, Jarvis, Gradstein, Ogg and Ogg2020; Jin et al., Reference Jin, Shi, Baranyi, Kemp, Han, Luo, Hu, He, Chen and Preto2020; Kemp et al., Reference Kemp, Selby and Izumi2020; Ruebsam & Al-Husseini, Reference Ruebsam and Al-Husseini2020; Nie et al., Reference Nie, Fu, Liang, Wei, Chen, Lin, Zeng, Wu, Zou and Mansour2023; Richey et al., Reference Richey, Nordt, White and Breecker2023; Gambacorta et al., Reference Gambacorta, Brumsack, Jenkyns and Erba2024) and commonly referred to as the Toarcian Oceanic Anoxic Event (T-OAE). The T-OAE was associated with widespread deposition of carbonaceous (typically ∼5 – 10% total organic carbon, TOC) marine mudstones (‘black shales’) (Jenkyns, Reference Jenkyns1985, Reference Jenkyns1988; Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b), linked to the development of regional bottom-water anoxia – euxinia. These conditions are typified by the basins of NW Europe where major black shale successions include the Jet Rock (England), Posidonia Shale (Posidonienschiefer; Germany, Netherlands and North Sea) and Schistes carton (France), but anoxia is now known to have extended more widely, including parts of the deep Panthalassa Ocean (Kemp et al., Reference Kemp, Chen, Cho, Algeo, Shen and Ikeda2022a; Chen et al., Reference Chen, Kemp, He, Newton, Xiong, Jenkyns, Izumi, Cho, Huang and Poulton2023).

The CIE has been ascribed to the release of large volumes of isotopically light carbon from the Karoo, Ferrar and Chon Aike Large Igneous Provinces (LIPs; Pálfy & Smith, Reference Pálfy and Smith2000; Percival et al., Reference Percival, Witt, Mather, Hermoso, Jenkyns, Hesselbo, Al-Suwaidi, Storm, Xu and Ruhl2015; Guex et al., Reference Guex, Pilet, Müntener, Bartolini, Spangenberg, Schoene, Sell and Schaltegger2016; Remirez & Algeo, Reference Remirez and Algeo2020; Bergman et al., Reference Bergman, Eldrett, Minisini, Ernst, Dickson and Bekker2021; Heimdal et al., Reference Heimdal, Goddéris, Jones and Svensen2021; Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022; Fleischmann et al., Reference Fleischmann, Picotti, Caves Rugenstein, Cobianchi and Bernasconi2022; Jiang et al., Reference Jiang, Jourdan, Olierook and Merle2023). The acceleration of Karoo–Ferrar LIPs activity and subsequent peak magmatism agree closely with the onset of early Toarcian warming and the T-OAE (Greber et al., Reference Greber, Davies, Gaynor, Jourdan, Bertrand and Schaltegger2020; Gaynor et al., Reference Gaynor, Svensen, Polteau and Schaltegger2022; Luttinen et al., Reference Luttinen, Kurhila, Puttonen, Whitehouse and Andersen2022; Jiang et al., Reference Jiang, Jourdan, Olierook and Merle2023; Ware et al., Reference Ware, Jourdan and Timms2023; Fendley et al., Reference Fendley, Frieling, Mather, Ruhl, Hesselbo and Jenkyns2024). It is argued that the release of volcanic CO2 drove global warming (cf. Jenkyns, Reference Jenkyns1999), with the rise in surface temperature (Gómez et al., Reference Gómez, Comas-Rengifo and Goy2016) causing dissociation of terrestrial and seafloor methane clathrates (cf. Dickens et al., Reference Dickens, Oneil, Rea and Owen1995; Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005), CH4 release and further warming (Suan et al., Reference Suan, Nikitenko, Rogov, Baudin, Spangenberg, Knyazev, Glinskikh, Goryacheva, Adatte, Riding, Follmi, Pittet, Mattioli and Lecuyer2011), an enhanced hydrological cycle (Chen et al., Reference Chen, Kemp, He, Huang, Jin, Xiong and Newton2021), increased rates of continental weathering and soil erosion (Brazier et al., Reference Brazier, Suan, Tacail, Simon, Martin, Mattioli and Balter2015; Ruebsam et al., Reference Ruebsam, Muller, Kovacs, Palfy and Schwark2018, Reference Ruebsam, Pienkowski and Schwark2020a; Kemp et al., Reference Kemp, Selby and Izumi2020), permafrost and glacier destabilization (Krencker et al., Reference Krencker, Lindström and Bodin2019; Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019), increased wetland methanogenesis (Them et al., Reference Them, Gill, Caruthers, Gröcke, Tulsky, Martindale, Poulton and Smith2017a), eustatic sea-level rise and marine transgression (Suan et al., Reference Suan, Mattioli, Pittet, Lecuyer, Sucheras-Marx, Duarte, Philippe, Reggiani and Martineau2010; Remirez & Algeo, Reference Remirez and Algeo2020; Reolid et al., Reference Reolid, Mattioli, Duarte, Ruebsam, Reolid, Mattioli, Duarte and Ruebsam2021). Thermogenic methane release from coals in the Karoo–Ferrar LIPs has also been postulated (McElwain et al., Reference McElwain, Wade-Murphy and Hesselbo2005; Heimdal et al., Reference Heimdal, Goddéris, Jones and Svensen2021), although the viability of this process to generate the observed negative δ13C excursion has been contested (Gröcke et al., Reference Gröcke, Rimmer, Yoksoulian, Cairncross, Tsikos and van Hunen2009; Rahman et al., Reference Rahman, Rimmer and Rowe2018).

The combined effects of these extensive palaeoenvironmental changes led to a marine biotic crisis and global second-order mass extinction affecting multiple groups including ammonites, belemnites, brachiopods, corals, ostracods, benthic foraminifera and calcareous nannofossils (Hallam, Reference Hallam1986; Little & Benton, Reference Little and Benton1995; Wignall et al., Reference Wignall, Newton and Little2005; Dera et al., Reference Dera, Neige, Dommergues, Fara, Laffont and Pellenard2010; Caruthers et al., Reference Caruthers, Smith and Gröcke2013; Danise et al., Reference Danise, Twitchett and Little2015; Jiang et al., Reference Jiang, Song, Kemp, Dai and Liu2020; Piazza et al., Reference Piazza, Ullmann and Aberhan2020; Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023) and also significantly impacted terrestrial ecosystems (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019; Danise et al., Reference Danise, Slater, Vajda and Twitchett2022; Galasso et al., Reference Galasso, Feist-Burkhardt and Schneebeli-Hermann2022; Jin et al., Reference Jin, Zhang, Baranyi, Kemp, Feng, Grasby, Sun, Shi, Chen and Dal Corso2022; Baranyi et al., Reference Baranyi, Jin, Dal Corso, Shi, Grasby and Kemp2023, Reference Baranyi, Jin, Dal Corso, Li and Kemp2024).

In this paper, we present new geochemical results for the upper Pliensbachian – middle Toarcian of the Dove’s Nest borehole in North Yorkshire, England, that was drilled in the Cleveland Basin, close to the classic coastal exposures that provide a global Lower Jurassic reference section. Despite the importance of the area, controversy continues regarding the interpretation of palaeoenvironmental conditions in the basin before, during and after the T-OAE (e.g. Hesselbo et al., Reference Hesselbo, Little, Ruhl, Thibault and Ullmann2020a; Remírez & Algeo, Reference Remírez and Algeo2020), the significance of geochemical elemental and isotopic proxies obtained from the sections as indicators of global processes (van de Schootbrugge et al., Reference van de Schootbrugge, McArthur, Bailey, Rosenthal, Wright and Miller2005; McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Pearce et al., Reference Pearce, Cohen, Coe and Burton2008; Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; Hesselbo et al., Reference Hesselbo, Little, Ruhl, Thibault and Ullmann2020a; Chen et al., Reference Chen, Kemp, He, Newton, Xiong, Jenkyns, Izumi, Cho, Huang and Poulton2023) and even the identification of the early Toarcian perturbations as an OAE (McArthur, Reference McArthur2019; Them et al., Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022).

New data from the Dove’s Nest core, combined with a review of extensive published literature, offer an improved understanding of palaeoenvironmental change in the Cleveland Basin during the Early Jurassic and confirm the global significance of the Yorkshire succession as a reference for defining the causes and consequences of the T-OAE.

2. North Yorkshire coast – a global reference for the Lower Jurassic

With a history of more than 200 years of geological research, including the seminal papers of Young & Bird (Reference Young and Bird1822), Phillips (Reference Phillips1829) and Tate & Blake (Reference Tate and Blake1876), and subsequent detailed mapping by the Geological Survey (Fox-Strangways, Reference Fox-Strangways1892), the Lower Jurassic exposed along the North Yorkshire coast between Redcar and Ravenscar (Fig. 1) is one of the most extensively studied marine sedimentary rocks globally (Hemingway, Reference Hemingway, Rayner and Hemingway1974; Cope et al., Reference Cope, Getty, Howarth, Morton and Torrens1980; Rawson & Wright, Reference Rawson, Wright and Taylor1995; Powell, Reference Powell2010; Rawson & Wright, Reference Rawson and Wright2018). It includes the Global Boundary Stratotype Section and Point (GSSP) for the Pliensbachian Stage (Meister et al., Reference Meister, Aberhan, Blau, Dommergues, Feist-Burkhardt, Hailwood, Hart, Hesselbo, Hounslow, Hylton, Morton, Page and Price2006) and is the type section for the Lias Group (Cox et al., Reference Cox, Sumbler and Ivimey-Cook1999).

Figure 1. Early Jurassic palaeogeography, regional setting and location of the Dove’s Nest study core in the Cleveland Basin. (a) Palaeogeographic reconstruction of Europe showing the location of the basin on the European epicontinental shelf; interpreted bottom-water redox conditions associated with the T-OAE are based on geological data and ocean circulation modelling (Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018). (b) Global palaeogeography of the Early Jurassic showing continent configuration, major ocean basins and location of the Karoo–Ferrar Large Igneous Provinces (LIPs) that were emplaced during the early – middle Toarcian (Heimdal et al., Reference Heimdal, Goddéris, Jones and Svensen2021; Gaynor et al., Reference Gaynor, Svensen, Polteau and Schaltegger2022). Yellow box shows the location of the Europe map. Palaeogeographic base maps in (a) and (b) modified from Blakey (Reference Blakey2012, Reference Blakey2016); palaeolatitude in (a) revised based on the online palaeolatitude calculator of van Hinsbergen et al. (Reference van Hinsbergen, de Groot, van Schaik, Spakman, Bijl, Sluijs, Langereis and Brinkhuis2015) at 183 Ma (https://paleolatitude.org) with the palaeomagnetic reference frame of Vaes et al. (Reference Vaes, van Hinsbergen, van de Lagemaat, van der Wiel, Lom, Advokaat, Boschman, Gallo, Greve, Guilmette, Li, Lippert, Montheil, Qayyum and Langereis2023). (c) Map of eastern North Yorkshire showing the geographic distribution of Jurassic sediments in the Cleveland Basin, isopachs for the Lias and location of the Dove’s Nest borehole. Redrawn after Kent (Reference Kent1980) and Rawson & Wright (Reference Rawson and Wright2000).

Detailed lithostratigraphic descriptions accompanied by a precise ammonite biostratigraphy (Buckman, Reference Buckman, Fox-Strangways and Barrow1915; Howarth, Reference Howarth1955, Reference Howarth1962, Reference Howarth1973, Reference Howarth1992; Howarth in Cope et al., Reference Cope, Getty, Howarth, Morton and Torrens1980; Powell, Reference Powell1984; Howard, Reference Howard1985; Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995; Hesselbo & King, Reference Hesselbo, King and Lord2019) have provided a framework for extensive palaeontological (e.g. Little & Benton, Reference Little and Benton1995; Harries & Little, Reference Harries and Little1999; Caruthers et al., Reference Caruthers, Smith and Gröcke2013; Danise et al., Reference Danise, Twitchett, Little and Clémence2013, Reference Danise, Clemence, Price, Murphy, Gomez and Twitchett2019; Ullmann et al., Reference Ullmann, Thibault, Ruhl, Hesselbo and Korte2014; Lord, Reference Lord2019; De Baets et al., Reference De Baets, Nätscher, Rita, Fara, Neige, Bardin, Dera, Duarte, Hughes, Laschinger, Garcia-Ramos, Piñuela, Übelacker and Weis2021; Ferrari et al., Reference Ferrari, Little and Atkinson2021; Atkinson et al., Reference Atkinson, Little and Dunhill2023; Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023) and geochemical studies (e.g. Jenkyns & Clayton, Reference Jenkyns and Clayton1997; Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000; Cohen et al., Reference Cohen, Coe, Harding and Schwark2004; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005, Reference Kemp, Coe, Cohen and Weedon2011; McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Pearce et al., Reference Pearce, Cohen, Coe and Burton2008; Bond et al., Reference Bond, Wignall, Wang, Izon, Jiang, Lai, Sun, Newton, Shao, Vedrine and Cope2010; Littler et al., Reference Littler, Hesselbo and Jenkyns2010; Gill et al., Reference Gill, Lyons and Jenkyns2011; French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; McArthur, Reference McArthur2019; Houben et al., Reference Houben, Goldberg and Slomp2021; Q Li et al., Reference Li, McArthur, Thirlwall, Turchyn, Page, Bradbury, Weis and Lowry2021; Wang, Reference Wang2022; Kovács et al., Reference Kovács, Ruhl, Silva, McElwain, Reolid, Korte, Ruebsam and Hesselbo2024) that have focussed particularly on the interval spanning the Pliensbachian – Toarcian stage boundary and the T-OAE, and the associated environmental change and biotic turnover.

The upper Pliensbachian – middle Toarcian interval that constitutes the focus of the present study comprises a neritic – epeiric mudstone-dominated succession that accumulated in the Cleveland Basin (Fig. 1), a small, largely onshore, extension of the Sole Pit Basin of the Southern North Sea (Powell, Reference Powell2010). North Yorkshire was at that time located at ∼44° N on the western margin of a shallow epeiric sea bordering the Pennine–Caledonian High (Scottish) landmass. The Mid North Sea High, a Palaeozoic ridge north of the Cleveland Basin, and the Pennine High to the west were the likely sources of coarser siliciclastic sediment (Wright, Reference Wright2022), particularly during periods of sea-level lowstand such as in the late Pliensbachian (Bradshaw et al., Reference Bradshaw, Cope, Cripps, Donovan, Howarth, Rawson, West and Wimbledon1992; Hesselbo, Reference Hesselbo2008).

For biostratigraphic reference, we adopt the regional ammonite stratigraphy (Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995; Rawson & Wright, Reference Rawson and Wright2018). Following wider stratigraphic convention (cf. Hesselbo et al., Reference Hesselbo, Ogg, Ruhl, Hinnov, Huang, Gradstein, Ogg, Schmitz and Ogg2020b), we employ the ammonite zones and subzones as traditional biozones and not chronozones as promoted for the Jurassic (Page, Reference Page2017). The Toarcian lacks a standardized ammonite zonal and subzonal scheme, and significant differences exist in the index taxa and the stratigraphic placement of Pliensbachian – Toarcian ammonite zones on a regional to global scale. Compare, for example, the Yorkshire ammonite stratigraphy (Fig. 2) with that applied in the SW German Basin (e.g. Ruebsam et al., Reference Ruebsam, Schmid-Röhl and Al-Husseini2023), the Lusitanian Basin Portugal and the Neuquén Basin Argentina (cf. Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022), noting the differences in marker species and boundary positions relative to the reference carbon isotope curves. This challenges the view that the Yorkshire zones and subzones can be regarded as reliable chronostratigraphic units.

Figure 2. Stratigraphic log of the studied section of the Dove’s Nest core and correlation to a composite outcrop section along the North Yorkshire coast between Hawsker Bottoms and Port Mulgrave. Organic carbon isotopes (δ13Corg) and whole-rock total organic carbon (TOCWR) profiles are shown with their correlation (modified from Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022). Organic-rich facies associated with the large negative δ13Corg excursion of the Toarcian Oceanic Anoxic Event (T-OAE; Jenkyns, Reference Jenkyns1985) together with the negative δ13Corg excursion defining the Pliensbachian – Toarcian Boundary Event (Littler et al., Reference Littler, Hesselbo and Jenkyns2010) provide prominent tie points. The δ13Corg maximum of the A. gibbosus Subzone is equated to the Late Pliensbachian Event positive excursion of Korte & Hesselbo (Reference Korte and Hesselbo2011), De Lena et al. (Reference De Lena, Taylor, Guex, Bartolini, Adatte, van Acken, Spangenberg, Samankassou, Vennemann and Schaltegger2019) and Hollaar et al. (Reference Hollaar, Hesselbo, Deconinck, Damaschke, Ullmann, Jiang and Belcher2023). Grain size scale: fm, fine mudstone; mm, medium mudstone; cm, coarse mudstone; and fs, very fine sandstone. Yorkshire coast ‘bed’ numbers and named marker beds from Hawsker Bottoms (Fig. 1) for the Pliensbachian (Howarth, Reference Howarth1955) and Whitby composite section for the Toarcian (Howarth, Reference Howarth1962, Reference Howarth1973, Reference Howarth1992). Dove’s Nest data from Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) and this study. Yorkshire coast δ13Corg data from Hawsker Bottoms: orange, Littler et al. (Reference Littler, Hesselbo and Jenkyns2010); turquoise, Cohen et al. (Reference Cohen, Coe, Harding and Schwark2004); and pink, DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005). Port Mulgrave: dark blue, DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005); turquoise, Cohen et al. (Reference Cohen, Coe, Harding and Schwark2004). Saltwick Bay: turquoise, Cohen et al. (Reference Cohen, Coe, Harding and Schwark2004). TOC data for the Yorkshire coast are composite section values from Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011; thin green high-resolution curve), Ruvalcaba Baroni et al. (Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; green-filled triangles) and McArthur (Reference McArthur2019; green-filled circles). Thick green line shows the trend of the two low-resolution datasets. Vertical dotted lines and numbers are the δ13Corg reference value for average Phanerozoic black shale (Meyers, Reference Meyers2014) and the TOCWR content of average shale (Law, Reference Law, Beaumont and Foster1999) and average black shale (Vine & Tourtelot, Reference Vine and Tourtelot1970). SB2 and SB3 are the middle and upper Sulphur Bands of the basal lower Toarcian (Salem, Reference Salem2013; McArthur, Reference McArthur2019). Abbreviations of biostratigraphic zonation: H. bifrons = Hildoceras bifrons; H. serpentinum = Harpoceras serpentinum; D. tenui. = Dactylioceras tenuicostatum; P. spin. = Pleuroceras spinatum; A. margaritatus = Amaltheus margaritatus; D. commune = Dactylioceras commune; H. falciferum = Harpoceras falciferum; C. exa. = Cleviceras exaratum; Ds = Dactylioceras semicelatum; * = Dactylioceras tenuicostatum; † = Dactylioceras clevelandicum; Pp = Protogrammoceras paltum; Ph = Pleuroceras hawskerense; Pa = Pleuroceras apyrenum; A. gib. = Amaltheus gibbosus; As = Amaltheus subnodosus; A. stokesi = Amaltheus stokesi. Ages after GTS2020 (Gradstein et al., Reference Gradstein, Ogg, Schmitz and Ogg2020) with revisions of Al-Suwaidi et al. (Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022). Chemostratigraphic units modified from Remírez & Algeo (Reference Remírez and Algeo2020) and defined by multi-element proxies (see text); note that a – d, to the left of the TOCWR profile for the Yorkshire coast, are subunits of Unit III, the T-OAE.

The upper Pliensbachian (Amaltheus margaritatus and Pleuroceras spinatum zones) comprises silty mudstones with hummocky cross-stratified fine sandstones and thin sideritic and berthierine (chamosite)-rich ooidal ironstones (Fig. 2; uppermost Staithes Sandstone and Cleveland Ironstone formations). The overlying lower – middle Toarcian mudstones (Whitby Mudstone Formation) include an interval (upper Dactylioceras semicelatum and Cleviceras exaratum subzones) that is characterized by very dark brown carbonaceous mudstones (terminology of Lazar et al., Reference Lazar, Bohacs, Macquaker, Schieber and Demko2015; traditionally referred to as ‘black shales’) of the uppermost Grey Shale and lower Mulgrave Shale (Jet Rock), which contain the large negative carbon-isotope excursion that accompanies the T-OAE (Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005). Fissile, dark grey carbonaceous mudstones (Bituminous Shales Harpoceras falciferum Subzone) and then paler-coloured silty mudstones (Alum Shale Member Hildoceras bifrons Zone) cap the study succession. The higher part of the Alum Shale (post Dactylioceras commune Subzone) and the uppermost members of the Whitby Mudstone (Peak Mudstone and Fox Cliff Siltstone) are not considered here.

Correlation of the coastal sections (Fig. 1), at a decimetre scale over distances of tens of km, has been demonstrated using numbered and/or named lithostratigraphic marker ‘beds’, principally layers of ironstone, sideritic concretions and other calcareous concretions with distinctive shapes, sizes and mineralization, constrained by ammonite biostratigraphy (e.g. Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995; Rawson & Wright, Reference Rawson and Wright2018). However, understanding of these correlations is hampered by the retention of unique ‘bed’ numbering schemes for different reference sections (Howarth, Reference Howarth1955, Reference Howarth1962, Reference Howarth1973, Reference Howarth1992). A correlation of these schemes was provided by Cope et al. (Reference Cope, Getty, Howarth, Morton and Torrens1980) and a graphical correlation of key sections has been presented by Caswell & Herringshaw (Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023, fig. 2).

Additionally, as noted by Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022), although the numbered ‘beds’ are the finest published stratigraphic subdivisions of the middle and upper Lias of Yorkshire and have been used by most authors studying the succession, the ‘beds’ are neither genetic beds sensu Campbell (Reference Campbell1967) nor geometric beds sensu McKee & Weir (Reference McKee and Weir1953). Beds may be defined, relative to laminae, based on their thickness (cf. McKee and Weir, Reference McKee and Weir1953; Tucker, Reference Tucker2011). However, the key limitation of this descriptive approach is that it does not distinguish between strata that are ‘separated from adjacent strata by surfaces of erosion, non-deposition, or abrupt change in character’ (McKee and Weir, Reference McKee and Weir1953, p. 382) and strata, whatever their thickness may be, that were formed by accretion of sediment onto surfaces of net sedimentation without intervening episodes of non-deposition, which is often accompanied by erosion (Campbell, Reference Campbell1967). For this reason, beds in a sedimentological sense (i.e. sensu Campbell, Reference Campbell1967) have no limiting thickness. Stratification within a bed should be called lamination ‘whatever the thickness of the individual layers’ (Kuenen Reference Kuenen1966, p. 527).

‘Beds’ in the Lower Jurassic of northeast England were defined by Howarth (Reference Howarth1955, Reference Howarth1962, Reference Howarth1973) based principally on the presence of calcareous concretions at regular intervals. The concretionary horizons coincide with horizons of faunal change, which were used to subdivide the succession into ammonite zones and subzones. Many of these ‘beds’ are in the order of metres thick and thus three orders of magnitude larger than bedding in the succession (Trabucho-Alexandre, Reference Trabucho-Alexandre, Smith, Bailey, Burgess and Fraser2015). They are, therefore, bedsets comprising thin beds of mudstone, which may constitute parasequence-scale successions (Macquaker and Taylor, Reference Macquaker and Taylor1996). Such ‘beds’ represent longer timescales (and processes) of sedimentation.

As illustrated by Caswell & Herringshaw (Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023, fig. 2), an individual marker level may have been attributed to several ‘bed’ numbers and may lie within ‘beds’ that have boundaries defined using different criteria in different places; a single ‘bed’ at one site may be equivalent to multiple differently numbered ‘beds’ elsewhere. It is essential therefore to always refer to the specific locality numbering scheme being applied when describing the stratigraphy.

The GSSP of the Toarcian Stage is located at the Peniche section in the Lusitanian Basin of Portugal (Fig. 1; da Rocha et al., Reference da Rocha, Mattioli, Duarte, Pittet, Elmi, Mouterde, Cabral, Comas-Rengifo, Gomez, Goy, Hesselbo, Jenkyns, Littler, Mailliot, de Oliveira, Osete, Perilli, Pinto, Ruget and Suan2016). There are no formally designated auxiliary sections, but the correlation between the Peniche and Almonacid de la Cuba section in the Iberian Range (Spain) provides additional ammonite records and a magnetostratigraphy across the boundary interval that offer additional criteria for supraregional correlation (Comas-Rengifo et al., Reference Comas-Rengifo, Arias, Gómez, Goy, Herrero, Osete and Palencia2010; de la Rocha et al., Reference da Rocha, Mattioli, Duarte, Pittet, Elmi, Mouterde, Cabral, Comas-Rengifo, Gomez, Goy, Hesselbo, Jenkyns, Littler, Mailliot, de Oliveira, Osete, Perilli, Pinto, Ruget and Suan2016). However, both successions have a strong Tethyan influence and were characterized throughout the early Toarcian by generally low primary productivity and deposition in oxic – dysoxic bottom waters. This is in stark contrast to the Boreal higher productivity and anoxic – euxinic environment that developed progressively in the Cleveland Basin and more widely in northern European basins during the earliest Toarcian, reaching a peak during the T-OAE but commonly extending into the middle Toarcian. Chemostratigraphy and macrofossil biostratigraphy provide a basis for correlation between the two areas at high stratigraphic resolution enabling the extensive geochemical data from Yorkshire to be tied to the GSSP (e.g. Ait-Itto et al., Reference Ait-Itto, Martinez, Price and Addi2018; Fantasia et al., Reference Fantasia, Adatte, Spangenberg, Font, Duarte and Föllmi2019).

The Pliensbachian – Toarcian boundary is well exposed at Hawsker Bottoms in Yorkshire (Fig. 1; Littler et al., Reference Littler, Hesselbo and Jenkyns2010), which offers a candidate for a Standard Auxiliary Boundary Stratotype (Head et al., Reference Head, Aubry, Piller and Walker2023) to represent the Boreal sections of northern Europe.

3. Carbon isotopes, TOC and the T-OAE

Studies by Küspert (Reference Küspert, Einsele and Seilacher1982), Jenkyns & Clayton (Reference Jenkyns and Clayton1997) and Hesselbo et al. (Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000) established the Yorkshire coast Toarcian succession as a global reference section for the large negative CIE associated with carbonaceous mudstones that comprise the T-OAE. As originally described, the isotopic expression of the T-OAE was considered to be the positive CIE observed in Tethyan pelagic carbonates (δ13Ccarb) of the H. falciferum Zone (= H. serpentinum Zone of this study) and its equivalents (Jenkyns, Reference Jenkyns1985, Reference Jenkyns1988) rather than the immediately underlying negative excursion, which was regarded as being of potential diagenetic origin (Jenkyns & Clayton, Reference Jenkyns and Clayton1997).

Subsequent work demonstrated that a positive δ13Corg excursion is weakly developed or absent in many curves derived from the analysis of bulk organic matter (e.g. Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000). However, a large negative δ13C excursion (typically 5 – 7‰ δ13Corg in NW Europe) is observed globally in marine carbonate, terrestrial and marine bulk organic matter and individual biomarkers (e.g. Suan et al., Reference Suan, Nikitenko, Rogov, Baudin, Spangenberg, Knyazev, Glinskikh, Goryacheva, Adatte, Riding, Follmi, Pittet, Mattioli and Lecuyer2011; Caruthers et al., Reference Caruthers, Smith and Gröcke2014; Ikeda et al., Reference Ikeda, Hori, Ikehara, Miyashita, Chino and Yamada2018; Remirez & Algeo, Reference Remirez and Algeo2020; Ruebsam & Al-Husseini, Reference Ruebsam and Al-Husseini2020; Richey et al., Reference Richey, Nordt, White and Breecker2023; Huang et al., Reference Huang, Jin, Pancost, Kemp and Naafs2024), situated immediately below or superposed on the positive excursion, where developed (Fig. 3). The interval of the negative excursion alone is now commonly considered to represent the T-OAE (e.g. Cohen et al., Reference Cohen, Coe, Harding and Schwark2004; Them et al., Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018; Remirez & Algeo, Reference Remirez and Algeo2020; Bodin et al., Reference Bodin, Fantasia, Krencker, Nebsbjerg, Christiansen and Andrieu2023).

Figure 3. Carbon isotope correlation of selected European Pliensbachian – Toarcian successions. The map (bottom right) shows the palaeogeographic location of the sites (see Fig. 1 for details). Cleveland Basin δ13Corg profiles from this study (Dove’s Nest = black, coast composite = grey; see Fig. 2 for sources). The top of the D. commune Subzone lies ∼16 m above the top of the Hard Shales on the Yorkshire coast (Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995). Mochras δ13Corg data from Xu et al. (Reference Xu, Ruhl, Jenkyns, Leng, Huggett, Minisini, Ullmann, Riding, Weijers, Storm, Percival, Tosca, Idiz, Tegelaar and Hesselbo2018) and Storm et al. (Reference Storm, Hesselbo, Jenkyns, Ruhl, Ullmann, Xu, Leng, Riding and Gorbanenko2020); δ13Ccarb after Ullmann et al. (Reference Ullmann, Szȕcs, Jiang, Hudson and Hesselbo2022). CIEs as Figure 2 with Stokesi Event of Peti et al. (Reference Peti, Thibault, Clemence, Korte, Dommergues, Bougeault, Pellenard, Jelby and Ullmann2017) and Storm et al. (Reference Storm, Hesselbo, Jenkyns, Ruhl, Ullmann, Xu, Leng, Riding and Gorbanenko2020). Sancerre δ13Corg data from Hermoso et al. (Reference Hermoso, Minoletti and Pellenard2013) and Peti et al. (Reference Peti, Thibault, Korte, Ullmann, Cachão and Fibæk2021); δ13Ccarb after Hermoso et al. (Reference Hermoso, Le Callonnec, Minoletti, Renard and Hesselbo2009a, Reference Hermoso, Minoletti, Le Callonnec, Jenkyns, Hesselbo, Rickaby, Renard, de Rafelis and Emmanuel2009b; Reference Hermoso, Minoletti and Pellenard2013) and Peti et al. (Reference Peti, Thibault, Korte, Ullmann, Cachão and Fibæk2021). Pliensbachian biostratigraphy follows Peti et al. (Reference Peti, Thibault, Clemence, Korte, Dommergues, Bougeault, Pellenard, Jelby and Ullmann2017, Reference Peti, Thibault, Korte, Ullmann, Cachão and Fibæk2021) and Zhang et al. (Reference Zhang, Kemp, Thibault, Jelby, Li, Huang, Sui, Wang, Liu and Jia2023). Peniche δ13Corg profile from Fantasia et al. (Reference Fantasia, Adatte, Spangenberg, Font, Duarte and Föllmi2019). Peniche Pliensbachian δ13Ccarb values after Oliveira et al. (Reference Oliveira, Rodriguez, Duarte and Lemos2006) with stratigraphic revisions and additional data from Silva et al. (Reference Silva, Duarte, Comas-Rengifo, Mendonça Filho and Azerêdo2011); Pliensbachian – Toarcian boundary and Toarcian δ13Ccarb after Hesselbo et al. (Reference Hesselbo, Jenkyns, Duarte and Oliveira2007). Yorkshire stratigraphy follows Figure 2. Other abbreviations: PlToBE = Pliensbachian – Toarcian Boundary Event; H. falcif. = Harpoceras falciferum; Dc = Dactylioceras commune; Pf = Peronoceras fibulatum; Cc = Catacoeloceras crissum; D. ten. = Dactylioceras tenuicostatum; Ast. = Amaltheus stokesi; Ps = Pleuroceras spinatum; Dp = Dactylioceras polymorphum; H. levisoni = Hildaites levisoni.

Recently, it has been proposed that the interval of the negative CIE would be better termed the Jenkyns Event and the use of T-OAE should be expanded to include intervals of higher δ13C values in the underlying upper D. tenuicostatum Zone and overlying lower H. falciferum Subzone (Erba et al., Reference Erba, Cavalheiro, Dickson, Faucher, Gambacorta, Jenkyns and Wagner2022), closer to the original concept of the T-OAE (Jenkyns, Reference Jenkyns1985, Reference Jenkyns1988) as a positive δ13C event. However, views differ widely on the use of the two terms (compare Müller et al., Reference Müller, Price, Bajnai, Nyerges, Kesjár, Raucsik, Varga, Judik, Fekete, May and Pálfy2017; Jin et al., Reference Jin, Shi, Baranyi, Kemp, Han, Luo, Hu, He, Chen and Preto2020; Reolid et al., Reference Reolid, Mattioli, Duarte and Marok2020; Erba et al., Reference Erba, Cavalheiro, Dickson, Faucher, Gambacorta, Jenkyns and Wagner2022; Gambacorta et al., Reference Gambacorta, Cavalheiro, Brumsack, Dickson, Jenkyns, Schnetger, Wagner and Erba2023, Reference Gambacorta, Brumsack, Jenkyns and Erba2024). Spatio-temporal variability in the deposition of organic-rich strata during the early Toarcian and their relationship to marine anoxia and the negative δ13C excursion have been reviewed by Ruebsam & Schwark (Reference Ruebsam and Schwark2024). In the absence of a consensus, we follow Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) in applying the term T-OAE to the interval of the negative δ13C excursion (Figs 2, 3).

The carbon isotope profile for the upper Pliensbachian – middle Toarcian of the Yorkshire coast (Fig. 2), one of the highest resolution global organic carbon (δ13Corg) records, comprises a composite of separate datasets derived from bulk organic matter sampled from different coastal sections (Cohen et al., Reference Cohen, Coe, Harding and Schwark2004; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005; Littler et al., Reference Littler, Hesselbo and Jenkyns2010). Cohen et al. (Reference Cohen, Coe, Harding and Schwark2004) analysed samples from the Dactylioceras tenuicostatum to lowermost H. bifrons zones; their data are a composite from Hawsker Bottoms, Port Mulgrave and Saltwick Bay (Figs 1, 2). DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005) analysed samples from the D. tenuicostatum and lower Harpoceras serpentinum zones; their data are a composite from Hawsker Bottoms and Port Mulgrave spliced at ‘bed’ 33 (Fig. 2). Littler et al. (Reference Littler, Hesselbo and Jenkyns2010) presented data from the P. spinatum and D. tenuicostatum zones (Pliensbachian–Toarcian boundary) at Hawsker Bottoms.

Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) included some additional δ13Corg data for the H. serpentinum and H. bifrons zones in their summary curve compiled from the above sources. We have not included these data in our reference profile (Fig. 2) due to uncertainties in the sample positions and splicing. Van de Schootbrugge et al. (Reference van de Schootbrugge, Houben, Ercan, Verreussel, Kerstholt, Janssen, Nikitenko and Suan2020) used published data to construct a Yorkshire coast composite record and incorporated new low-resolution δ13Corg results from below and above previously published curves. Their curve extends below and above our study interval, including values from the lower and upper Pliensbachian (Prodactylioceras davoeiA. margaritatus zones) and from the middle Toarcian (H. bifrons Zone to above the Grammoceras thouarsense Zone) but lacks stratigraphic resolution comparable to this study.

Figure 4. Geochemical profiles for lithofacies proxies Al2O3 (aluminosilicates, principally clay minerals), CaCO3e (carbonates; calcite, siderite) and TOCWR (organic fraction) through the upper Pliensbachian – middle Toarcian of the Dove’s Nest core, with selected detrital proxies. ‘Bed’ numbers, names and biostratigraphy are derived from chemostratigraphic correlation to Hawsker Bottoms for the Pliensbachian (Howarth, Reference Howarth1955) and a Whitby composite section for the Toarcian (Howarth, Reference Howarth1962, Reference Howarth1973, Reference Howarth1992): red, sideritic beds; blue, limestones; and black other beds (see Fig. 2). Vertical dotted lines and numbers are reference values for Post-Archean Average Shale (PASS; = average mud of Taylor & McLennan, Reference Taylor, McLennan and Meyers2001). Prominent limestone ‘bed’ 35 (Whale Stones) and ‘beds’ 39 – 40 (Top Jet Dogger and Millstones) are clearly expressed by their high CaCO3e contents. Significant shifts in the elemental (Figs S1, S2) and element-ratio profiles (Figs 4, 5) combined with coincident changes in δ13Corg and TOC (Fig. 2) are used to define the chemostratigraphic units (see text), modified from the scheme of Remírez & Algeo (Reference Remírez and Algeo2020). LPlE = Late Pliensbachian Event; other abbreviations as Figure 2. Detrital proxies show multiple stacked CU cycles superimposed on a longer-term fining-upward trend through the top Staithes Sandstone to mid-Cleveland Ironstone, followed by a marked upward increase in grain size comprising 3 stacked CU cycles (cf. Macquaker & Taylor, Reference Macquaker and Taylor1996). Cycle 3 is most prominent and coincides with an interval of high δ13Corg and TOCWR values ascribed to the LPlE. The base of the Whitby Mudstone is a sharp facies break to clay-mineral dominated sediments illustrated by a steep rise in K/Al, Rb/Al and Cs/Al (not plotted) ratios. The T-OAE is expressed by a sharp increase and peak in TOCWR (maximum 9.2%), although elevated organic matter contents continue upward through the Whitby Mudstone. Upward-coarsening, 405 ka cycles in the Jet Rock (Si/Al profile) derived from cycle analysis of the coastal section by Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018, fig. 2) are also displayed in the Dove’s Nest record.

Figure 5. Geochemical profiles for redox and productivity proxies through the upper Pliensbachian – middle Toarcian of the Dove’s Nest core. Stratigraphic framework as in Figure 4. SB2 – SB3 are the middle and upper Sulphur Bands, consisting of laminated pyritic carbonaceous mudstones.

Low-resolution terrestrial wood (δ13Cwood) and compound-specific δ13C records across the T-OAE interval at Hawsker Bottoms show negative CIEs that are smaller in magnitude (∼1.5 – 5‰ in Yorkshire) compared to the associated bulk organic δ13C records (Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000; French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014). Nonetheless, coincident negative excursions in biomarkers from both marine and terrestrial plants demonstrate a carbon cycle perturbation that affected both the atmospheric and marine systems.

In strong contrast to the δ13Corg data, bulk carbonate carbon isotopes (δ13Ccarb) show no clear stratigraphic pattern in the Yorkshire succession due to a pervasive and highly variable diagenetic overprint in both limestones and organic-rich siliciclastic rocks (Jenkyns & Clayton, Reference Jenkyns and Clayton1997). However, a small negative shift below a marked positive δ13Ccarb excursion at the top of the C. exaratum Zone is identifiable in scattered isotope data obtained from belemnite rostra (McArthur et al., Reference McArthur, Donovan, Thirlwall, Fouke and Mattey2000; Bailey et al., Reference Bailey, Rosenthal, McArthur, van de Schootbrugge and Thirlwall2003; Gill et al., Reference Gill, Lyons and Jenkyns2011; Ullmann et al., Reference Ullmann, Thibault, Ruhl, Hesselbo and Korte2014; compilation of Korte et al., Reference Korte, Hesselbo, Ullmann, Dietl, Ruhl, Schweigert and Thibault2015). A change in life habits has been proposed to explain the subdued nature of the δ13Ccarb signal recorded in belemnites which, nonetheless, preserve a strong negative δ13Corg excursion of ∼4‰ in the organic matrix of their rostra (Ullmann et al., Reference Ullmann, Thibault, Ruhl, Hesselbo and Korte2014).

Whole-rock Total Organic Carbon (TOCWR) data for the upper Pliensbachian – middle Toarcian in Yorkshire have been published by several authors, including Jenkyns & Clayton (Reference Jenkyns and Clayton1997), Hesselbo et al. (Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000), McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008), Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011), Ruvalcaba Baroni et al. (Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018) and McArthur (Reference McArthur2019). The Yorkshire data presented by the last three of these sources are plotted in Figure 2. A TOCWR maximum, with some values exceeding 10%, coincides with the T-OAE interval, a characteristic of many other NW European sections, although organic matter enrichment is not universal across Europe (Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018, fig. 3; Remirez & Algeo, Reference Remirez and Algeo2020, fig. 3; Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b, fig. 2).

4. Pliensbachian – Toarcian boundary

Following Howarth (Reference Howarth1973, Reference Howarth1992), the base of the Toarcian in North Yorkshire is placed at the bottom of the ‘Sulphur Band’ of Chowns (Reference Chowns1968), which lies between the stratigraphically highest Pleuroceras and the stratigraphically lowest Dactylioceras (Howarth in Cope et al., Reference Cope, Getty, Howarth, Morton and Torrens1980; Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995; Littler et al., Reference Littler, Hesselbo and Jenkyns2010; Hesselbo & King, Reference Hesselbo, King and Lord2019), marking the base of the D. tenuicostatum Zone, Protogrammoceras paltum Subzone (Fig. 2). The Sulphur Band is a 15-cm-thick bed of pyrite-rich finely laminated carbonaceous mudstone (‘black shale’) with very low-angle cross-lamination and broad scour-and-fill structures towards the top (Salem, Reference Salem2013). The bed displays a symmetrical grain-size variation from silt to planar laminated mud and back to silt. Sedimentary structures in the upper two-thirds of the bed have been partially destroyed by Chondrites and Diplocraterion burrows penetrating down from the overlying bioturbated siltstones. The Sulphur Band lies in the middle of ‘bed’ 43 of Hawsker Bottoms but is recognized as a discrete ‘bed’ 26 at Kettleness (Howarth, Reference Howarth1955). Subaerial weathering of pyrite in the mudstone generates the distinctive yellow ‘sulphur’ colour of the bed when exposed in outcrops.

Howarth (Reference Howarth1973; in Cope et al., Reference Cope, Getty, Howarth, Morton and Torrens1980) and Powell (Reference Powell1984) assigned Hawsker Bottoms ‘beds’ 43 and 44 to the Cleveland Ironstone Formation. According to this placement, the Pliensbachian – Toarcian boundary falls within the Kettleness Member, and many authors have followed this convention (e.g. McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Littler et al., Reference Littler, Hesselbo and Jenkyns2010; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; Hesselbo & King, Reference Hesselbo, King and Lord2019; McArthur, Reference McArthur2019). However, these two ‘beds’ consist of carbonaceous mudstone interbedded with coarse mudstone and calcareous nodules, which are characteristic of the overlying Grey Shale Member of the Whitby Mudstone. For this reason, Howard (Reference Howard1985) and Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) assigned these ‘beds’ to the Grey Shales and placed the base of the member at the top of ‘bed’ 42, a thin (13 cm) red ironstone. This is underlain by a 45-cm-thick calcareous sandstone (‘bed’ 41) at Hawsker Bottoms. We follow this definition (Fig. 2): ‘bed’ 43 contains the base of the Toarcian Stage at the base of the Sulphur Band, 33 cm above the base of the Grey Shale.

The Sulphur Band (also referred to as the ‘Lower Sulphur Band’ or ‘Sulphur Band 1’; Salem, Reference Salem2013; Agbi et al., Reference Agbi, Ozibo and Newton2015; McArthur, Reference McArthur2019) constitutes a carbon isotope minimum (< −28‰ δ13Corg) and TOCWR maximum (up to 6.8%) at the top of the first trough within a broad double negative δ13Corg excursion (Pliensbachian – Toarcian Boundary Event, Fig. 2) that spans Hawsker Bottoms ‘bed’ 40 to Port Mulgrave ‘bed’ 4 at Hawsker Bottoms (Littler et al., Reference Littler, Hesselbo and Jenkyns2010, fig. 2). The amplitude of the negative CIE spanning the stage boundary approaches 2.5‰ (Fig. 2). A second δ13Corg minimum of similar magnitude occurs 1 m above, in a higher interval of pyritic laminated black shale (Port Mulgrave ‘beds’ 2 – 3) in the lower P. paltum Subzone; this ‘Middle Sulphur Band’ (Salem, Reference Salem2013; Agbi et al., Reference Agbi, Ozibo and Newton2015) or ‘Sulphur Band 2’ (McArthur, Reference McArthur2019) contains up to 4% TOCWR.

A negative δ13C excursion, with a double peak resolved in higher-resolution records, spans the stage boundary in many other high-resolution carbonate and organic carbon records, including the Llanbedr (Mochras Farm) borehole (Wales; subsequently referred to as Mochras); the Lorraine Sub-basin (Luxembourg); Sancerre (France); Peniche (Portugal); Brasa and Cima Benon (Italy); Es Cosconar (Spain); Talghemt and Foum Tillicht (Morocco); Chacay Melehue (Argentina); and Inuyama (Japan) (Fig. 3; Caruthers et al., Reference Caruthers, Smith and Gröcke2014; Martinez et al., Reference Martinez, Krencker, Mattioli and Bodin2017; Ikeda et al., Reference Ikeda, Hori, Ikehara, Miyashita, Chino and Yamada2018; Rosales et al., Reference Rosales, Barnolas, Goy, Sevillano, Armendáriz and López-García2018; Boulila et al., Reference Boulila, Galbrun, Sadki, Gardin and Bartolini2019; Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019; Ruebsam & Al-Husseini, Reference Ruebsam and Al-Husseini2020; Menini et al., Reference Menini, Mattioli, Hesselbo, Ruhl, Suan, Reolid, Duarte, Mattioli and Ruebsam2021; Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022; Fleischmann et al., Reference Fleischmann, Picotti, Caves Rugenstein, Cobianchi and Bernasconi2022). This confirms the Pliensbachian – Toarcian Boundary Event (PlToBE) as being a key global chemostratigraphic marker, estimated to have had a duration of ∼ 200 ka (Ruebsam et al., Reference Ruebsam, Munzberger and Schwark2014; Martinez et al., Reference Martinez, Krencker, Mattioli and Bodin2017; Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022).

5. Elemental chemostratigraphy – previous studies

Several authors reported partial geochemical analyses of Yorkshire upper Pliensbachian – middle Toarcian iron ores during the mid-nineteen to mid-twentieth centuries, as summarized by Whitehead et al. (Reference Whitehead, Anderson, Wilson and Wray1952) and Gad (Reference Gad1966). The compositions of three mudstone samples from the Jet Rock and Alum Shales were published by Dunham (Reference Dunham1961). A more detailed geochemical study of the Yorkshire Lias was undertaken by Gad (Reference Gad1966) using wet chemical techniques, the main results of which were presented by Gad et al. (Reference Gad, Catt and Le Riche1969) and Catt et al. (Reference Catt, Gad, Le Riche and Lord1971). Major-element and semi-quantitative trace-element data were provided for 28 mudstones from upper Pliensbachian – middle Toarcian sections. Selected data were also included for a range of ironstones (Gad, Reference Gad1966; Catt et al., Reference Catt, Gad, Le Riche and Lord1971).

Subsequently, Pye & Krinsley (Reference Pye and Krinsley1986) published major- and trace-element data, determined by X-ray fluorescence (XRF), for 26 mudstone samples from 15 beds through the Grey Shale – Alum Shales; these data were plotted stratigraphically by Imber et al. (Reference Imber, Armstrong, Clancy, Daniels, Herringshaw, McCaffrey, Rodriguez, Trabucho-Alexandre and Warren2014). Maynard (Reference Maynard1986) and Aggett (Reference Aggett1990) reported microprobe analyses of berthierine and siderite from the Cleveland Ironstone at Staithes. Myers & Wignall (Reference Myers, Wignall, Leggett and Zuffa1987), Myers (Reference Myers, Young and Gordon Taylor1989) and Parkinson (Reference Parkinson1996) employed portable spectral gamma-ray spectrometry to generate U, K and Th profiles at 0.5 – 1 m resolution for the Grey Shale and Jet Rock, the Cleveland Ironstone Formation and the entire Pliensbachian – Toarcian of the Yorkshire coast, respectively. Young et al. (Reference Young, Aggett, Howard and Young1990) plotted some whole-rock and electron microprobe mineral data from the Cleveland Ironstone, and Mücke & Farshad (Reference Mücke and Farshad2005) published XRF whole-rock and electron microprobe mineral data from the five main ironstone seams. Worden et al. (Reference Worden, Utley, Butcher, Griffiths, Wooldridge, Lawan, Dowey, Osborne and Volk2020) presented QEMSCAN® scanning electron microscope – energy-dispersive X-ray spectroscopy (SEM–EDX) data demonstrating a predominance of ‘granular’ siderite and minor pore-filling Fe-rich clay, identified by X-ray diffraction (XRD) as berthierine, in a Cleveland Ironstone sample from an unspecified horizon.

Recent geochemical work, despite employing more sensitive instrumental methods and much higher sampling resolutions, has generally focused on limited suites of elements considered to provide palaeoenvironmental information, principally redox-sensitive trace metals (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Harding in Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; McArthur, Reference McArthur2019; Remírez & Algeo, Reference Remírez and Algeo2020; Houben et al., Reference Houben, Goldberg and Slomp2021) or isotopes (Jenkyns et al., Reference Jenkyns, Gröcke and Hesselbo2001; Cohen et al., Reference Cohen, Coe, Harding and Schwark2004; Pearce et al., Reference Pearce, Cohen, Coe and Burton2008; Nielsen et al., Reference Nielsen, Goff, Hesselbo, Jenkyns, LaRowe and Lee2011; Q Li et al., Reference Li, McArthur, Thirlwall, Turchyn, Page, Bradbury, Weis and Lowry2021) but has included some major-element studies (Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; Wang, Reference Wang2022). Most recently, Kovács et al. (Reference Kovács, Ruhl, Silva, McElwain, Reolid, Korte, Ruebsam and Hesselbo2024) compiled the elemental data of Percival et al. (Reference Percival, Witt, Mather, Hermoso, Jenkyns, Hesselbo, Al-Suwaidi, Storm, Xu and Ruhl2015) and Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) and supplemented these with new Hg, TOC, carbon isotope and XRF data for 24 samples from the Pliensbachian–Toarcian boundary interval at Staithes and Port Mulgrave.

Remírez & Algeo (Reference Remírez and Algeo2020) subdivided the upper Pliensbachian – middle Toarcian (upper Staithes Sandstone to lower Alum Shale, A. margaritatus – H. bifrons zones) of the coast into five chemostratigraphic units: I – V (Fig. 2). These were loosely defined as being based on a ‘similar patterns of variation in multiple proxies’, derived from data published by McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008), with new boron and gallium determinations. However, samples were characterized for a restricted number of constituents, generally those judged to have significance for the reconstruction of bottom-water and sediment redox conditions, palaeoproductivity and/or palaeosalinity. Comprehensive major and trace-element data were lacking.

Unit I of Remírez & Algeo (Reference Remírez and Algeo2020) comprises the Staithes Sandstone and Cleveland Ironstone formations, incorporating upper Pliensbachian strata (A. margaritatus and P. spinatum zones) and the lowest P. paltum Subzone of the basal Toarcian (Fig. 2). Unit II represents the lower part of the Grey Shale Member of the Whitby Mudstone Formation, comprising most of the D. tenuicostatum Zone, including the P. paltum, D. clevelandicum, D. tenuicostatum and mid-D. semicelatum subzones. Unit III contains the upper Grey Shale and the highly pyritic carbonaceous mudstones of the Jet Rock at the base of the Mulgrave Shale Member and extends from the mid-D. semicelatum Subzone to the top of the C. exaratum Subzone. It incorporates the large negative δ13Corg excursion in the Jet Rock that defines the T-OAE. Unit IV comprises the lower half of the Bituminous Shales in the middle Mulgrave Shale, representing the lower H. falciferum Subzone. Unit V constitutes the upper Bituminous Shales and overlying Alum Shale Member, comprising the upper H. falciferum Subzone and most of the overlying H. bifrons Zone (D. commune Subzone).

Units I – V of Remírez & Algeo (Reference Remírez and Algeo2020) equate to Zones 1 – 4 of McArthur (Reference McArthur2019): Units I, II = Zone 1; Unit III = Zone 2; Unit IV = Zone 3; and Unit V = Zone 4. These schemes show limited similarity to the geochemical ‘units’ derived by cluster analysis of hand-held XRF elemental data, TOC and δ13Corg values, in a study of the coastal sections by Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018), despite near-identical stratigraphic trends being apparent for constituents that are in common.

Below, we present a suite of elemental geochemical data derived from the upper Pliensbachian – middle Toarcian of the Dove’s Nest borehole and correlate these to published data derived from the classic Yorkshire coastal sections (Fig. 1c). The core provides a single vertical section through rocks that are directly comparable in facies and thickness to the adjacent outcrops (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022; Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023). Our study provides the first comprehensive elemental data suite for an extended Staithes Sandstone – Whitby Mudstone (Alum Shale) succession in North Yorkshire.

Major-element and selected trace-element data, combined with lithostratigraphy and carbon-isotope stratigraphy, provide the basis for the high-resolution correlation of the Dove’s Nest core to the coastal sections. The resulting integrated chemostratigraphic framework offers increased stratigraphic coverage and additional geochemical parameters to previous studies and provides new constraints on the environmental changes that occurred before, during and after the T-OAE in the Cleveland Basin.

6. Material and methods

6.a. Dove’s Nest core

The Dove’s Nest Farm North Shaft NS1 borehole (Fig. 1), located ∼5.5 km south of Whitby, North Yorkshire (NZ 89297 05434; 0.62480° W, 54.43650° N), was drilled in February 2013 by Fugro Geoservices UK for the AngloAmerican Woodsmith Project (formerly Sirius Minerals). The drill site is now situated in Woodsmith Mine (https://uk.angloamerican.com/the-woodsmith-project). The cored vertical 251 m borehole sampled a representative Lower – Middle Jurassic succession between 125 and 220 m with full recovery.

The stratigraphy of the upper Pliensbachian (top Staithes Sandstone Formation) to middle Toarcian (lower Alum Shale Member, Whitby Mudstone Formation) in the Dove’s Nest core has been described by Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) and the trace-fossil assemblages were logged by Caswell & Herringshaw (Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023, figs 2, 7). Technical details of the studied section of the core, including core box numbers and depth intervals, were provided in the Supplementary Material of Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022). The thickness of the middle and upper Lias Group formations in the core is approximately the same as that exposed on the coast between Hawsker Bottoms and Whitby. However, unlike the rocks exposed along the coast, the core provides a single, continuous vertical section through unweathered sedimentary rocks, with no structural complexity, and so it offers optimum potential for the preservation of primary geochemical signals.

Figure 6. Field photograph of the lower cliff face immediately west of the old harbour of Port Mulgrave (Fig. 1) annotated with the TOC, CaCO3 and δ13Corg data of DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005) and Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011) and the stratigraphic framework of Howarth (Reference Howarth1962). The five distinctive concretionary horizons in the Jet Rock, extinction level iii of Caswell et al. (Reference Caswell, Coe and Cohen2009) and chemostratigraphic units (IIIa – IVa) with key intervals of change are indicated. D. semi. = Dactylioceras semicelatum; C. exaratum = Cleviceras exaratum; H.f. = H. falciferum. Chemostratigraphic correlation to Dove’s Nest is illustrated in Figure 7. Note that the shore platform in the foreground occurs at the level of the base Jet Rock (‘bed’ 33). The sedimentary log and geochemical profiles below this (‘bed’ 32) represent variations in the subsurface at this site.

Figure 7. Chemostratigraphic correlation of the T-OAE interval in the Yorkshire coastal outcrop reference sections with the Dove’s Nest core. Stratigraphy after Howarth (Reference Howarth1962, Reference Howarth1973, Reference Howarth1992) and Howarth (in Cope et al., Reference Cope, Getty, Howarth, Morton and Torrens1980). Dt = Dactylioceras tenuicostatum Subzone. Lithological log is based on the Hawsker Bottoms and Port Mulgrave sections from DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005): lithologies are dark-grey laminated mudrocks (dark-grey shading), medium-grey mudrocks (pale-grey shading) and carbonate bands and nodules (brick pattern). Major carbonate markers – ‘Stone’ bands and ‘Doggers’ – are indicated. Sample heights of Hesselbo et al. (Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000, fig. 3) were recalculated based on the positions of major bed contacts. Data sources: Hesselbo et al. (Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000); DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005); Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011); Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018); Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022); this study. Si/Al and Ti/Al ratios for the coastal sections were recalculated by Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018 supplementary data) after correction for analytical bias. Shaded intervals a – d represent subdivisions of chemostratigraphic Unit III. This unit corresponds to the interval displaying the large negative carbon-isotope excursion that characterizes the T-OAE. Dashed horizontal grey lines show the correlation of major bed bases; dotted horizontal grey lines correlate significant chemostratigraphic tie points. Cyclostratigraphic filtered output for carbon isotopes (orange curve) and the detrital fraction (yellow curve, derived from Zr/Rb data) are plotted after Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018). Vertical dotted lines indicate: the δ13Corg value of average Phanerozoic black shales (grey; Meyers, Reference Meyers2014); the oxic–anoxic- and anoxic–euxinic-facies boundaries defined by TOC content (green, 2.5% and 10%) proposed by Algeo & Maynard (Reference Algeo and Maynard2004).

6.b. Sampling

The Dove’s Nest core was logged in 2013 (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) and 276 samples were collected for geochemical analysis at 20 – 50 cm intervals between 126.18 and 219.06 m depth. For this study, a chip with a mass of ∼1 g from each sample was taken and ground to a fine powder (c. 10 µm) using a Retsch agate mortar grinder RM100 at Durham University.

6.c. Geochemical analysis

Sample solutions were prepared from the dried powders using the lithium-metaborate fusion method of Jarvis (Reference Jarvis, Jarvis, Gray and Houk2003, section 7.3.3). The elemental composition of the solutions was determined by inductively coupled plasma-atomic emission spectrometry (ICP-AES) and ICP-mass spectrometry (ICP-MS) analyses using JY Ultima 2C and Agilent 7700x spectrometers at Kingston University London.

Major elements (Si, Ti, Al, Fe, Mn, Mg, Ca, Na, K and P) and selected trace elements (Ba, Cr, Sr and Zr) were determined by ICP-AES on the full sample suite (Atar, Reference Atar2015). The Ca data have been reported previously by Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022), together with carbon isotope values for bulk organic matter (δ13Corg) and TOCWR for the same samples. The last of these were calculated from acid-insoluble residue TOC values using CaCO3 values derived from ICP-AES Ca determinations, based on the assumption that most Ca resides in carbonate (referred to herein as CaCO3e). Elemental data are reported here as weight per cent oxides for the major elements and as ppm (parts per million; µg/g) for trace elements (Supplementary Material Table S1). Total Fe is expressed as Fe2O3T. By reference to international reference materials run with samples, accuracy and precision for major elements determined by ICP-AES were judged to be generally <3% for major elements and <5% for trace elements.

An additional trace-element suite (Cs, Rb, Sc, Th, U, V and Y) was determined by ICP-MS in a subset of 98 lithium-metaborate fusion solutions, chosen at ∼1 m intervals through the study interval (Supplementary Material Table S2). Splits of powders for this smaller sample suite were additionally prepared using partial digestion with aqua regia for the determination of Mo by ICP-MS (modified after Moor et al., Reference Moor, Lymberopoulou and Dietrich2001). The ICP-MS data reported here are derived from the instrument data files of Atar (Reference Atar2015) with results for elements yielding poor accuracy and/or reproducibility rejected and reported values (ppm, parts per million; µg/g) recalculated to correct some errors identified in the original document. Accuracy and precision for elements determined by ICP-MS were judged to be generally <10%. Results for Mo are reported as Mo* (Table S2) to reflect the greater uncertainty of these data, where many values are close to or below the limit of quantitative determination.

6.d. Data presentation

Stratigraphic variation in elemental chemistry can be obscured by the closed-sum nature of major-element percentage datasets, particularly where there are large fluctuations in a dominant mineral fraction. Increases in carbonate (Ca) and/or quartz (Si) will generate coincident falls in the values of most other constituents. Element ratios or normalization may be used to address this problem. Primary mineralogical controls on elements commonly used in chemostratigraphy of carbonaceous mudstones and element ratio grain-size and palaeoredox proxies have been summarized by LaGrange et al. (Reference LaGrange, Konhauser, Catuneanu, Harris, Playter and Gingras2020).

Aluminium and/or Rb are generally considered to be stable conservative elements and are favoured as the divisor to better reveal geochemical variation in the aluminosilicate and heavy mineral fractions or trace metals, independent of quartz or carbonate dilution (e.g. Algeo, Reference Algeo2004; Hermoso et al., Reference Hermoso, Minoletti and Pellenard2013; Fantasia et al., Reference Fantasia, Föllmi, Adatte, Bernárdez, Spangenberg and Mattioli2018a, b; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; Xu et al., Reference Xu, Ruhl, Jenkyns, Leng, Huggett, Minisini, Ullmann, Riding, Weijers, Storm, Percival, Tosca, Idiz, Tegelaar and Hesselbo2018; Bennett & Canfield, Reference Bennett and Canfield2020; Houben et al., Reference Houben, Goldberg and Slomp2021; Gambacorta et al., Reference Gambacorta, Cavalheiro, Brumsack, Dickson, Jenkyns, Schnetger, Wagner and Erba2023). Alternatively, normalization to a standard shale, typically Post-Archean Average Shale (PAAS) (Taylor & McLennan, Reference Taylor, McLennan and Meyers2001) or average upper continental crust (Rudnick & Gao, Reference Rudnick, Gao, Holland and Turekian2014) and expression of element concentrations as enrichment factors (EF), where EFelementX = [(X/Al)sample/(X/Al)PASS], are commonly used for plotting redox-sensitive trace-element data (e.g. Brumsack, Reference Brumsack2006; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; Algeo & Liu, Reference Algeo and Liu2020; Tribovillard, Reference Tribovillard2020; Yano et al., Reference Yano, Yasukawa, Nakamura, Ikehara and Kato2020). Both approaches are adopted here.

The Chemical Index of Alteration (CIA) of Nesbitt & Young (Reference Nesbitt and Young1982) was calculated using molecular proportions:

$${\rm{CIA}} = {\rm{[A}}{{\rm{l}}_2}{{\rm{O}}_{{\rm{3}}}}/\left( {{\rm{A}}{{\rm{l}}_{\rm{2}}}{{\rm{O}}_{\rm{3}}} + {\rm{CaO^*}} + {\rm{N}}{{\rm{a}}_{\rm{2}}}{\rm{O}} + {{\rm{K}}_{\rm{2}}}{\rm{O}}} \right)] \times 100$$

where CaO* is the amount of CaO incorporated in the silicate fraction of the rock, with a correction made for carbonate and apatite content.

Carbonate content varies substantially in the study section and is associated with a range of carbonate minerals including calcite, dolomite and siderite. CO2 contents were not determined, preventing direct carbonate correction. CaO* values were calculated using the method of McLennan (Reference McLennan1993): CaO was corrected for phosphate using the CaO/P2O5 value of 1.625 of unweathered marine carbonate fluorapatite (Jarvis et al., Reference Jarvis, Burnett, Nathan, Almbaydin, Attia, Castro, Flicoteaux, Hilmy, Husain, Qutawnah, Serjani and Zanin1994). If the remaining number of moles was less than that of Na2O, this CaO* value was adopted. If the number of moles was greater than Na2O, CaO* was assumed to be equivalent to Na2O. Since Ca is typically lost more rapidly than Na during weathering, this is likely to yield minimum CIA values, by up to about 3 units (McLennan, Reference McLennan1993).

Transition metal ratios (e.g. Ni/Co, V/Cr and V/Mo) have long been applied as palaeoredox proxies (e.g. Ernst, Reference Ernst1970; Jones & Manning, Reference Jones and Manning1994; Piper & Calvert, Reference Piper and Calvert2009), but Algeo & Liu (Reference Algeo and Liu2020) have demonstrated that they perform poorly compared to single element enrichment factors for such purposes. We provide limited new transition metal data here and generally do not make use of bimetal ratios of these. By contrast, the molar ratio of TOC to total phosphorus (PT), commonly referred to as the Corg/P ratio, has become established as a robust palaeoredox proxy in sediments and sedimentary rocks (e.g. Algeo & Ingall, Reference Algeo and Ingall2007; Algeo & Li, Reference Algeo and Li2020; Papadomanolaki et al., Reference Papadomanolaki, Lenstra, Wolthers and Slomp2022). Data obtained from Dove’s Nest may be compared to published results from elsewhere.

For the interpretation of bottom water redox condition, we use a four-stage classification scheme for deoxygenation modified from Algeo & Liu (Reference Algeo and Liu2020) comprising: oxic (>2 mL O2 L−1), dysoxic (2 to ∼0 mL O2 L−1), anoxic (∼0 mL O2 L−1, Fe2+ > 0, H2S = 0; also termed ‘anoxic-ferruginous’ or ‘suboxic’) and euxinic (0 mL O2 L−1, Fe2+ = 0, H2S > 0, also termed anoxic-euxinic) conditions. It is commonly difficult to unambiguously separate the redox state of bottom waters as opposed to those of the underlying surface or shallow subsurface sediments based on bulk geochemistry – particularly in the rock record. Sulfidic (‘euxinic’) conditions necessary for pyrite formation, for example, may occur in sediments accumulating under fully oxygenated water columns, although pyrite framboid size offers a potential criterion to identify water-column euxinia (Wilkin et al., Reference Wilkin, Barnes and Brantley1996; Wignall & Newton, Reference Wignall and Newton1998; Rickard, Reference Rickard2019).

The distinction between oxic and dysoxic facies in sediments is generally made based on body fossils and ichnofabric rather than geochemical criteria, but anoxic and euxinic redox facies are associated with geochemical reactions that can potentially leave diagnostic signals in the sediment. However, it is recognized that redox conditions in an area are typically subject to temporal variation on a range of time scales and of varying magnitude. This variation may generate mixed geochemical and other proxy records in the sedimentary record. Sedimentological evidence for at least weak or intermittent oxygenation in the form of fossil benthic fauna and bioturbation in conjunction with geochemical proxies indicating euxinic conditions is common (Algeo & Liu, Reference Algeo and Liu2020). This pattern may be due, for example, to the transient influx of oxygenated hyperpycnal flows into a basin or seasonal overturning of the water column, resulting in short-term colonization of an otherwise lifeless seafloor, or to longer term oxygenation events.

Different geochemical redox proxies respond at different thresholds to redox changes and acquire a signature characteristic of a specific redox state at different rates. It is essential therefore to combine data from multiple proxies to reliably identify the dominant redox conditions during the deposition of a bed.

6.e. Statistical analysis

Quantitative assessment of inter-element relationships was made by Pearson linear correlation and multivariate principal component analysis (PCA) of the geochemical data using Past 4.11 software (Hammer et al., Reference Hammer, Harper and Ryan2001). Constant-sum constraints of the elemental data (Aitchison, Reference Aitchison1986) were removed using a centred log-ratio (clr) transformation prior to statistical analysis.

6.f. Elemental chemostratigraphy

Published elemental profiles for the Yorkshire coastal outcrops (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; McArthur, Reference McArthur2019; Remírez & Algeo, Reference Remírez and Algeo2020; Wang, Reference Wang2022) together with carbon isotopes and TOC data (Jenkyns & Clayton, Reference Jenkyns and Clayton1997; Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000; Cohen et al., Reference Cohen, Coe, Harding and Schwark2004; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005; McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Littler et al., Reference Littler, Hesselbo and Jenkyns2010; Kemp et al., Reference Kemp, Coe, Cohen and Weedon2011; McArthur, Reference McArthur2019) provide the basis for a bed-scale correlation of the lower Whitby Mudstone Formation (lower Toarcian Dactylioceras tenuicostatum – lower H. bifrons zones) to the Dove’s Nest core. However, only very limited and widely spaced geochemical data are available for the upper Pliensbachian Staithes Sandstone and Cleveland Ironstone formations on the coast (Gad, Reference Gad1966; McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; McArthur, Reference McArthur2019; Remírez & Algeo, Reference Remírez and Algeo2020; Kovács et al., Reference Kovács, Ruhl, Silva, McElwain, Reolid, Korte, Ruebsam and Hesselbo2024), where the Kettleness Member in particular shows significant lithological and thickness variation between sections (e.g. Howarth, Reference Howarth1955 pl. 13). Chemostratigraphic correlation of the lower section of Dove’s Nest to the coast is only possible at member scale, but the core data offer new insights into the stratigraphy of the succession.

For consistency, we employ the five chemostratigraphic units (I – V) of Remírez & Algeo (Reference Remírez and Algeo2020) with minor modification and designate 14 subunits (including 4 subdivisions of the T-OAE interval: IIIa – d; Fig. 2) derived from the recognition of patterns in our more comprehensive and stratigraphically extended geochemical data from the Dove’s Nest core (Tables S1, S2).

7. Results – δ13Corg and TOC stratigraphy of Dove’s Nest

7.a. Upper Pliensbachian – lower Toarcian

Organic carbon isotope and TOCWR curves for the upper Pliensbachian – middle Toarcian of the Dove’s Nest core have been described by Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022, fig. 2). These data are reviewed here to provide a stratigraphic framework and context for the elemental chemostratigraphy.

Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) demonstrated that the carbon isotope and TOC profiles, when combined with lithostratigraphy, provide a sound basis for the correlation of the Dove’s Nest core with the Yorkshire coastal succession (Fig. 2). The data from the Dove’s Nest core are fully compatible with previous high-resolution studies from the Yorkshire coast, confirm a consistent regional chemostratigraphy, and offer new higher-resolution reference profiles for the upper Pliensbachian A. margaritatus – P. spinatum zones and for the lower middle Toarcian H. bifrons Zone.

Carbon isotopes and organic carbon display large amplitude variation through the study interval at Dove’s Nest: −31.5‰ to −24.5‰ δ13Corg and 0.2 – 9.2% TOCWR, respectively (Fig. 2). Employing the standard Yorkshire coast lithostratigraphy and biostratigraphy (Simms et al., Reference Simms, Chidlaw, Page, Morton and Gallois2004b; Hesselbo & King, Reference Hesselbo, King and Lord2019), key features of the long-term trends, from bottom to top, are:

  1. (1) The upper Staithes Sandstone Formation and Cleveland Ironstone Penny Nab Member (A. margaritatus Zone) show a somewhat erratic plateau in the δ13Corg profile, with values of −25.7 ± 0.5‰ (Fig. 2). Maximum values for the section of ∼ −25‰ occur at the top of the Penny Nab Member in the A. gibbosus Subzone. This maximum corresponds to the level of the Late Pliensbachian Event (LPlE) positive CIE (Korte & Hesselbo, Reference Korte and Hesselbo2011; De Lena et al., Reference De Lena, Taylor, Guex, Bartolini, Adatte, van Acken, Spangenberg, Samankassou, Vennemann and Schaltegger2019; Cramer & Jarvis, Reference Cramer, Jarvis, Gradstein, Ogg and Ogg2020; Hollaar et al., Reference Hollaar, Hesselbo, Deconinck, Damaschke, Ullmann, Jiang and Belcher2023; Zhang et al., Reference Zhang, Kemp, Thibault, Jelby, Li, Huang, Sui, Wang, Liu and Jia2023), although the δ13Corg peak at Dove’s Nest is small and likely truncated by the disconformity at the base of the Pecten Seam. A much larger positive excursion of δ13Cwood values (>4‰) has been recorded from the upper A. margitatus Zone of Robin Hood’s Bay (Korte & Hesselbo, Reference Korte and Hesselbo2011). TOCWR contents show a stepped rise from 0.7 ± 0.3% in the Staithes Sandstone to 1.7 ± 0.6% at the top of the Penny Nab.

  2. (2) Carbon isotope values decline from −24.8‰ to −26.9‰ δ13Corg through the Kettleness Member (P. spinatum Zone) and a marked double negative excursion with values falling to −29.6‰ spans the Cleveland Ironstone – Whitby Mudstone formation boundary. This is the PlToBE negative CIE (Littler et al., Reference Littler, Hesselbo and Jenkyns2010; Fantasia et al., Reference Fantasia, Adatte, Spangenberg, Font, Duarte and Föllmi2019; Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019; Cramer & Jarvis, Reference Cramer, Jarvis, Gradstein, Ogg and Ogg2020; Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022; Ullmann et al., Reference Ullmann, Szȕcs, Jiang, Hudson and Hesselbo2022; Bodin et al., Reference Bodin, Fantasia, Krencker, Nebsbjerg, Christiansen and Andrieu2023). TOCWR contents display a concave pattern (Fig. 2), falling from 1.8% at the base to a minimum of 0.6% in the middle of the Kettleness Member and then rising back to 1.5% at the top.

  3. (3) δ13Corg values plateau at 25.8 ± 0.3‰ in the Grey Shale Member (upper P. paltum – lower D. tenuicostatum subzones) above the PlToBE, accompanied by a TOCWR rise from 1.2% to 2.2%.

  4. (4) A sharp change in the geochemical profiles occurs with the appearance of black laminated carbonaceous mudstones (‘black shales’) in the middle of ‘bed’ 31 in the mid-D. semicelatum Subzone (Fig. 2). This marks the base of a broad large negative δ13Corg excursion in the upper Grey Shale and Jet Rock beds of the lower Mulgrave Shale Member, with values falling to −31.5‰, accompanied by a large TOCWR peak and values up to 9.2% (Fig. 2). This is an expression of the T-OAE (Section 3), which terminates at the top of the C. exaratum Subzone.

  5. (5) δ13Corg values rise and plateau at −27.4 ± 0.4‰ in the Bituminous Shales (H. falciferum Subzone), continuing upward into the lower Alum Shale (H. bifrons Zone) and are accompanied by generally falling TOCWR from 3.9% to 1.8%. A coincident small, stepped rise in δ13Corg and fall in TOCWR occurs in the upper Bituminous Shales above ‘bed’ 44 (Peak Stones) in the mid-H. falciferum Subzone (Fig. 2).

  6. (6) An interval of elevated TOC values bounded at the bottom and top by peaks of up to 4% occurs between ‘beds’ 48 (Ovatum Band) and 50, the interval of the Hard Shales (lower D. commune Subzone) at the base of the Alum Shale Member.

The most prominent features of the chemostratigraphy, therefore, are the presence of two intervals with negative δ13Corg excursions (Fig. 2): the PlToBE and the T-OAE, the latter accompanied by a thick interval of laminated carbonaceous mudstones with high TOC contents. Additionally, a large step down to lower δ13Corg and step up to higher TOC contents in the long-term profiles occur at the base of the T-OAE interval. It is notable that δ13Corg values below the T-OAE are consistently ∼1‰ higher than the average Phanerozoic black shale value of −27‰ (Meyers, Reference Meyers2014), while the succession above yields marginally lower values than this average.

TOCWR values in the lower section lie close to an average shale value of 0.8% (Law, Reference Law, Beaumont and Foster1999), while those above are significantly higher, at >2% (characteristic of a very good petroleum source rock). Average black shale TOCWR values of 3% (Vine & Tourtelot, Reference Vine and Tourtelot1970) are exceeded in the three laminated mudstones of the Sulphur Bands (SB1 – 3) in the Grey Shale (Fig. 2), throughout the upper Grey Shales to mid-Mulgrave Shale, and at the bottom and top of the Hard Shales.

7.b. Pliensbachian – Toarcian boundary

The amplitude of the PlToBE negative CIE approaches 4‰ δ13Corg at Dove’s Nest (Fig. 2). The Sulphur Band (SB1) is not obvious in the core, probably due to the lack of weathering. However, a distinctive bioturbated layer with abundant Chondrites burrows at ∼185 m is similar to that developed above the Sulphur Band in the top of ‘bed’ 43 at Kettleness and Runswick Bay (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022, fig. 4). The Pliensbachian–Toarcian boundary is therefore placed below the Chondrites interval, at 185.5 m in the core. This placement of the boundary is consistent with the δ13Corg record (Fig. 2): a large negative excursion is developed below the boundary (minimum −29.6‰ at 185.75 m) and a second negative excursion above it (minimum −27.8‰ at 184.41 m).

The δ13Corg minimum identified at Dove’s Nest occurs in an Fe-rich (12% Fe2O3) calcareous (8.9% CaCO3e) sample at 185.75 m that is only moderately enriched in TOCWR (1.5%). Coincident enrichment in Fe, Mg, Ca and Mn indicates the presence of siderite. Comparison to the succession at Hawsker Bottoms (Littler et al., Reference Littler, Hesselbo and Jenkyns2010, fig. 2) suggests that our sampling missed the Sulphur Band in the core, while sampling the sideritic ironstone a short distance below (Hawsker Bottoms ’bed’ 42), or the successions differ.

7.c. Toarcian Oceanic Anoxic Event (T-OAE)

The thickness (9.5 m) of the lower Toarcian δ13Corg negative excursion constituting the T-OAE in the Dove’s Nest core, which occurs between 175.1 and 165.2 m with δ13Corg values of −31.5‰ to −25.8‰ (Fig. 2), is very similar to that in the curves of Hesselbo et al. (Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000) and DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005), which were derived from the lower Toarcian outcrops at Hawsker Bottoms and Port Mulgrave (=mid-‘bed’ 31 to top ‘bed’ 38) on the Yorkshire coast. The shape of the excursion, which extends from the mid-D. semicelatum to upper C. exaratum subzones and the δ13Corg values, are most similar to those of Hesselbo et al. (Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000); values reported by DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005) are more 13C-depleted, down to −33.2‰.

The Dove’s Nest δ13Corg profile is overlain on the Yorkshire coast composite curve and correlated to a compilation of published upper Pliensbachian – middle Toarcian carbon isotope data (Oliveira et al., Reference Oliveira, Rodriguez, Duarte and Lemos2006; Hesselbo et al., Reference Hesselbo, Jenkyns, Duarte and Oliveira2007; Hermoso et al., Reference Hermoso, Le Callonnec, Minoletti, Renard and Hesselbo2009a, Reference Hermoso, Minoletti, Le Callonnec, Jenkyns, Hesselbo, Rickaby, Renard, de Rafelis and Emmanuel2009b, Reference Hermoso, Minoletti, Rickaby, Hesselbo, Baudin and Jenkyns2012, Reference Hermoso, Minoletti and Pellenard2013; Silva et al., Reference Silva, Duarte, Comas-Rengifo, Mendonça Filho and Azerêdo2011; Peti et al., Reference Peti, Thibault, Clemence, Korte, Dommergues, Bougeault, Pellenard, Jelby and Ullmann2017; Xu et al., Reference Xu, Ruhl, Jenkyns, Leng, Huggett, Minisini, Ullmann, Riding, Weijers, Storm, Percival, Tosca, Idiz, Tegelaar and Hesselbo2018; Fantasia et al., Reference Fantasia, Adatte, Spangenberg, Font, Duarte and Föllmi2019; Storm et al., Reference Storm, Hesselbo, Jenkyns, Ruhl, Ullmann, Xu, Leng, Riding and Gorbanenko2020; Ullmann et al., Reference Ullmann, Szȕcs, Jiang, Hudson and Hesselbo2022) for three well-characterized European sections in Figure 3. This exemplifies how the PlToBE and T-OAE offer robust tie points for correlation that complement the biostratigraphy.

The δ13Corg excursion interval at Dove’s Nest corresponds to a large TOCWR peak with values of up to 9.2% (Fig. 2), which extends from the base of the T-OAE interval to the top of the Jet Rock at 164.0 m, the top calcareous beds of which (Top Jet Dogger and Millstones, ‘beds’ 39, 40) exhibit small δ13Corg and TOCWR peaks.

8. Elemental chemostratigraphy of Dove’s Nest

Major-element and selected trace-element profiles for the upper Pliensbachian – middle Toarcian of the Dove’s Nest core are plotted in Supplementary Material Figures S1 and S2 and selected data ratioed to Al or Th are plotted in Figures 4 and 5. These profiles illustrate a clear lithogeochemical differentiation of the regional lithostratigraphic units identified in the core based on sedimentological criteria, carbon isotope and TOC profiles (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022). The elemental chemistry offers additional criteria for the identification of marker beds. It provides evidence of significant stratigraphic changes in sediment composition that do not necessarily coincide with lithostratigraphic unit boundaries but are considered to have palaeoenvironmental significance.

The chemostratigraphic units proposed by Remírez & Algeo (Reference Remírez and Algeo2020) are readily identified in the Dove’s Nest core. The succession is further divided here, with a total of 14 chemostratigraphic intervals recognizable on the δ13Corg and TOCWR profiles (Fig. 2) but characterized by a wide range of elemental properties (Figs 4, 5, S1, S2): Unit I, which was not sampled in detail for previous geochemical studies, is divided into three subunits: Ia corresponds to the top of the Staithes Sandstone, Ib to the Penny Nab Member of the Cleveland Ironstone and Ic to the Kettleness Member. The base of Unit II is taken immediately above the base of the Grey Shale of the Whitby Mudstone at the correlated level of the Sulphur Band. The unit comprises Subunits IIa and b; the base of IIb is placed at the step increase in TOCWR associated with the laminated black shale of SB3.

Unit III, the interval of the T-OAE, which corresponds to the top of the Grey Shale and the Jet Rock at the base of the Mulgrave Shale, is divided into four subunits based primarily on the δ13Corg curve but coincident with significant shifts in the elemental chemostratigraphy (Figs 4, 5). Subunit IIIa corresponds to the main falling limb of the negative δ13C excursion, ‘phase A’ of Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018). Subunit IIIb is the δ13C minimum (‘phase B’), and IIIc is the rising limb of the CIE (‘phase C’). Subunit IIId is the small δ13C peak immediately above the negative excursion interval, comprising the top Jet Rock (‘beds’ 39, 40; Top Jet Dogger and Millstones) in the coastal sections.

Unit IV constitutes the lower Bituminous Shales characterized by high TOC content and a low plateau in δ13C values. It is divided into a lower subunit (IVa) with falling TOC and an upper subunit (IVb) with no long-term TOC trend. Unit V is divided into three subunits: Va corresponds to the upper Bituminous Shales, Vb to the Hard Shales at the base of the Alum Shale Member, an interval of carbonate-cemented shales with elevated TOC contents (Figs 2, 4) and Vc to the ‘Main Alum Shale’ beds.

The litho- and biostratigraphic framework, sedimentology and geochemical characteristics of the chemostratigraphic units are described below.

8.a. Unit I Staithes Sandstone and Cleveland Ironstone

The lowest beds sampled during this study are assigned to the uppermost Staithes Sandstone Formation (A. stokesi Subzone, Fig. 2; Powell, Reference Powell1984, Reference Powell2010; Howard, Reference Howard1985; Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995; Hesselbo & King, Reference Hesselbo, King and Lord2019). In the coastal sections, the formation consists of bioturbated fossiliferous, fine- to medium-grained, micaceous, calcareous sandstones and sandy siltstones. At outcrop, colours range from blue-grey when fresh to yellow-brown when weathered. Thicker sandstones are commonly cross-bedded, including hummocky cross-stratification (HCS). Thinner sandstone sheets have erosional bases and fine-upward to mudstone. Parallel lamination, low-angle cross-lamination and wave-ripple lamination are common in the sandstones where not destroyed by bioturbation. Sideritic and calcitic concretions are present in the more argillaceous beds. The sandstones were deposited from sediment-laden, storm-surge and ebb currents in inner- and mid-shelf settings (Powell, Reference Powell2010). The formation is 28.6 m thick in the type section at Staithes and 25.7 m at Hawsker Bottoms (Howard, Reference Howard1985).

The Staithes Sandstone is interpreted to have accumulated predominantly in a shoreface environment above storm wave-base, likely 25 – 30 m depth in an epeiric setting (Walker & Plint, Reference Walker, Plint, Walker and James1992; Immenhauser, Reference Immenhauser2009), followed by long-term deepening through the Cleveland Ironstone, which was deposited mostly below storm wave-base (Van Buchem & Knox, Reference Van Buchem, Knox, Graciansky, Hardenbol, Jacquin and Vail1998). Following Howard (Reference Howard1985), the base of the Cleveland Ironstone Formation at Hawsker Bottoms is placed at the base of ‘bed’ 17 of Howarth (Reference Howarth1955), where hard ferruginous and calcareous sandstone (‘bed’ 16) passes up into mudstone (‘bed’ 17). This lies towards the summit of the A. stokesi Subzone (A. margaritatus Zone) of the lowest upper Pliensbachian (Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995).

Blue-grey coarse mudstones and very fine sandstones predominate from the base of the Dove’s Nest core study interval at 223.64 m. Cross-lamination and lenticular bedding are common. A lithological change to very thin – thin beds of grey medium- to coarse-grained mudstone at 209.7 m is correlated to the base of ‘bed’ 17 at Hawsker Bottoms (Fig. 2), marking the base of the Cleveland Ironstone Formation (Howard, Reference Howard1985). The Staithes Sandstone in the core contains more mud than at the coast and the sandstones are finer and thinner; the medium to thick HCS beds of the coastal exposures are absent. By comparison to the coast, the 13.5 m thick section studied at Dove’s Nest represents the upper half of the formation.

8.a.1. Subunit 1a (Staithes Sandstone) 219.1 – 209.7 m

Few geochemical data are available for the Staithes Sandstone from the coast: Gad (Reference Gad1966) analysed one sample from Staithes (’bed’ 12 of Howarth, Reference Howarth1955) and McArthur (Reference McArthur2019) and Remírez & Algeo (Reference Remírez and Algeo2020) reported a partial analysis of another sample from the same locality (‘bed’ 17). The formation lies within chemostratigraphic Unit I of the latter authors (Fig. 2). We distinguish the Staithes Sandstone as Subunit 1a (base not defined), which is characterized by the lowest and most erratic Al2O3, TiO2, K2O, Cs, Rb, Sc and Th values (typical phyllosilicate-associated elements) and low TOC and U contents (Figs S1, S2). Coarse detrital mineral proxy elements SiO2 (quartz), Na2O (Na-plagioclase) and Zr (zircon) are notably enriched (Fig. S1), up to 70%, 1.8% and 320 ppm, respectively.

Very high but stratigraphically upward decreasing values of Si/Al, Zr/Al and Na/Al with rising K/Al, Cs/Al and Rb/Al (e.g. Fig. 4) are interpreted to reflect a broad fining-upward trend towards a sharp facies break at the formation top. Beds containing carbonate concretions are revealed by isolated spikes in CaO and MnO, the former representing up to 17% CaCO3e (Fig. 4). A weakly mineralized calcareous ‘ironstone’ (14% Fe2O3T, 5.7% CaO) was sampled at 218.02 m. Elevated MnO (0.11%) and a lack of Mg enrichment point to a likely oxide – hydroxide (e.g. goethite) mineralogy in this bed. The sampling resolution was insufficient to resolve clear bed-scale trends in this part of the succession.

8.a.2. Subunit 1b (Penny Nab Member, Cleveland Ironstone) 209.7 – 196.3 m

In the type area of the Cleveland Hills, the Cleveland Ironstone Formation contains thick ironstone beds (‘seams’) that were mined extensively in the nineteenth and early twentieth centuries (Hallimond et al., Reference Hallimond, Ennos and Sutcliffe1925; Whitehead et al., Reference Whitehead, Anderson, Wilson and Wray1952; Chowns, Reference Chowns1968). The ironstones thin and the intervening mudstones thicken to the SE, where the formation consists dominantly of silty mudstones with interbedded thin seams of sideritic and berthierine-rich (‘chamosite’) oolitic ironstone (Howard, Reference Howard1985; Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995). The latter constitute up to 33% of the stratal thickness of the formation in north Cleveland, decreasing to only 5% at Hawsker Bottoms (Chowns, Reference Chowns1966, Reference Chowns1968; Howard, Reference Howard1985), immediately to the NE of Dove’s Nest (Fig. 1).

Regional ironstone marker beds (Fig. 2) include the Osmotherley (‘bed’ 18), Avicula (‘bed’ 20) and Raisdale (‘bed’ 23) seams. A fourth marker bed, the Two Foot Seam, is absent at Hawsker Bottoms due to the disconformity at the base of the Pecten Seam (‘beds’ 25 – 27). The seams cap coarsening-upward cycles typically 2 – 5 m thick, with individual cycles being laterally continuous over much of the basin (Shalaby, Reference Shalaby1980; Rawson et al., Reference Rawson, Greensmith and Shalaby1983; Howard, Reference Howard1985). The ironstones have been interpreted to represent marine flooding surfaces (Young et al., Reference Young, Aggett, Howard and Young1990) with each cycle constituting an individual depositional sequence, incorporating 0.1 – 1.0 m thick parasequences (Macquaker & Taylor, Reference Macquaker and Taylor1996).

Howard (Reference Howard1985) defined two members in the Cleveland Ironstone, the Penny Nab (top A. margaritatus Zone) and Kettleness (P. spinatum Zone) separated by the base Pecten Seam disconformity. The base of the Penny Nab Member is formed by a transgressive lag of reworked and bored carbonate nodules, internal moulds of the ammonite Amaltheus stokesi (Sowerby) and eroded and bored belemnites. The Penny Nab Member consists dominantly of medium- to coarse-grained mudstone and argillaceous silty sandstone with interbedded thin seams of oolitic ironstone. The member is built up of five small-scale cycles. The lowest three coarsen upward and are capped by ironstones, while the upper two are erosionally truncated below oolitic ironstone seams (Howard, Reference Howard1985; Young et al., Reference Young, Aggett, Howard and Young1990).

The Kettleness Member represents the ironstone facies that dominate the upper Cleveland Ironstone in its type area: in north and west Cleveland the member consists entirely of oolitic ironstone. The base of the Kettleness Member (P. spinatum Zone) rests with an angular unconformity on the Penny Nab Member across the whole of the Cleveland Basin (Chowns, Reference Chowns1966; Howard, Reference Howard1985). The top of the Penny Nab Member has been progressively removed by increasing erosion towards the south-east of the basin, with a whole parasequence missing in the Hawsker Bottoms section. To the east and southeast, the ironstone interfingers with and undergoes a lateral facies change into coarse-grained mudstone and subordinate very fine-grained sandstone. Minor coarsening-upward cycles, 0.35 – 2.5 m thick, are capped by nodular siderite mudstone.

The number of cycles within the member decreases to the NW from a maximum of six at Hawsker Bottoms. Sedimentary facies in the two members have been interpreted to reflect the progradation of sediments from the NW during deposition of the Penny Nab followed by a switch to progradation from the SE for the Kettleness (Young et al., Reference Young, Aggett, Howard and Young1990).

Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) placed the base of the Cleveland Ironstone at 210.15 m in the Dove’s Nest core. Based on our geochemical data, we place it slightly higher at 209.7 m corresponding to a level (sample 209.67) marking a sharp fall in SiO2 and rise in Al2O3 (Fig. S1) indicative of a significant increase in clay mineral content and representing a lithological transition to mudstone. The immediately underlying sample at 209.81 m shows very high Si/Al, Ti/Al and Na/Al ratios (Fig. 4) with a spike of 2300 ppm Ba and elevated MnO (Fig. S1) suggesting the presence of a former redox boundary. The Cleveland Ironstone Formation is 24.2 m thick in the core, very similar to the 23.4 m thickness recorded at Hawsker Bottoms by Howard (Reference Howard1985). Very thin to thin beds of grey medium- to coarse-grained mudstones dominate this interval. Beds have sharp erosional bases overlain by siltstone lags and are commonly bioturbated.

Five ironstone horizons were identified visually in the lower 13.4 m of the Cleveland Ironstone at Dove’s Nest by Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) which was ascribed to the Penny Nab Member. The thickest and most prominent ironstone was sampled at 200.39 m and coincides with a minimum and subsequent step increase in TOCWR values (Fig. 2) and a marked breakpoint in the Si/Al, Ti/Al, Zr/Al and Na/Al profiles (Fig. 4). This ironstone is particularly fossiliferous with a prominent oolitic texture and has been correlated to the Avicula Seam at the top of the Amaltheus subnodosus Subzone (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022).

The high 47.2% Fe2O3 but low SiO2 (4.9%) and Al2O3 (2.9%) contents of the Avicula Seam in the core point to dominant siderite mineralogy, with no significant pyrite observed, while coincident high MgO, MnO and CaO contents (Fig. S1) are indicative of associated magnesite and rhodochrosite substitution typical of Cleveland Ironstone siderites (Maynard, Reference Maynard1986; Aggett, Reference Aggett1990). Recalculation of the Dove’s Nest Avicula Seam analysis yields a total carbonate content of 89% with a composition of 76.5% FeCO3, 13.8% MgCO3, 8.6% CaCO3 and 1.1% MnCO3. This corresponds closely with electron microprobe analyses of Cleveland Ironstone siderites obtained by Aggett (Reference Aggett1990): 77.8 – 79.7% FeCO3, 10.8 – 13.9% MgCO3, 7.0 – 8.9% CaCO3 and 0.8 – 1.3% MnCO3. Based on a revision of the carbonate content (previously reported as 7.72% CaCO3 based solely on the Ca value; Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) to accommodate siderite rather than calcite, the TOCWR value calculated for this sample is 0.37%.

The siderite-dominated mineralogy of the Avicula Seam at Dove’s Nest is consistent with its equivalents in the coastal successions described by Chowns (Reference Chowns1968) and Mücke & Farshad (Reference Mücke and Farshad2005), although Gad (Reference Gad1966) emphasized the relatively berthierine-rich (chamosite) composition of some intervals within the ‘bed’. However, significant mineralogical and geochemical variation occurs within the 60 cm thick seam at Staithes (Gad, Reference Gad1966; Catt et al., Reference Catt, Gad, Le Riche and Lord1971) and siderite replacement of berthierine ooids and groundmass is commonly extensive in the Cleveland Ironstone Formation (Hallimond et al., Reference Hallimond, Ennos and Sutcliffe1925; Kearsley, Reference Kearsley, Young and Taylor1989; Aggett, Reference Aggett1990; Mücke & Farshad, Reference Mücke and Farshad2005).

Elevated P and P/Al (Figs 5, S1) point to the presence of associated phosphates, as observed in the Avicula Seam elsewhere (Young et al., Reference Young, Aggett, Howard and Young1990). The high V/Al ratio (Fig. 5) is consistent with the general V enrichment in the ironstones (Gad, Reference Gad1966; Mücke & Farshad, Reference Mücke and Farshad2005). A high Th/K ratio (Table S1) compared to the associated mudstones is also worth noting, as generally observed in Cleveland Basin ironstones (Parkinson, Reference Parkinson1996), which has been attributed to the presence of Th-rich kaolinite derived from a tropically weathered source area (Myers, Reference Myers, Young and Gordon Taylor1989). Despite some other samples from the Penny Nab showing limited Fe enrichment (>10% Fe2O3; Fig. S1), none of these display significant deviations in other constituents that characterize the Avicula Seam ironstone (Figs S1, S2).

The Avicula Seam marks the top of a 5 m thick interval of increasing upward Si/Al, Ti/Al, Zr/Al and Na/Al values that fall sharply at, or immediately above, the seam (Fig. 4). This pattern is interpreted to represent a major coarsening-upward cycle (cf. Rawson et al., Reference Rawson, Greensmith and Shalaby1983; Howard, Reference Howard1985), reflecting increasing proportions of coarser silt-size detrital minerals (e.g. quartz, Na-plagioclase, ilmenite, titanite, zircon) in the mudstones (cf. Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018). Comparable cycles occur below, at the base of the Penny Nab Member and above with values rising upward towards the top of the member at the disconformity at the base of the Pecten Seam (Fig. 4). A Si/Al peak immediately below the Pecten Seam has been documented previously at Staithes (Young et al., Reference Young, Aggett, Howard and Young1990, fig. 4). These three stacked coarsening-upward cycles generate an asymmetric saw-tooth pattern in the geochemical profiles (Fig. 4).

The coarsening-upward cycles developed in the Penny Nab Member at Dove’s Nest correspond broadly to Type 1 cycles I – III of Howard (Reference Howard1985) and ‘parasequences’ I – III of Young et al. (Reference Young, Aggett, Howard and Young1990), considered to be of sequence scale by Macquaker & Taylor (Reference Macquaker and Taylor1996). They represent the 4th-order cycles of Van Buchem & Knox (Reference Van Buchem, Knox, Graciansky, Hardenbol, Jacquin and Vail1998). Howard (Reference Howard1985) placed ironstones (Osmotherley, Avicula, Raiswell) at the top of each coarsening-upward cycle, while Young et al. (Reference Young, Aggett, Howard and Young1990) interpreted the ironstones as occurring above a disconformity (marine flooding surface) at the base of each ‘parasequence’. Macquaker & Taylor (Reference Macquaker and Taylor1996) considered the ironstones to have formed during prolonged breaks in sediment accumulation, either as the downdip equivalents of sequence boundaries or during the deposition of forced regressive systems tracts. The finer structure within the geochemical profiles at Dove’s Nest is consistent with the presence of additional smaller scale and lower amplitude cycles (parasequences), but these are inadequately resolved by the sampling resolution.

The top of the Penny Nab Member in the Dove’s Nest core comprises a succession of 30 thin beds with erosional bedding planes and silt lags. Comparable structures, some of which are gutter casts, form a distinctive 2 m interval in the Lower Gibbosus Shales (Hawsker Bottoms top ‘bed’ 21 and ‘bed’ 22 of Howarth, Reference Howarth1955) called the ‘upper striped bed’ by Greensmith et al. (Reference Greensmith, Rawson and Shalaby1980). This interval was considered by Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) to represent the top of the member at Dove’s Nest since the Raisdale Seam, an ironstone regionally developed between the Lower Gibbosus Shales and the Middle Gibbosus Shales (Cope et al., Reference Cope, Getty, Howarth, Morton and Torrens1980), is either indistinct or absent in the core.

8.a.3. Subunit 1c (Kettleness Member, Cleveland Ironstone) 196.3 – 185.6 m

Across most of the Cleveland Basin, the base of the Kettleness Member, marking the contact between the A. margaritatus and P. spinatum zones (A. gibbosus and P. apyrenum subzones), is a disconformity that becomes more pronounced toward the south and east (Cope et al., Reference Cope, Getty, Howarth, Morton and Torrens1980; Howard, Reference Howard1985), but is lithologically indistinguishable from a normal bedding plane. In the Dove’s Nest core, as in the coastal outcrops, the bottom of the Kettleness Member consists of coarse mudstones with lenticular bedding formed by starved combined-flow ripples of siltstone encased in finer mudstones (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022). These mudstones are part of a group of beds referred to as the Pecten Seam (Marley, Reference Marley1857; Chowns, Reference Chowns1966; Cope et al., Reference Cope, Getty, Howarth, Morton and Torrens1980; Simms et al., Reference Simms, Chidlaw, Page, Morton and Gallois2004b). The Pecten Seam has been defined differently by different authors. Howarth (Reference Howarth1955) and Cope et al. (Reference Cope, Getty, Howarth, Morton and Torrens1980) included Hawsker Bottoms ‘beds’ 25 – 31 in the Pecten Seam comprising a lower 43 cm thick Grosmont Pecten Bed (‘beds’ 25 – 27) at the base, overlain by the 2.6 m thick Eston Shell Beds (‘beds’ 28 – 31), above. Chowns (Reference Chowns1966) additionally included ‘bed’ 32, the 1.1 m thick ‘Black Hard’ in the Seam. By contrast, Howard (Reference Howard1985), Young et al. (Reference Young, Aggett, Howard and Young1990) and Hesselbo & Jenkyns (Reference Hesselbo, Jenkyns and Taylor1995) restricted the Pecten Seam to ‘beds’ 25 – 27.

The Pecten Seam at Dove’s Nest was placed between 196.36 – 193.29 m at Dove’s Nest by Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022), corresponding to Hawsker Bottoms ’beds’ 25 – 31 (cf. Howarth, Reference Howarth1955). The bottom ironstone at 196.30 m (<10 cm thick) is oolitic, fossiliferous and bioturbated, whereas the remaining ironstone horizons, picked out by a reddish hue at 195.18 m (25 cm) and 193.69 m (30 cm) in an otherwise black succession, are indistinct. The mudstones between the ironstone horizons vary between medium-grained and homogenous, with pyrite-filled small burrows, to coarser, thin-bedded and (low-angle) cross-laminated. The top of the Kettleness Member (∼187 m) is characterized by Chondrites burrows above laminated mudstones. The Sulphur Band (SB1) is not obvious in the core, probably due to the lack of weathering, but the distinctive bioturbated layer (top ‘bed’ 43) above the Sulphur Band seen at Kettleness and Runswick was identified by Trabucho-Alexandre et al. (Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022).

The Kettleness Member and the bottom of ‘bed’ 43 constitute chemostratigraphic Subunit Ic (top ∼185.6 m). Iron profiles demonstrate that the basal and second ironstones of the Pecten Seam were not captured by our geochemical sampling (Figs 5, S1) but sample 194.02 shows the top ironstone (25.2% Fe2O3T) to be again dominantly sideritic (50% carbonate) with an estimated composition of 72.9% FeCO3, 14.3% MgCO3, 11.8% CaCO3, 0.96% MnCO3. A trend of stratigraphically upward decreasing, then rising, TOCWR and δ13Corg values through the member at Dove’s Nest (Fig. 2) is accompanied by rising detrital proxy element ratios, Si/Al, Ti/Al, Zr/Al and Na/Al (Fig. 4) that peak around the TOC minimum, decrease above and then rise upward again forming peaks at the base of the Grey Shale. These are interpreted to represent two major stacked coarsening-upward cycles above the base Kettleness disconformity (Fig. 4). The P2O5 peaks within the Pecten Seam (Figs 5, S1) at Dove’s Nest may correspond to the francolite (carbonate fluorapatite) concentrations observed in the seam at Staithes by Young et al. (Reference Young, Aggett, Howard and Young1990, fig. 4).

Howard (Reference Howard1985) distinguished seven (Pecten Seam and i – vi) Type 2 cycles (= ‘parasequences’ of Young et al., Reference Young, Aggett, Howard and Young1990) in the Kettleness Member at Hawsker Bottoms. As in the underlying Penny Nab Member, finer structure within the profiles at Dove’s Nest suggests that additional smaller scale cycles are present in addition to the two that are documented, but the sampling resolution is inadequate to capture the detail of these. The coarsest-grained interval in the middle of the Kettleness Member at Dove’s Nest likely corresponds to cycle/‘parasequence’ iv on the coast.

In total, five major coarsening-upward cycles through the Cleveland Ironstone are resolved by our data (Fig. 4). Cycle 3 is bounded by the Avicula Seam at the base and Pecten Seam at the top, cycle 5 terminates in the Pliensbachian – Toarcian boundary interval below the Sulphur Band.

8.b. Unit II Whitby Mudstone lower – middle Grey Shale

The Grey Shale (Powell, Reference Powell1984) is the lowest member of the Whitby Mudstone Formation in Yorkshire (Fig. 2), which is the type area for the D. tenuicostatum Zone, the lowest zone of the Toarcian (Howarth, Reference Howarth1973; Hesselbo et al., Reference Hesselbo, Ogg, Ruhl, Hinnov, Huang, Gradstein, Ogg, Schmitz and Ogg2020b). Hawsker Bottoms ‘bed’ 43 contains the Pliensbachian – Toarcian boundary at the base of the Sulphur Band (SB1), 33 cm above the base of the Grey Shale (= base of Kettleness ‘bed’ 25). The member consists of 13.6 m of grey mudstone containing 15 rows of calcareous concretions. The lower beds include micaceous mudstones and ripple-laminated coarse-grained mudstones indicative of current activity, but these pass up into laminated mudstones in the upper part of the member. Evidence for storm-influenced deposition includes normally graded layers with silt lags interpreted as tempestites, as well as ripples, gutter casts and HCS (Kemp et al., Reference Kemp, Fraser and Izumi2018), indicative of deposition above storm wave-base and likely < 30 m water depth (Section 8a; cf. Walker & Plint, Reference Walker, Plint, Walker and James1992; Immenhauser, Reference Immenhauser2009).

The mudstones of the Grey Shale in the Dove’s Nest core are similar to those of the Cleveland Ironstone Formation, but they are a much darker grey and are the least fossiliferous interval of the studied section. By contrast, the Cleveland Ironstone and the Bituminous Shales are the two most fossiliferous intervals albeit with very different assemblages.

The intervals containing the Sulphur Band marker bed (SB1) and overlying Sulphur Band 2 (SB2), with the twin δ13Corg minima of the Pliensbachian – Toarcian Boundary Event, are clearly picked out by TOC, Fe2O3, Fe/Al (Figs 4, 5, S1) and S (not determined here, but well displayed in the coastal succession (e.g. Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018, fig. 8) peaks.

Figure 8. Box plot summary of stratigraphic trends in major- and trace-element contents and Al-ratio data of chemostratigraphic units through the upper Pliensbachian – middle Toarcian of the Dove’s Nest core. Boxes represent 25–75% quartiles with median values shown by the vertical line inside the box. Whiskers are drawn from the top of the box up to the largest data point less than 1.5 times the box height from the box (the ‘upper inner fence’) and similarly below the box (Hammer et al., Reference Hammer, Harper and Ryan2001). Outlier values outside the inner fences are shown as circles, values further than 3 times the box height from the box (the ‘outer fences’) are shown as stars.

8.b.1. Subunit IIa (lower Grey Shale Member) 185.6 – 180.4 m

The base of the Whitby Mudstone, the base of Subunit IIa, marks a major long-term shift in the chemostratigraphy of the succession with coincident falls in coarser grained detrital proxies Si/Al, Ti/Al, Na/Al, Zr/Al and accompanying rises in phyllosilicate (principally clay minerals) associated element ratios K/Al, Rb/Al (Fig. 4) and Cs/Al (not plotted). The succession between Sulphur Bands 2 and 3 (SB2 – SB3, P. paltum Subzone) includes a distinctive 1.8 m succession of six rows of deep-red weathering sideritic nodules and interbedded decimetric grey mudstones (Port Mulgrave ‘beds’ 7 – 17), the Six Red Nodule beds of Howarth (Reference Howarth1973) and Hesselbo & Jenkyns (Reference Hesselbo, Jenkyns and Taylor1995).

8.b.2. Subunit IIb (middle Grey Shale Member) 180.4 – 174.5 m

A marked step increase in the TOC profile above the red nodule beds in the basal D. clevelandicum Subzone (Fig. 2) accompanies the laminated bituminous mudstones of Sulphur Band 3 (SB3, ‘bed’ 19a) at the base of Subunit IIb. Coincident peaks in Fe2O3T, MgO and MnO (Fig. S1) point to a siderite component in this bed, in addition to pyrite observed in the core. Accompanying the rise in TOC, are step increases in Al2O3, MgO, K2O and P2O5 and step falls in SiO2, Na2O, Si/Al, Ti/Al, Zr/Al and Na/Al (Figs 4, S1) indicating a consistently higher proportion of phyllosilicates and relatively stable, finer grain sizes throughout Subunit IIb.

8.c. Unit III (T-OAE) Whitby Mudstone, Grey Shale – Mulgrave Shale

Excellent agreement exists between the high-resolution chemostratigraphic records published for the interval of the T-OAE on the Yorkshire coast (Figs 6, 7; Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005, Reference Kemp, Coe, Cohen and Weedon2011; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; Wang, Reference Wang2022) and the new data from Dove’s Nest (Fig. 7).

8.c.1. Subunit IIIa (upper Grey Shale Member) 174.5 – 171.2 m

The onset of rising upward TOC values and falling δ13Corg values occur from the base of Port Mulgrave ‘bed’ 31 in the upper Grey Shale D. semicelatum Subzone, with increasing rates of change starting at the transition to darker-coloured laminated sediments in the middle of ‘bed’ 31 (Fig. 7). The colour change occurs at the top of an interval containing increasing siltstone laminae, reflected by increasing Si/Al ratios (Fig. 7), with HCS recorded at Kettleness (cf. Wignall et al., Reference Wignall, Newton and Little2005, fig. 3). Benthic faunal diversity crashes and bioturbation disappears progressively through the lowest 1.5 m of ‘bed’ 31 (Wignall et al., Reference Wignall, Newton and Little2005; Caswell et al., Reference Caswell, Coe and Cohen2009; Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023). Pyrite framboid median and maximum sizes decrease above the onset of the TOC increase and mudstone lamination (Wignall et al., Reference Wignall, Newton and Little2005). Framboid maximum sizes fall to <5 μm and then remain constant upwards, indicative of sulfide precipitation in a euxinic water column (Wignall & Newton, Reference Wignall and Newton1998).

The base of Subunit IIIa is placed at the onset of the TOC rise which at Dove’s Nest is coincident with the start of generally falling upward trends in detrital and broader aluminosilicate proxies, SiO2, TiO2, Al2O3, Na2O, Cs, Rb, Sc, Th, Zr, Si/Al, Ti/Al, Zr/Al and Na/Al (Figs 4, 5, 7, S1, S2) but increases in parameters associated with reducing organic-rich sediments Fe/Al (pyrite), P/Al (phosphate), U/Th and Mo/Al (organic matter). Subunit IIIa extends upward to the δ13Corg minimum (‘bed’ 33) at the base of the Jet Rock (Figs 6, 7).

The top of the Grey Shale (‘beds’ 31, 32), in addition to becoming increasingly organic-rich upwards, is characterized by the presence of widespread shell beds (Howarth, Reference Howarth1973) and represents an interval of significant biotic turnover (Caswell et al., Reference Caswell, Coe and Cohen2009; Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023) with a progressive fall in benthic macroinvertebrate diversity upwards, towards a minimum in the lower Jet Rock (Danise et al., Reference Danise, Twitchett and Little2015).

8.c.2. Subunit IIIb (lower Jet Rock) 171.2 – 167.2 m

The Mulgrave Shale Member (Rawson & Wright, Reference Rawson and Wright1992) overlies the Grey Shale with the boundary on the coast being placed below a line of calcareous concretions, the Cannon Ball Doggers, ‘bed’ 33, at the base of the C. exaratum Subzone (Figs 6, 7). The Mulgrave Shale comprises the Jet Rock and Bituminous Shales beds of the H. serpentinum Zone (Fig. 2). The Jet Rock (Figs 6, 7) extends from ‘beds’ 33 – 40 (Millstones) inclusive (Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995) and is named after the occurrence of jet (compressed wood, traditionally ascribed to araucarian species but likely including a wide range of taxa; Caldwell Steele, Reference Caldwell Steele2020) that was previous mined from these beds (Hemingway, Reference Hemingway1933). The interval is coincident with the C. exaratum Subzone (Howarth, Reference Howarth1992; Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995; Simms et al., Reference Simms, Chidlaw, Morton and Page2004a), although Howarth (Reference Howarth1962) originally included ‘bed’ 32 (top D. semicelatum Zone) in the Jet Rock.

The Jet Rock, which consists of dark grey finely laminated pyritic bituminous mudstones, is 7.09 m thick on the coast at Port Mulgrave (Figs 2, 6, 7; Howarth, Reference Howarth1962). Laminae contain coccolith-rich pellets and organo-mineral aggregates and include normally graded layers with sharp bases that show erosional truncation. Discontinuous wavy laminae exhibit downlapping geometries. These features are consistent with deposition by advective dispersal of sediment from wave-enhanced gravity flows of fluid mud triggered by storms (Ghadeer & Macquaker, Reference Ghadeer and Macquaker2011).

A characteristic feature of the Jet Rock is the presence of common cm- to m-sized calcareous concretions, some with pyritic skins (Coleman & Raiswell, Reference Coleman and Raiswell1981). Key marker horizons within the Jet Rock (Figs 6, 7) are the: (1) Cannon Ball Doggers (‘bed’ 33), scattered 10 – 15 cm diameter spherical calcareous nodules with thick pyritic skins; (2) Whale Stones (‘bed’ 35), up to 45 cm thick, 1 × 3 m elongate doggers with thin pyritic skins, associated with irregular linear arrays of smaller pyritic-skinned tubular nodules (Lower Pseudovertebrae); and (3) Curling Stones (‘bed’ 37), 20 × 40 cm oblate spheroidal doggers with pyritic skins. The top of the Jet Rock comprises a prominent 50 cm thick interval of strongly cemented argillaceous limestone: ‘bed’ 39, the Top Jet Dogger, incorporating a 20 – 30 cm layer with very large (up to 4 m diameter) almost perfectly circular discoidal concretions set in its top, ‘bed’ 40, the Millstones (Howarth, Reference Howarth1962).

Significantly, although clearly of diagenetic origin, the most prominent named concretion beds in the Cleveland Basin succession exactly match the stratigraphic positions of equivalent carbonate marker horizons in the Posidonia Shale of Germany with respect to the corresponding carbon isotope curves: Whale Stones = Unterer Stein; Curling Stones = Steinplatte; Top Jet Dogger–Millstones = Oberer Stein, for example at Dotternhausen Quarry in the SW German Basin (e.g. see Röhl et al., Reference Röhl, Schmid-Röhl, Oschmann, Frimmel and Schwark2001; Ruebsam et al., Reference Ruebsam, Schmid-Röhl and Al-Husseini2023), 950 km to the SE (Fig. 1). Comparable concretionary horizons are seen in the Schistes carton of the Truc de Balduc area of southern France (Riegraf, Reference Riegraf, Einsele and Seilacher1982) and a carbonate maximum at the top of the T-OAE interval, at a level equivalent to the Top Jet Dogger–Millstones, is developed in the Schistes carton at Sancerre in the southern Paris Basin (Hermoso et al., Reference Hermoso, Minoletti, Rickaby, Hesselbo, Baudin and Jenkyns2012, Reference Hermoso, Minoletti and Pellenard2013). Other less prominent carbonate beds are also seen in consistent positions in Germany, France and elsewhere, demonstrating the wider regional significance of the Yorkshire lithostratigraphy.

The Jet Rock contains the highest TOCWR contents in the Pliensbachian – Toarcian boundary succession (Figs 2, 6, 7) with values typically around 6 – 10% but rising to >18% in one layer on the coast (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Kemp et al., Reference Kemp, Coe, Cohen and Weedon2011; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; Houben et al., Reference Houben, Goldberg and Slomp2021). High-resolution (2.5 cm spaced) CaCO3, TOC, S and δ13Corg profiles for the Jet Rock and immediately under- and overlying strata (DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005; Kemp et al., Reference Kemp, Coe, Cohen and Weedon2011) on the coast provide a basis for the definition and precise placement of chemostratigraphic subunits, which can be correlated, refined and further interpreted using the multi-element dataset from the Dove’s Nest core.

The base of the Jet Rock (‘bed’ 33) and of the H. serpentinum Zone mark a significant chemostratigraphic boundary with a sharp fall and minimum of δ13Corg that we use to define the base of Subunit IIIc (Figs 6, 7). By contrast, Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) placed the base of their ‘phase B’ ∼65 cm lower at an underlying sharp δ13Corg decrease in the upper half of ‘bed’ 32, below. Subunit IIIb represents the δ13C minimum of the T-OAE, within which a change from falling to rising Si/Al and Ti/Al in the lower half of ‘bed’ 34 accompanies a further step increase in TOC values (Fig. 7). These all reach a maximum in mid-‘bed’ 35, the Whale Stones, and remain high up into the base of ‘bed’ 36. These high TOC values are associated with an interval temporarily but overwhelmingly dominated by dense amorphous marine organic matter (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019; Houben et al., Reference Houben, Goldberg and Slomp2021).

Subunit IIIb shows prominent negative excursions in many elements towards the top due to the diluting effects of the high carbonate content of the Whale Stones interval (Figs S1, S2). By contrast, Al- and Th-normalized values (Figs 4, 5) generally peak towards the top of the subunit, linked to increasing terrestrial detritus and grain size (Si, Ti, Na, Zr), changes in clay mineral composition (K, Rb) and phosphate deposition (P). High pyrite contents evidence increased sulfate reduction driven by higher organic matter contents and increasingly euxinic conditions (indicated by V, U, Mo enrichment in the shales).

8.c.3. Subunits IIIc and IIId (upper Jet Rock) 167.2 – 164.1 m

The sharp drop in TOC towards the base of ‘bed’ 36 coincides with the onset of rising δ13Corg towards a maximum at the top of the Jet Rock (Figs 6, 7). The interval of δ13Corg rise and lower TOC plateau is differentiated as Subunit IIIc. A δ13Corg maximum and TOCIR peak occur in the limestones and calcareous shales of the Top Jet Dogger and Millstones (‘beds’ 39 – 40), above, which are designated as Subunit IIId. It is observed that CaCO3 contents increase steadily upward from the middle of ‘bed’ 38, forming a broad peak that reaches a maximum in the Top Jet Dogger but continues upward into the lower part of ‘bed’ 41 at the base of the Bituminous Shales and the H. falciferum Subzone.

High MgO but low Fe2O3 values (Fig. S1) point to the likely presence of dolomite in addition to siderite in the Whale Stones and Top Jet Dogger – Millstones intervals, which also show Fe/Al and Mn/Al enrichment indicative of the latter mineral (Fig. 5), while high P2O5 and P/Al peaks evidence the presence of phosphates and high Ba and Ba/Al peaks, concentrations of barite. Erratic high MgO and Mg/Al ratios throughout the Jet Rock (Figs 8, S1) indicate the occurrence of dolomite in the black shales (cf. Pye, Reference Pye1985). However, cements in the Whale Stones concretions consist almost exclusively of calcite (Hallam, Reference Hallam1962). Erratic high P values here and elsewhere in the succession may be associated with levels of phosphatic fish debris which, on the coast, are particularly common in the Jet Rock and lower Bituminous Shales (Caswell & Coe, Reference Caswell and Coe2014) Unit III – IV interval, and are also a common feature in coeval black shale successions like the Posidonia Shale of Germany (Riegraf, Reference Riegraf1985; Burnaz et al., Reference Burnaz, Littke, Grohmann, Erbacher, Strauss and Amann2024) and the Schistes carton of southern France (Bomou et al., Reference Bomou, Suan, Schlögl, Grosjean, Suchéras-Marx, Adatte, Spangenberg, Fouché, Zacaï, Gibert, Brazier, Perrier, Vincent, Janneau, Martin, Reolid, Duarte, Mattioli and Ruebsam2021).

8.d. Unit IV Whitby Mudstone, middle Mulgrave Shale

The Bituminous Shales, ‘beds’ 41 – 48, which constitute the middle and upper portions of the Mulgrave Shale Member, overlie the Jet Rock and are 22.9 m thick at Saltwick Nab (Howarth, Reference Howarth1962). These mainly laminated grey carbonaceous mudstones, contain few carbonate concretions but more abundant pyrite and crushed pyritized fossils than the underlying Jet Rock. Belemnites and fossilized wood are common, the latter including logs up to 3.7 m long at the outcrop. In the Dove’s Nest core, the base of the Bituminous Shales comprises thin beds of coarser mudstone with fossiliferous bedding planes characterized by monospecific assemblages of Pseudomytiloides dubius (Sowerby). Cross-lamination reappears in these mudstones (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022), which include abrupt alternations of thin beds of fine and coarse mudstone (striped beds).

Three beds in the Bituminous Shales provide markers in the coastal sections (e.g. Rawson & Wright, Reference Rawson and Wright2018): ‘bed’ 42, a row of scattered oval doggers with pyritic skins and Pseudomytiloides; ‘bed’ 44, a row of scattered doggers with pyritic aggregates; ‘bed’ 46, a red-weathering sideritic mudstone (Fig. 2). A distinctive 36 cm interval comprising a double row of red-weathering pyritic doggers at the top of the H. falciferum Subzone (‘bed’ 48) yields a unique ammonite fauna that includes Ovaticeras ovatum (Young and Bird). Following Simms et al. (Reference Simms, Chidlaw, Page, Morton and Gallois2004b), we place this ‘bed’, the Ovatum Band, at the top of the Bituminous Shales, although Howarth (Reference Howarth1962) listed it separately and did not assign it to a specific member. The Bituminous Shales (including the Ovatum Band) lie entirely within the H. falciferum Subzone, H. serpentinum Zone (Hesselbo & Jenkyns, Reference Hesselbo, Jenkyns and Taylor1995).

Geochemical profiles (e.g. Figs 2, S1) demonstrate that a major facies change occurs at ‘bed’ 44. A sharp step fall in TOCWR occurs above this bed, from 3.4 ± 0.5% below to 2.4 ± 0.3% above (Fig. 2), with coincident marked falls documented for redox-sensitive trace elements in the coastal sections, including total S, Cd, Co, Mo and Mn (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; McArthur, Reference McArthur2019). The boundary between chemostratigraphic Units IV and V based on this major redox-proxy shift (following Remírez & Algeo, Reference Remírez and Algeo2020), is placed at the base of ‘bed’ 45. A more subtle compositional change occurs within the lower part of the Bituminous Shales around the level of ‘bed’ 42, which marks a change from falling TOCWR values below, to a plateau above (Fig. 2) and coincident breakpoints occur in most redox-sensitive trace-element profiles (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; McArthur, Reference McArthur2019). We subdivide these beds into chemostratigraphic Subunits IVa below and IVb above.

8.d.1. Subunit IVa (lower Bituminous Shales, Mulgrave Shale Member) 164.1 – 157.4 m

Subunit IVa is characterized by relative plateaus in most major-element profiles at Dove’s Nest (Fig. S1). The unique feature of the subunit is the very high levels of redox-sensitive trace metals, U, V and Mo, with the last of these peaking markedly (53 ppm Mo with 5 ppm U, 190 ppm V) towards the bottom of the subunit and remaining high above (Figs 5, S2). Strontium concentrations are also notably enriched in the lower section (Fig. S1) but display no clear correspondence to carbonate-associated elements, suggesting an association with phosphate (e.g. carbonate fluorapatite) and possibly diagenetic celestite.

8.d.2. Subunit IVb (middle Bituminous Shales, Mulgrave Shale Member) 157.4 – 151.2 m

The base of Subunit IVb is marked by small steps in all major element profiles (Fig. S1) driven largely by an increase in CaO contents accompanying a facies change to more sideritic shales. The subunit boundary coincides with a breakpoint in the long-term Ti/Al profile (Fig. 4) with a change from erratically falling values below to a generally increasing upward trend, attributed to the upward coarsening of the grain sizes towards the study section top. A significant broad Sr peak occurs at the base of the subunit (Fig. S1). Redox-sensitive trace metals U, V and Mo remain high throughout and then fall sharply with the TOC and CaCO3e declines above ‘bed’ 44 (Peak Stones) at the top of the subunit (Fig. S2). Subunit IVb is further characterized by relatively high P2O5 and P/Al values indicative of a higher phosphate mineral content than adjacent beds.

8.e. Unit V Whitby Mudstone, Mulgrave Shale – Alum Shale

The chemostratigraphic boundary between Units IV and V at ‘bed’ 44 (Peak Stones) within the Bituminous Shales previously observed on the Yorkshire coast (Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; McArthur, Reference McArthur2019; Remírez & Algeo, Reference Remírez and Algeo2020) is clearly developed in the Dove’s Nest core above 151.9 m, which is correlated to ‘bed’ 44.

8.e.1. Subunit Va (upper Bituminous Shales, Mulgrave Shale Member) 151.2 – 143.7 m

Here, a coincident rise in Al2O3 and marked falls in CaCO3e (Figs 2, S1), TOCWR and redox-sensitive trace-metal proxies, including Fe/Al, V/Al, U/Th and Mo/Al ratios (Fig. 5), is observed. Our new high-resolution carbon isotope profile from the core (Fig. 2; Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) confirms a step increase in δ13Corg at the same level, previously hinted at by low-resolution data from the coast (Cohen et al., Reference Cohen, Coe, Harding and Schwark2004).

The geochemistry of the mudstones forming Subunit Va is uniform, with relatively flat plateaus on most elemental and Al-ratio profiles (Figs 4, 5, S1, S2) although Ti/Al displays an increasing upward trend.

8.e.2. Subunit Vb (Hard Shales, Alum Shale Member) 143.7 – 137.2 m

‘Bed’ 49 marks the base of the Alum Shale Member of the Whitby Mudstone on the Yorkshire coast. The basal part of the Alum Shale was termed the Hard Shales by Howarth (Reference Howarth1962), reflecting the presence of carbonate cements in ‘beds’ 49 and 50. This 6.32 m thick interval displays a distinctive geochemical signature which, in addition to an elevated CaCO3e content compared to the immediately underlying and overlying succession, includes higher TOCWR and Mo contents with peak values at the bottom and top of the unit (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008, fig. 2; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018, fig. 2). ‘Bed’ 48, the Ovatum Band and ‘bed’ 50, a 20 cm thick red-weathering sideritic mudstone, show associated Mn enrichment (McArthur, Reference McArthur2019, fig. 2).

A correlative interval of weakly carbonate-cemented mudstones (0.6 – 4.7% CaCO3e) and increased TOCWR (2.1 – 3.9%) bracketed by peaks at the bottom and top is observed in the Dove’s Nest core between 143.8 and 137.6 m (Figs 2, S1). This interval is assigned to the Hard Shales and constitutes the base of the middle Toarcian H. bifrons Zone (D. commune Subzone; Fig. 2). It corresponds to chemostratigraphic Subunit Vb. The high-Mn signature of the Ovatum Band on the coast is not seen in the core, but this part of the bed may not have been sampled or the high Mn content in the coast sample of McArthur (Reference McArthur2019) may be a weathering artefact. The TOC peak at the base of Subunit Vb displays U and Mo enrichment and the interval in general is characterized by relatively elevated Mo and Mo/Al ratios (Figs 5, S2).

8.e.3. Subunit Vc (‘Main Alum Shales’, Alum Shale Member) 137.2 – 126.2 m

Howarth (Reference Howarth1962) referred to the 15.47 m of mudstone overlying the Hard Shales as the ‘Main Alum Shales’ (beds 51 – 64) that are in turn overlain by the carbonate cemented beds of the Cement Shales at the top of the Alum Shale. The ‘Main Alum Shales’ are very fossiliferous with abundant thick-shelled shallow-burrowing bivalves, Dacryomya ovum (Sowerby) and common belemnites and ammonites (e.g. Atkinson et al., Reference Atkinson, Little and Dunhill2023). ‘Beds’ 51 – 59 (12.24 m) lie within the D. commune Subzone (Fig. 2).

The top of our study interval in the Dove’s Nest core (137.2 – 126.2 m) corresponds to the D. commune Subzone of the ‘Main Alum Shales’ and is referred to chemostratigraphic Subunit Vc. The geochemistry of these beds is very comparable to Subunit Va at the summit of the Bituminous Shales, but with higher Ti/Al ratios indicative of a higher silt content. The top of Subunit Vc is not defined.

8.f. Elemental chemostratigraphy overview

The general characteristics of the chemostratigraphic units described above are summarized in Figure 8, showing log-scaled box plots of the elemental and Al-ratio data for all major and trace elements determined in the Dove’s Nest core. Elements showing an association with the coarser detrital fraction (e.g. Si, Ti, Zr) or phyllosilicates (e.g. K, Rb, Sc, Th) commonly show significantly different trends for the Al-ratio data, while those associated strongly with organic matter content and sediment redox (TOC, P, Mn, Mo, U, V) are less affected by normalization. The figure illustrates the geochemical integrity of the different units.

9. Statistical analysis

9.a. Pearson r correlation

Inter-element relationships were assessed using Pearson correlation of the high-resolution sample set, and the low-resolution samples with additional trace-element data. Very similar coefficients were obtained for elements present in both sample suites. Results for the low-resolution samples are provided in Supplementary Material Table S3. Correlations between major elements are attributed to their partitioning in the dominant minerals in the rocks. Mineralogical analyses were not performed as part of the present study, but the interpretation is informed by comparison to published mineralogical data from equivalent intervals on the Yorkshire coast (e.g. Gad et al., Reference Gad, Catt and Le Riche1969; Pye & Krinsley, Reference Pye and Krinsley1986). Quartz and clay mineral contents for 21 samples from the Dove’s Nest core were provided by de Vos (Reference de Vos2017).

High correlation coefficients (r >0.85) are observed between Al2O3 and other aluminosilicate-associated elements TiO2, K2O, Cs and Rb (Table S3). The highest correlation coefficients (r > 0.9) are between K2O, Rb and Cs, typically associated with phyllosilicates, especially illite (e.g. González López et al., Reference González López, Bauluz, Yuste, Mayayo and Fernández-Nieto2005) and mica. Lower correlation coefficients for Al2O3 (r = 0.5 – 0.8) with SiO2, Cr, Sc and Th indicate additional mineralogical controls (e.g. quartz, heavy minerals) on these elements. SiO2 shows the highest correlation with TiO2 (r = 0.8) and lower values (r = 0.5 – 0.8) for Al2O3 Na2O, Cr, Sc, Th and Zr, an element suite likely associated with quartz, Na-plagioclase and heavy minerals (e.g. rutile, ilmenite, titanite and zircon) in a coarser-grained terrigenous clastic fraction.

CaO, principally controlled by carbonate, predictably has a high negative correlation (r = < −0.8) with most of the above elements (Table S3) but shows modest positive correlations with MnO and P2O5 (r = 0.52) attributable to Mn substitution in calcite and the presence of carbonate fluorapatite. Fe2O3 is positively correlated only with MnO and MgO (r = ≥0.6) and V (r = 0.4) reflecting its disparate mineral associations that include siderite but also Fe–Mn oxyhydroxides, pyrite and berthierine.

9.b. PCA results

Loading coefficients of the different variables of the variance-covariance matrix for the first three principal components of the PCA for selected Dove’s Nest data (96 samples with a full element suite) are listed in Supplementary Material Table S4. The first two principal components together account for 71.3% of the variance in the data. PC1 (48.4% of the variance) shows positive loadings for siliciclastic mineral-associated (quartz, feldspars, phyllosilicates, heavy minerals) elements, SiO2, TiO2, Al2O3, Na2O, K2O, Cs, Cr, Rb, Th, Zr and a large negative loading for carbonate (CaO). This negative loading highlights the significance of carbonate impacting the geochemical composition of samples and that this is not solely an artefact of the closed-sum nature of the original dataset. Further discrimination between samples is offered by PC2 (22.9% of the variance) with large negative values for Mo* and TOCWR but positive loadings for MnO and Fe2O3, amongst others. PC3 accounts for a further 9.8% of the variance and is distinguished particularly by its high positive loading for Mo*.

A biplot of the first two principal components (Fig. 9) illustrates clear inter-element associations that can be interpreted as reflecting mineralogical controls on the bulk geochemistry of the rocks. It further demonstrates how these relate to the stratigraphic units comprising the upper Pliensbachian – middle Toarcian succession in the Cleveland Basin.

Figure 9. Principal component analysis biplot of PC1 vs PC2 for geochemical data from the upper Pliensbachian – middle Toarcian of the Dove’s Nest core. Compositional data for samples (n = 96) having a full major- and trace-element dataset were transformed using a Centre Log-Ratio to remove closure effects prior to PCA. The first two principal components account for 48.4% and 22.9% of the variance, respectively (Table S4).

The close proximity of vectors for individual constituents on the biplot indicates associations between (1) SiO2, Na2O, Zr and V, together with Cr, Sc and Th, representing siliciclastic detritus; (2) Al2O3, K2O, Rb and Cs is a clay-mineral association; (3) Fe2O3 and Mn reflect enrichment in a Fe–Mn component – likely including siderite, goethite, other Fe–Mn oxyhydroxides and berthierine; (4) CaO and P2O5 represent carbonate cements and phosphate; (5) TOCWR, Mo* and U are associated with organic matter. It is notable that TiO2 falls between the main siliciclastic and clay-mineral association vectors, MgO and Y lie between the siliciclastic and Fe–Mn associations and Sr plots between carbonate cements and organic matter. This pattern indicates the presence of multiple mineralogical and/or geochemical controls on these elements.

The distribution of samples on the biplot (Fig. 9) shows good separation of the stratigraphic units, reflecting the changing proportions of mineralogical constituents and geochemical environments through the succession. Outliers are caused principally by Fe – Mn enrichment in ironstone intervals within the Staithes Sandstone and Cleveland Ironstone, and extreme carbonate cementation of individual samples in the Jet Rock (Whale Stones and Top Jet Dogger). It is worth noting that the Hard Shales (Subunit Vb) cluster with samples displaying strong organic matter enrichment (upper Grey Shales – lower Bituminous Shales, Units II and IV).

10. Discussion – Geochemistry as a grain-size proxy

In the previous text, it has been inferred that positive correlations between detrital proxy element ratios (Si/Al, Ti/Al, Zr/Al and Na/Al), and their negative correlation to typical clay-mineral-associated elements (K/Al, Rb/Al, Cs/Al), provide a grain-size proxy in the mudstone succession at Dove’s Nest (Fig. 4). Limited grain-size data are available to test this interpretation. Gad et al. (Reference Gad, Catt and Le Riche1969) published grain-size data for 3 beds from the Whitby Mudstone at Whitby: Grey Shale ‘bed’ 19, 0.4% sand (>50 μm), 49.0% silt (2–50 μm), 50.6% clay (<2 μm); Jet Rock ‘bed’ 32, 0.01% sand, 48.0% silt, 52.0% clay; and Bituminous Shales ‘bed’ 43, 0.1% sand, 44.7% silt, 55.2% clay; together with results for two beds in the Alum Shale from above our study interval. These data indicate a generally fining-upward trend in the Whitby Mudstone consistent with our interpretation.

Pye & Krinsley (Reference Pye and Krinsley1986) noted a positive correlation between the silt/clay ratio and Si/Al ratio but provided no numerical data. Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) similarly argued that Si/Al, Zr/Al and Zr/Rb ratios offer proxies for coincident changes in fluvial transport and siliciclastic grain size in the Whitby Mudstone. However, mudstones may accumulate as coarser composite grains that reflect elevated bottom-water energy conditions (e.g. Schieber et al., Reference Schieber, Miclaus, Seserman, Liu and Teng2019; Z Li et al., Reference Li, Schieber and Pedersen2021), so increased grain size may not necessarily generate elevated values of detrital proxy elements.

Grain-size values determined in thin-sections of 73 samples from the Grey Shale and Jet Rock of the coastal sections at Port Mulgrave were plotted by Ghadeer (Reference Ghadeer2011, fig. 8.3). These demonstrate an overall fining-upward trend (Fig. 10) with 2 – 7% fine sand and 12 – 23% silt in the basal beds of the Grey Shale Member (‘beds’ 1 – 6) and clay dominating (<2% silt) the top of the Jet Rock (‘beds’ top 36 – 40). A large peak in silt (16 – 24%), including one sample with 10% fine sand, occurs in ‘bed’ 31 that matches the prominent Si/Al and Ti/Al peaks observed at this level both on the coast (Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; Wang, Reference Wang2022) and in the Dove’s Nest core (Figs 7, 10), and the overall silt profile displays a strong similarity to the geochemical detrital proxies.

Figure 10. Ti/Al and K/Al ratio profiles from Dove’s Nest compared to quartz grain size, proportion of silt, illite content, CIA, and Os isotopes. Quartz grain size and illite content (peak-area integration with height ratios method) after de Vos (Reference de Vos2017). Silt percentage was determined from Port Mulgrave (Ghadeer, Reference Ghadeer2011). CIA = Chemical Index of Alteration (Nesbitt & Young, Reference Nesbitt and Young1982); CIA = [AI2O3/(Al2O3 + CaO* + Na2O + K2O)] × 100 (molecular proportions). CaO* moles assumed to be equivalent to Na2O (filled blue circles) with additional values derived from CaO determinations (open blue circles), where a number of moles was less than that of Na2O (McLennan, Reference McLennan1993). Osmium isotope plot from Cohen et al. (Reference Cohen, Coe, Harding and Schwark2004, fig. 1), with composite data from 3 coastal sections: Hawsker Bottoms; Port Mulgrave; Saltwick Bay. Cleveland Basin relative sea-level curve (Hesselbo, Reference Hesselbo2008) replotted relative to biostratigraphic zones interpreted for the Dove’s Nest core. 187Os/188Os ratio of early Toarcian ocean water (0.377 ± 0.065) after van Acken et al. (Reference van Acken, Tütken, Daly, Schmid-Röhl and Orr2019). Cycles are based principally on the Ti/Al profile.

De Vos (Reference de Vos2017) undertook quartz abundance and grain-size analyses of 21 samples from the Dove’s Nest core using SEM–EDX of polished sample splits of material analysed in this study. The resulting quartz means grain-size profile closely matches the shape of the Ti/Al ratio curve (Fig. 10) and the Ti/Al ratio shows a statistically significant (p <0.001) positive correlation with grain size (Fig. 11c). The coarsest grain size, mean very fine silt but incorporating fine – coarse silt fractions, characterizes the upper Penny Nab and Kettleness members of the Cleveland Ironstone. The succession then fines upward towards a minimum in the middle of the Bituminous Shales, which are fine mudstones (Figs 10, 11), before coarsening upward (CU) into coarse mudstones of the Alum Shale.

Figure 11. Geochemical cross-plots for selected detrital proxies and grain size in the Dove’s Nest core. (a) Si/ Al vs Ti/Al. (b) Na/Al vs K/Al. (c) Ti/Al vs quartz mean grain size (de Vos, Reference de Vos2017). Mineral reference compositions from webmineral.com, average shale (PAAS) after Taylor & McLennan (Reference Taylor, McLennan and Meyers2001). Plotted regression lines are (a) ordinary least square and (b, c) reduced major axis, with 95% confidence envelope in (c). Grey shading in (b) represents the field of the Subunit IIIb illite and weathering pulse including Whale Stones ‘bed’ 35, characterized by anomalous low CIA values (Fig. 10).

11. Mineralogical trends

Detrital minerals identified in the Whitby Mudstone of the coastal succession comprise quartz, micas (muscovite, biotite), feldspars (plagioclase, K-feldspar), illite-smectite, kaolinite and chlorite (Pye & Krinsley, Reference Pye and Krinsley1986). Authigenic anatase and kaolinite are also present. The detrital sand and coarsest silt comprise mostly quartz, with minor feldspar and muscovite in both the Cleveland Ironstone and Whitby Mudstone (Ghadeer, Reference Ghadeer2011). Detrital Ti-bearing minerals including rutile and anatase, and their replacement product, leucoxene, are a consistent accessory phase. Houben et al. (Reference Houben, Barnhoorn, Wasch, Trabucho-Alexandre, Peach and Drury2016b) provided quantitative data for 4 samples from the Whitby Mudstone that yielded: 13.1 – 16.3% quartz; 0 – 1% plagioclase; 1.3 – 15.1% calcite; 0.1 – 2.8% ankerite; 0.6 – 0.9% anatase; and 8.0 – 11.0% pyrite.

11.a. Clay minerals

Gad et al. (Reference Gad, Catt and Le Riche1969) reported that clay fractions in the Whitby Mudstone contain approximately equal amounts of kaolinite, mica (illite) and swelling clays (vermiculite and/or smectite) with minor chlorite (<5%). The silt fraction includes clay-mineral aggregates. Houben et al. (Reference Houben, Barnhoorn, Lie-A-Fat, Ravestein, Peach and Drury2016a, b) recorded 31 – 54% illite–smectite, 17 – 27% illite, 26 – 40% kaolinite and 3 – 4% chlorite in four samples from the formation.

The clay mineralogy of the 21 samples from the Dove’s Nest core was determined by de Vos (Reference de Vos2017). Quantitative clay mineral analysis employing the XRD peak-area integration with peak-height ratio method showed kaolinite (27 ± 5%), illite (48 ± 4%) and vermiculite (25 ± 5%), with a dominance of irregularly interstratified illite–vermiculite (I/V), to be the principal components of the clay fraction (<4 μm). Muscovite and chlorite are present in minor amounts, the latter only in the lower part of the succession below 175 m (Staithes Sandstone – Grey Shale; Units I and II).

The results of de Vos (Reference de Vos2017) are broadly consistent with earlier clay mineral studies of Pliensbachian – Toarcian sedimentary rocks in the Cleveland Basin by Morris (Reference Morris1980), Sellwood & Sladen (Reference Sellwood and Sladen1981), Jeans (Reference Jeans2006) and Houben et al. (Reference Houben, Barnhoorn, Lie-A-Fat, Ravestein, Peach and Drury2016a, b). Those authors emphasized the dominance of illite and kaolinite with mixed-layer clays, although they did not specify the presence of a I/V phase. Kemp & McKervey (Reference Kemp and McKervey2001) and SJ Kemp et al. (Reference Kemp, Merriman and Bouch2005) reported illite–smectite as the interstratified species but acknowledged difficulties in the interpretation of the XRD spectra. Precise comparison between the studies is hampered by differences in analytical methodology and clay mineral terminology.

Most authors have regarded the Lias Group clay mineral assemblages in the Cleveland Basin as being principally detrital (Sellwood & Sladen, Reference Sellwood and Sladen1981; Jeans, Reference Jeans2006; Dera et al., Reference Dera, Pellenard, Neige, Deconinck, Pucéat and Dommergues2009) despite a maximum burial depth of 3 – 4 km and temperatures of 100 – 120° C (SJ Kemp et al., Reference Kemp, Merriman and Bouch2005), although evidence exists for the development of some authigenic kaolinite (Pye & Krinsley, Reference Pye and Krinsley1986).

The relatively high abundance of kaolinite in the Upper Lias throughout northern Europe suggests warm temperatures and near-constant annual humid conditions on adjacent landmasses (Singer, Reference Singer1984; Chamley, Reference Chamley1989; Fagel, Reference Fagel, Hillaire-Marcel and De Vernal2007; Dera et al., Reference Dera, Pellenard, Neige, Deconinck, Pucéat and Dommergues2009). Local factors may also have played a role. For example, in Scotland, the sand and clay mineralogy of Jurassic sedimentary rocks suggests that they were derived from erosion of Devonian and Carboniferous strata (Hurst, Reference Hurst1985; Hall & Bishop, Reference Hall, Bishop, Doré, Cartwright, Stoker, Turner and White2002). The high detrital kaolinite content of Rhaetian to Sinemurian mudstones in northeast Scotland has been attributed to the reworking of Carboniferous regoliths, which were uplifted and eroded at the time of formation of the Viking Graben (Hurst, Reference Hurst1985).

The kaolinite content in the Whitby Mudstone of the Dove’s Nest core (de Vos, Reference de Vos2017) is lower (21 – 30%) than in the underlying Staithes Sandstone and Cleveland Ironstone (24 – 40%) which might imply a transition to cooler and dryer climate. However, this is contrary to trends elsewhere, including at Mochras in the Cardigan Bay Basin (Fig. 1a), that show widespread kaolinite enrichment in the lower Toarcian (Raucsik & Varga, Reference Raucsik and Varga2008; Dera et al., Reference Dera, Pellenard, Neige, Deconinck, Pucéat and Dommergues2009; Brański, Reference Brański2012; Hermoso & Pellenard, Reference Hermoso and Pellenard2014; Xu et al., Reference Xu, Ruhl, Jenkyns, Leng, Huggett, Minisini, Ullmann, Riding, Weijers, Storm, Percival, Tosca, Idiz, Tegelaar and Hesselbo2018) and palaeotemperature records (Section 18) that indicate generally cool climate in the late Pliensbachian, a short-lived hyperthermal during the T-OAE, with persistence of warm climate conditions thereafter (Suan et al., Reference Suan, Mattioli, Pittet, Lecuyer, Sucheras-Marx, Duarte, Philippe, Reggiani and Martineau2010; Korte et al., Reference Korte, Hesselbo, Ullmann, Dietl, Ruhl, Schweigert and Thibault2015; Bougeault et al., Reference Bougeault, Pellenard, Deconinck, Hesselbo, Dommergues, Bruneau, Cocquerez, Laffont, Huret and Thibault2017; Ruebsam & Schwark, Reference Ruebsam, Schwark, Reolid, Duarte, Mattioli and Ruebsam2021).

The lower kaolinite content in the lower Toarcian at Dove’s Nest may reflect clay mineral fractionation with increasing distance to shore (Gibbs, Reference Gibbs1977; Godet et al., Reference Godet, Bodin, Adatte and Föllmi2008; Deconinck et al., Reference Deconinck, Hesselbo and Pellenard2019), driven by relative sea-level rise accompanying the early Toarcian transgression (Hallam, Reference Hallam1997, Reference Hallam2001; Hesselbo, Reference Hesselbo2008) and the impact of extreme seasonal aridity during the T-OAE (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019). Similarly, a pulse of kaolinite at the base of the Kettleness Member (Fig. 10) with falling values above, may reflect nearer-shore conditions associated with the depositional hiatus that resulted in the regional disconformity at the base of the P. spinatum Zone, followed by sea-level rise. However, changes in sediment provenance cannot be discounted as a mechanism to explain the mineralogical trends.

Illite abundance is associated with weak weathering intensities and increased levels of mechanical weathering, implying terrestrial erosion (Singer, Reference Singer1984). Illite contents at Dove’s Nest, ascribed to a I/V mixed layer component (de Vos, Reference de Vos2017), are relatively constant throughout the succession except for the T-OAE interval (Subunits IIa, b), which shows a marked rise to 54 – 64% illite from an average of 46% below, before falling back to an average of 47% above the event (Fig. 10). The increase in illite in the I/V fraction has been ascribed to higher terrestrial erosion rates leading to deeper erosion of podzol soil profiles (de Vos, Reference de Vos2017). The implied elevated levels of continental erosion during the earlier phases of the T-OAE match the increase of continental weathering flux of >40% indicated by the increase in 187Os/188Os ratio (Section 16.b.4) from ∼0.4 to ∼1.0 in the top Grey Shale and Jet Rock on the coast (Fig. 10; Cohen et al., Reference Cohen, Coe, Harding and Schwark2004), and coincident increases in the Mochras borehole (Percival et al., Reference Percival, Cohen, Davies, Dickson, Hesselbo, Jenkyns, Leng, Mather, Storm and Xu2016), in western North America (Them et al., Reference Them, Gill, Selby, Gröcke, Friedman and Owens2017b) and in Japan (Kemp et al., Reference Kemp, Selby and Izumi2020).

11.b. Detrital proxies in the silt fraction

Mineralogical controls on bulk-rock geochemistry may be investigated using element-ratio Pearson correlations. Detrital proxy element ratios Si/Al, Ti/Al, Na/Al and Zr/Al show high positive correlations with each other of r = ≥0.62 (p = 0.0001), most >0.75 and are negatively correlated to K/Al (r = <−0.46, p = 0.0001). Of these, Ti/Al has the highest correlation with mean grain size (r = 0.73) and particularly with the medium silt fraction (r = 0.76), although all other proxies show similar relationships.

Figure 11 shows example element-ratio cross-plots of (a) Si/Al vs Ti/Al; (b) Na/Al vs K/Al; and (c) Ti/Al vs quartz mean grain size for Dove’s Nest. The geochemical units and their constituent lithostratigraphic divisions display discrete clusters. Si/Al in the Staithes Sandstone (Subunit Ia) shows the greatest scatter (Fig. 11a) attributable to large sample-to-sample differences in quartz abundance in the coarser fractions. The Cleveland Ironstone (Subunits Ib, Ic) shows a tight positive array on the Si/Al vs Ti/Al plot attributable to a close association between Ti-bearing heavy minerals (rutile, anatase) and quartz in a changing silt fraction. Claystones (Fig. 11c) of the Bituminous Shales (Unit IV, Subunit Va) display low and constant Si/Al and Na/Al ratios (Fig. 11a, b) but significant variation in Ti/Al and K/Al reflecting variable clay mineralogy.

The remarkable similarity between the Si/Al and Na/Al profiles (Fig. 4) and a high positive correlation between Si/Al and Na/Al (r = 0.93) point to Na-rich plagioclase as a dominant mineralogical control, likely present as both a primary detrital fraction and as a replacement of K-feldspar (cf. Min et al., Reference Min, Zhang, Li, Zhao, Li, Lin and Wang2019). The negative correlation between Na/Al and K/Al (Fig. 11b) is attributed to the presence of Na-plagioclase in the silt fractions and dominance of illite in the clay fraction, with scatter generated particularly by the presence of variable K-feldspar and micas. The large scatter and high values of K/Al in the mudstones from the T-OAE are particularly significant and are explained by pulses of illitic clay during the event (cf. Fig. 10).

Zr/Al shows the strongest correlation to Si/Al but a more irregular profile (Fig. 4) which may be explained by the nugget effect of zircon grains present in the small sample sizes analysed. The resulting element cross-plots (not shown) are like those of Si/Al.

11.c. Chemical Index of Alteration - CIA

The CIA (Nesbitt & Young, Reference Nesbitt and Young1982; McLennan, Reference McLennan1993) is a widely applied chemical weathering proxy and has been used in many recent T-OAE studies (e.g. Fu et al., Reference Fu, Wang, Zeng, Feng, Wang and Song2017; Fantasia et al., Reference Fantasia, Föllmi, Adatte, Bernárdez, Spangenberg and Mattioli2018a, b; Bomou et al., Reference Bomou, Suan, Schlögl, Grosjean, Suchéras-Marx, Adatte, Spangenberg, Fouché, Zacaï, Gibert, Brazier, Perrier, Vincent, Janneau, Martin, Reolid, Duarte, Mattioli and Ruebsam2021; Fu et al., Reference Fu, Wang, Wen, Song, Wang, Zeng, Feng and Wei2021; Alnazghah et al., Reference Alnazghah, Koeshidayatullah, Al-Hussaini, Amao, Song and Al-Ramadan2022; Liu et al., Reference Liu, Cao, He, Liang, Pu and Wang2022). A ternary plot of Al2O3 – CaO*+Na2O – K2O for the Dove’s Nest data is presented in Figure 12 with the corresponding CIA values. These are plotted stratigraphically in Figure 10. Due to the large variation in carbonate content between samples, particularly in the T-OAE interval (Unit III, Fig. 7), Figure 10 plots CIA values calculated using CaO* = Na2O (blue filled circles and blue line) for all samples and using CaO* = phosphate corrected CaO (open blue circles) for samples where the remaining number of moles was less than that of Na2O (following McLennan, Reference McLennan1993). It is recognized that the CIA profile based on the former method is identical to that for Ln(Al2O3/Na2O) in molar proportions (not plotted) which is an alternative weathering proxy applicable for moderate to high carbonate samples (Von Eynatten et al., Reference Von Eynatten, Barceló-Vidal and Pawlowsky-Glahn2003; Xia & Mansour, Reference Xia and Mansour2022). The CIX proxy of Garzanti et al. (Reference Garzanti, Padoan, Setti, López-Galindo and Villa2014) that uses a modification of the CIA formula excluding CaO generates the same trend.

The CIA data for very low carbonate samples that enable the use of CaO* values for calculation generally display higher CIA values and lower amplitude change through the succession but show the same stratigraphic trends as the Na2O substituted values (Fig. 10). The reduced amplitude is due primarily to the high Na2O content of the upper Pliensbachian sediments (Unit I, Figs 4, S1). The long-term trend shows increasing upwards CIA values representing weak to moderate chemical weathering of <70 (Na2O-substituted values) in the Staithes Sandstone (Subunit Ia), moderate chemical weathering ∼70 – 75 in the Cleveland Ironstone (Subunits Ib, Ic), rising to 75 – 80 in the basal Toarcian (Units II, III) including the T-OAE, reaching high and constant values of >80, indicative of intense chemical weathering, through the remainder of the lower Toarcian – middle Toarcian (Unit IV, V) section.

The CIA profile for the upper Pliensbachian displays a cyclicity that is inversely correlated to Si/Al, Ti/Al and Zr/Al and positively correlated to K/Al and Rb/Al (compare Figs 4 and 10), indicating a strong grain-size control. For comparison, Dinis et al. (Reference Dinis, Garzanti, Vermeesch and Huvi2017, fig. 3) documented a 10-unit offset to lower CIX values in river sands compared to associated river muds in their study of weathering proxies along the modern SW African margin, although parallel trends that broadly followed the regional weathering trends were observed for the two size fractions.

For Dove’s Nest, coarser-grained Pliensbachian sedimentary rocks have lower CIA values due to lower clay and higher plagioclase contents that might conceivably represent cycles of chemical weathering intensity but are attributed principally to grain-size variation linked to cyclic changes in bottom energy conditions and shoreline proximity. The sequence stratigraphy (Section 8.a.2) indicates that these were driven by changes in relative sea level (Howard, Reference Howard1985; Young et al., Reference Young, Aggett, Howard and Young1990; Macquaker & Taylor, Reference Macquaker and Taylor1996).

The general anticorrelation of CIA with Ti/Al and the other detrital proxies continues through the Toarcian at Dove’s Nest but the positive correlation to K/Al and Rb/Al is reversed in T-OAE Subunits IIIb – c where increasing K/Al ratios are linked to a pulse of illitic clay (Fig. 10). Notably, a Th/K minimum at this level is a feature of the field spectral gamma-ray stratigraphic profiles generated for the Yorkshire coast Pliensbachian – Toarcian by Myers & Wignall (Reference Myers, Wignall, Leggett and Zuffa1987) and Parkinson (Reference Parkinson1996). The latter considered Th/K to be a likely proxy for the kaolinite/illite ratio. The illite pulse correlates to a major global weathering event documented by a large increase in 187Os/188Os values in the coastal succession. Si/Al and K/Al profiles obtained for the T-OAE interval and adjacent beds by Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) and Wang (Reference Wang2022) on the coast display identical trends to those seen in the core (Fig. 7) but those authors were unable to calculate CIA values due to an absence of Na2O data.

The continental weathering pulse indicated by the positive Os excursion coincides with a negative excursion in CIA values at Dove’s Nest (Fig. 10). This is the opposite of what might be expected since increased chemical weathering would lead to higher CIA values. The T-OAE interval elsewhere is commonly characterized by a prominent increase in CIA values and other chemical weathering proxies (Fantasia et al., Reference Fantasia, Föllmi, Adatte, Bernárdez, Spangenberg and Mattioli2018a, Reference Fantasia, Adatte, Spangenberg, Font, Duarte and Föllmi2019; Bomou et al., Reference Bomou, Suan, Schlögl, Grosjean, Suchéras-Marx, Adatte, Spangenberg, Fouché, Zacaï, Gibert, Brazier, Perrier, Vincent, Janneau, Martin, Reolid, Duarte, Mattioli and Ruebsam2021; Wang, Reference Wang2022) attributable to a warmer and wetter climate during the Toarcian hyperthermal.

In the Cleveland Basin, a pulse of detrital illite combined with a short-term upward increase in sediment grain size drove falling CIA values through the δ13C minimum of the T-OAE (Subunit IIIb), despite evidence of an increasingly hot climate with extreme wet/dry seasonality based on terrestrial palynomorph assemblages (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019). Seasonality peaked during the later phases of the T-OAE (Subunits IIIc – d), associated with a sharp increase in CIA and lower illite content (Fig. 10). A peak of charcoal abundance at this level in the Peniche and Mochras successions (Baker et al., Reference Baker, Hesselbo, Lenton, Duarte and Belcher2017) is consistent with dryer climate conditions (increased wildfires) in the European region following a wetter phase during the initial stages of the T-OAE.

The charcoal maximum correlates to a coincident peak in the relative abundance of Cerebropollenites. The proliferation of Cerebropollenites during multiple past hyperthermal events suggests that the source plants were adapted to hot, arid climates able to survive in warm, drought-like conditions (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019). In the case of the Cleveland Basin, therefore, it is suggested that a wetter climate during the interval of the δ13C minimum led to increased detrital input that overprinted the normal chemical weathering proxies. A change from dominantly wet to dominantly dry conditions indicated by the charcoal and palynological proxy records may reflect a temporal shift in the relative importance of dry vs wet in a strongly seasonal wet/dry cycle.

11.d. Fe–Mn enrichment and carbonate cements

The Pearson correlation and PCA results for Dove’s Nest show a close association between Fe, Mn and Mg, with r values of >0.6 (Table S3) and with Ca and P falling in the same sector of negative PC1 and positive PC2 scores (Fig. 9). The relationship between Fe–Mn enrichment and carbonate may be illustrated using a Fe+Mn–Ca–Mg ternary plot (Fig. 13). As with the detrital proxy elements (Fig. 11), the chemostratigraphic units display distinctive clusters and arrays on the Fe+Mn–Ca–Mg plot. The Pliensbachian formations of Unit I form a low-Ca cluster adjacent to muscovite and kaolinite mineral analyses from the Cleveland Ironstone (Aggett, Reference Aggett1990); Ca values lower than average shale (PASS) indicate negligible carbonate in most samples. Individual analyses falling outside the cluster incorporate calcite grains and/or cements, generating an array of scattered points towards the Ca vertex (Fig. 13) or are siderite-rich, exemplified by the Avicula Seam Ironstone (sample 200.39) the bulk chemistry of which displays high Fe+Mn and falls close to the siderite analyses of Aggett (Reference Aggett1990). Generally low Si values of Fe-rich samples indicate that berthierine is not a significant mineral component in the sample suite analysed here.

Figure 12. Ternary diagram of Al2O3–(CaO*+Na2O)–K2O in Dove’s Nest rock samples. Values are molecular proportions. (a) CIA = Chemical Index of Alteration (Nesbitt & Young, Reference Nesbitt and Young1982) with CaO* moles assumed to be equivalent to Na2O (McLennan, Reference McLennan1993; see text). Tonalite, granodiorite and granite compositions after Condie (Reference Condie1993). (b) Enlargement of plotted data in the top sector of the diagram (red outline in a) with selected sample details. Grey shading represents the field of the Subunit IIIb illite and weathering pulse with high K/Al and anomalous low CIA values.

Figure 13. Ternary plot of iron and carbonate-associated elements Fe+Mn–Ca–Mg in Dove’s Nest rock samples compared to constituent mineral compositions. Stars are electron microprobe determinations of mineral fractions from the Cleveland Ironstone of Staithes (Aggett, Reference Aggett1990). Open triangles are calculated pure mineral values (webmineral.com). Average shale (PASS) composition after Taylor & McLennan (Reference Taylor, McLennan and Meyers2001). Plot is scaled based on the maximum and minimum values of the three components (cf. de Lange et al., Reference de Lange, Jarvis and Kuijpers1987).

Unit II, the lower Grey Shale, forms a tight cluster close to average shale (Fig. 13). The carbonaceous mudstones of Units III and IV form long scattered arrays between moderate Fe+Mn–Mg contents and the Ca vertex. The very high carbonate content of the Whale Stones sample (167.89), interpreted to represent a carbonate concretion, is demonstrated by its proximity to the calcite mineral analysis (Aggett, Reference Aggett1990). Samples from Units III and IV fall along a mixing line between calcite, ankerite and kaolinite. This may be fortuitous since the scatter towards the Fe+Mn vertex is likely attributable in part to the presence of pyrite which is abundant in these units (e.g. Houben et al., Reference Houben, Barnhoorn, Wasch, Trabucho-Alexandre, Peach and Drury2016b) but was not quantified in this study. Nonetheless, ankerite has been recorded as a minor mineral phase in these intervals (Houben et al., Reference Houben, Barnhoorn, Wasch, Trabucho-Alexandre, Peach and Drury2016b). Offset of Unit III samples to higher Mg values is attributed to the presence of dolomite (cf. Pye, Reference Pye1985).

Unit V samples cluster in the same low-Ca area of the Fe+Mn–Ca–Mg plot as those from the Cleveland Ironstone (Subunits Ib, Ic) indicating a return to less Fe–Mn enrichment and generally low carbonate contents (compare Fig. 9). Samples from the Hard Shales (Subunit Vb) are variably displaced to higher Ca values (Fig. 13), reflecting the presence of calcite cements.

12. Chemostratigraphic correlation and palaeoredox proxies

Chemostratigraphic correlation between Dove’s Nest and the Yorkshire coast may be further illustrated by profiles for the redox-sensitive constituents TOC/PT, FeEF, MnEF and PEF (Fig. 14), which provide insights into facies and bottom-water evolution in the Cleveland Basin during the late Pliensbachian – middle Toarcian. High-resolution data are available for the lower Toarcian of both sections and excellent agreement exists between the fine structure and values of the profile pairs. Key marker bed horizons (e.g. Sulphur Bands 2, 3, Whale Stones, Top Jet Dogger, Millstones, Peak Stones, Ovatum Band; see also Fig. 7) provide a framework for bed-scale correlation between the sections. The Dove’s Nest data additionally offer higher resolution and extended stratigraphic coverage of the upper Pliensbachian and basal middle Toarcian.

Figure 14. Correlation of TOC/PT, DOPT, FeEF, MnEF and PEF between the Dove’s Nest core and Yorkshire coastal outcrop sections. Coast geochemical profiles from McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008), McArthur (Reference McArthur2019) and Remírez & Algeo (Reference Remírez and Algeo2020) with additional high-resolution data (Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018); Yorkshire chemostratigraphic Units I – V modified from Remírez & Algeo (Reference Remírez and Algeo2020). Stratigraphy as in Figs 2, 4. WS = Whale Stones (‘bed’ 35’); TJD = Top Jet Dogger (‘bed’ 39). Enrichment factors (EF) are calculated relative to PASS. Data of Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) are recalibrated relative to McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008) and McArthur (Reference McArthur2019): stratigraphic heights are increased by 1.1 m; Al values are increased by 20% to remove analytical bias. Vertical grey dotted lines are EF values of 1. Vertical green dotted lines indicate the Redfield ratio, a TOC/P ratio of ∼106:1, typical of marine plankton biomass (Redfield et al., Reference Redfield, Ketchum, Richards and Hill1963). Values of >106 indicate P-release from the sediment under reducing conditions. Vertical dotted lines on the TOC/PT and DOPT plots mark the positions of redox boundaries typically associated with values of 50 and 0.25 (red, oxic/suboxic), 106 and 0.5 (green, dysoxic/anoxic) respectively, following Algeo & Ingall (Reference Algeo and Ingall2007) and Algeo & Maynard (Reference Algeo and Maynard2004). More conservative threshold DOP values of <0.45 for oxic or dysoxic depositional environments and >0.75 for a euxinic environment (gold vertical dotted line) have been proposed by Raiswell et al. (Reference Raiswell, Hardisty, Lyons, Canfield, Owens, Planavsky, Poulton and Reinhard2018).

12.a. TOC/PT molar ratio

The TOC/PT ratio is a redox proxy based on the differing processes and rates of organic carbon and phosphorus remineralization under varying redox conditions (Ingall et al., Reference Ingall, Bustin and Van Cappellen1993; Van Cappellen & Ingall, Reference Van Cappellen and Ingall1994; Algeo & Ingall, Reference Algeo and Ingall2007; Algeo & Li, Reference Algeo and Li2020; Papadomanolaki et al., Reference Papadomanolaki, Lenstra, Wolthers and Slomp2022). Marine plankton biomass has a remarkably constant TOC/P molar ratio of 106:1, the Redfield ratio (Redfield et al., Reference Redfield, Ketchum, Richards and Hill1963), and lacustrine particulates have a comparable composition (Hecky et al., Reference Hecky, Campbell and Hendzel1993). In modern oxic marine bottom waters and surficial sediments, organic carbon is mostly returned to the water column during bacterial respiration, but P is trapped by adsorption onto Fe-oxyhydroxides and the precipitation of carbonate-fluorapatite or its precursor in the sediment during iron reduction (Jarvis et al., Reference Jarvis, Burnett, Nathan, Almbaydin, Attia, Castro, Flicoteaux, Hilmy, Husain, Qutawnah, Serjani and Zanin1994, fig. 3; März et al., Reference März, Riedinger, Sena and Kasten2018). These processes lead to a fall in the sediment TOC/PT molar ratio with values typically < 50 in oxic facies (Algeo & Ingall, Reference Algeo and Ingall2007). Under anoxic conditions the preservation of organic matter is enhanced and, in the absence of Fe-oxyhydroxides, reactive P is released back to seawater leading to an increased TOC/PT molar ratio of > 106. Modern anoxic sediments generally display ratios of 100 – 200, with values exceeding 1000 in some ancient black shales (Algeo & Ingall, Reference Algeo and Ingall2007; Papadomanolaki et al., Reference Papadomanolaki, Lenstra, Wolthers and Slomp2022).

TOC/PT molar ratios in the upper Pliensbachian Unit I at Dove’s Nest are indicative of oxic facies with a mean value of 52 ± 26 (Fig. 14). A small upward increase in the average ratio is seen from the Staithes Sandstone to the Kettleness Member of the Cleveland Ironstone, between three Subunits Ia – c. A mean TOC/PT molar ratio of 57 ± 24 in lower Grey Shale Unit II indicates predominantly oxic conditions with the notable exceptions of Sulphur Bands 2 and 3 that exceed 120 (SB1 was not sampled in the core but displays a TOC/PT ratio of >200 on the coast), evidencing short-lived episodes of anoxia – euxinia.

High TOC/PT molar ratios that consistently exceed the Redfield ratio characterize T-OAE Unit III of the top Grey Shale and Jet Rock (Fig. 14), with the highest values occurring in Subunit IIIb (230 ± 69). These indicate that significant quantities of P must have been released back to anoxic bottom waters during the peak of the T-OAE. TOC/PT ratios fall sharply in Subunit IIId but remain higher in Units IV and V than in the pre-T-OAE section (Units I and II) indicating continuing oxygen depletion in bottom waters during the later early Toarcian and early middle Toarcian, fluctuating within the dysoxic field with a molar ratio of 85 ± 18 (Fig. 14).

An identical pattern is displayed by Yorkshire coast data but with higher TOC/PT molar ratios recorded by the ultra-high-resolution data obtained for T-OAE Unit III (Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) with values of 390 ± 104 for Subunit IIIb (Figs 14, 15).

Figure 15. Bottom-water redox proxy interpretation for the upper Pliensbachian – lower Toarcian of the Cleveland Basin derived from TOC, TOC/P, DOP and Fe speciation. δ13Corg profile for Dove’s Nest (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022, black) rescaled to match coast composite data (see Fig. 2 for sources). Rescaled whole-rock TOC profile for Dove’s Nest (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022, dark green) with coast composite data of Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011; thin yellow-green high-resolution curve) and trend of the low-resolution coast datasets of Ruvalcaba Baroni et al. (Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018, open triangles) and McArthur (Reference McArthur2019) (thin pale green low-resolution curve; see Fig. 2). Average shale and black shale values as in Figure 2; ‘anoxic threshold’ of TOCWR = 2.5 wt% follows Algeo & Maynard (Reference Algeo and Maynard2004). Low-resolution TOC/PT (dark green) and DOPT (dark orange) curves from McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008) and Remírez & Algeo (Reference Remírez and Algeo2020) with high-resolution data (thin pale curves) from Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018); see Figure 14 for further information. Iron speciation data from Salem (Reference Salem2013, cream-filled circles) and Houben et al. (Reference Houben, Goldberg and Slomp2021, yellow-filled circles) with redox field boundaries after Raiswell et al. (Reference Raiswell, Hardisty, Lyons, Canfield, Owens, Planavsky, Poulton and Reinhard2018). Extinction levels (i)– (iii) after Caswell et al. (Reference Caswell, Coe and Cohen2009). Pliensb. = Pliensbachian; Bitumin. Sh. = Bituminous Shales; D. semic. = Dactylioceras semicelatum; Dt = D. tenuicostatum; Dc = D. clevelandicum; P. hawk. = Pleuroceras hawskerense; Pa = P. apyrenum.

12.b. Degree of pyritization DOPT

The ‘degree of pyritization’ (DOP, Raiswell & Berner, Reference Raiswell and Berner1985) is an established proxy for bottom-water redox conditions (Raiswell et al., Reference Raiswell, Buckley, Berner and Anderson1988, Reference Raiswell, Hardisty, Lyons, Canfield, Owens, Planavsky, Poulton and Reinhard2018). DOP values exceeding 0.5 have been suggested to provide evidence of euxinia (free H2S in the water column) driving the addition of syngenetic pyrite precipitated in the water column to the sediment (Raiswell et al., Reference Raiswell, Newton and Wignall2001; Algeo & Maynard, Reference Algeo and Maynard2004), although more conservative threshold DOP values of <0.45 for oxic or dysoxic depositional environments and >0.75 for a euxinic environment were proposed by Raiswell et al. (Reference Raiswell, Hardisty, Lyons, Canfield, Owens, Planavsky, Poulton and Reinhard2018).

Although originally defined as DOP = pyrite iron/total reactive iron (Berner, Reference Berner1970; Raiswell et al., Reference Raiswell, Buckley, Berner and Anderson1988), values may be closely approximated using whole-rock total iron and total sulfur data. Algeo & Li (Reference Algeo and Li2020) adopted the equation:

$${\rm{DO}}{{\rm{P}}_{\rm{T}}} = {{\rm{S}}_{\rm{T}}} \times \left( {{\rm{55}}{\rm{.85/64}}{\rm{.12}}} \right){\rm{/F}}{{\rm{e}}_{\rm{T}}}$$

where total sulfur (ST) and total iron (FeT) and the coefficient 55.85/64.12 represent the weight ratio of Fe/S in stoichiometric pyrite. The use of ST and FeT results in the inclusion of some non-pyrite sulfur (e.g. organic sulfur) and some non-reactive iron (e.g. silicate- or carbonate-bound Fe) in the calculation of the DOPT ratio. These amounts are typically small but, nonetheless, they may result in different values of DOP and DOPT for a given sample (e.g. Algeo et al., Reference Algeo, Rowe, Hower, Schwark, Merrmann and Heckel2008; Algeo & Liu, Reference Algeo and Liu2020). DOP vs DOPT correlations are formation-specific, so DOPT palaeoredox thresholds determined in the literature cannot be applied universally without validation.

Several studies have been undertaken on the Yorkshire coastal succession that include DOP data. DOP values of 0.84 ± 0.03 were reported for the Jet Rock (C. exaratum Subzone) by Raiswell & Berner (Reference Raiswell and Berner1985) and between 0.8 and 0.9 by Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008). Recent studies on the Yorkshire coastal succession have used ST and FeT data, where either DOPT = 0.95 × ST/FeT, the decrease in S content incorporated to allow for the presence of organic S (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008) or simply DOPT = ST/FeT (Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018). DOPT values calculated by McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008) and Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) are comparable (0.84 ± 0.09 for Subunit IIIb) to published DOP results. Such high values significantly exceed the 0.75 threshold for euxinic environments (Figs 14, 15; Raiswell et al., Reference Raiswell, Hardisty, Lyons, Canfield, Owens, Planavsky, Poulton and Reinhard2018) and offer very strong evidence for long-term euxinic bottom-water conditions in the Cleveland Basin during the T-OAE (Raiswell & Berner, Reference Raiswell and Berner1985; Raiswell et al., Reference Raiswell, Bottrell, Al-Biatty and Tan1993; Wignall et al., Reference Wignall, Newton and Little2005; McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Pearce et al., Reference Pearce, Cohen, Coe and Burton2008; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018). Nonetheless, petrographic evidence shows levels of fine-scale bioturbation even during this interval, despite an absence of visible trace fossils (Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023).

Sulfur contents and Fe speciation were not determined in the Dove’s Nest samples. Additionally, it should be noted that the presence of significant siderite and berthierine in parts of the Yorkshire succession, particularly in the Cleveland Ironstone, will be particularly significant for raising FeT values (e.g. Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018, fig. 7A) and will generate lower than true DOP values in samples containing these minerals.

DOPT profiles for the coastal succession from McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008) and Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) are illustrated in Figures 14 and 15. The DOPT curve morphology and redox interpretation are very similar to those provided by TOC/PT. In both cases, largely oxic conditions with possible minor dysoxic intervals characterized the late Pliensbachian (Unit I). Oxic conditions prevailed during the earliest Toarcian but brief intervals of anoxia – euxinia first appeared at the start of the Toarcian (Sulphur Band 1) and reoccurred during the deposition of Sulphur Bands 2 and 3. A general drift towards persistent dysoxia preceded the onset of T-OAE which was marked by a rapid shift to anoxia – euxinia that prevailed in the area, with varying intensity, throughout the event.

T-OAE Unit III is the only interval where DOPT values consistently fluctuate above the threshold of >0.75 for euxinic environments proposed by Raiswell et al. (Reference Raiswell, Hardisty, Lyons, Canfield, Owens, Planavsky, Poulton and Reinhard2018), with the highest values present in Subunits IIIb and c. A decrease in pyrite framboid mean size to ≤5 μm accompanying rising TOC and TOC/PT trends in Subunit IIIa is indicative of a fine fraction component precipitated from a euxinic water column (Wignall & Newton, Reference Wignall and Newton1998; Wignall et al., Reference Wignall, Newton and Little2005). Song et al. (Reference Song, Littke and Weniger2017) reported >95% of <7 μm syngenetic framboidal pyrite in T-OAE carbonaceous mudstones from Runswick Bay.

The TOC/PT and DOPT proxies show different trends following the OAE (Fig. 14): the former indicates a rapid return to fluctuating dysoxic conditions that prevailed through the remainder of the early Toarcian into the early middle Toarcian. The DOPT proxy implies that anoxia persisted through most of the H. falciferum Subzone (Unit IV) prior to a shift to dysoxia in late H. falciferum Subzone time (Unit V). This interval of divergence coincides with a period of weaker basin restriction indicated by Mo vs TOC proxy (Section 13a) but also occurs with a step increase in δ34SCAS values (Section 16.a.2) interpreted to reflect the impact of global pyrite burial and indicate a significant fall in oceanic sulfate content.

12.c. Iron speciation FeHR/FeT and Fepy/FeHR

Bottom-water redox interpretations for the Cleveland Basin derived from DOPT values and pyrite framboid size may be further assessed using sedimentary rock Fe-speciation data obtained from the Yorkshire coast succession by Salem (Reference Salem2013) and Houben et al. (Reference Houben, Goldberg and Slomp2021). These proxies are based on the presence or absence of enrichments in key iron minerals (Raiswell & Canfield, Reference Raiswell and Canfield1998; Raiswell et al., Reference Raiswell, Hardisty, Lyons, Canfield, Owens, Planavsky, Poulton and Reinhard2018) with data derived from chemical extracts of bulk rock samples. Highly reactive iron (FeHR) comprises ferric oxides, iron carbonates, magnetite and pyrite. A highly reactive iron to total iron ratio (FeHR/FeT) of below 0.38 is indicative of oxic – dysoxic bottom waters with values <0.22 unambiguously oxic. Values >0.38 provide strong evidence for anoxic depositional conditions (Fig. 15).

The extent of pyritization of the highly reactive Fe pool (Fepy/FeHR) indicates whether the bottom water was anoxic (i.e. oxygen depleted and ferruginous) or euxinic (Fe-free and sulfidic), provided that the FeHR/FeT ratio is above 0.38. Fepy/FeHR values of <0.7 reflect ferruginous water column conditions. The threshold between anoxic and euxinic bottom-water conditions is less well-defined but is generally placed between 0.7 and 0.8 (Fig. 15; Raiswell et al., Reference Raiswell, Hardisty, Lyons, Canfield, Owens, Planavsky, Poulton and Reinhard2018). In cases where Fepy/FeHR ratios of >0.7 indicate sulfidic conditions but FeHR/FeT is <0.38 (i.e. oxic – dysoxic bottom water) H2S production is thought to occur close to but below the sediment/water interface.

The use of Fe-speciation proxies needs to be assessed critically (Raiswell et al., Reference Raiswell, Hardisty, Lyons, Canfield, Owens, Planavsky, Poulton and Reinhard2018). For example, carbonate-rich samples may include additional Fe2+ incorporated into diagenetic carbonates, which may lead to an overestimation of anoxia when Fe-speciation is applied to sediments with low FeT or low TOC contents (<0.5%). In this study, the Fe content of upper Pliensbachian – middle Toarcian rocks from Dove’s Nest, for example, fluctuates around the average shale value of 5.1% (Fig. S1), with the lowest value of 1.7% Fe (with 3.9% TOC) obtained from the Whale Stones concretion sample, so palaeoredox interpretations should be robust.

FeHR/FeT and Fepy/FeHR data from the Yorkshire coast (Fig. 15) support interpretations derived from the TOC/PT and DOPT profiles. Highly variable redox conditions during the earliest Toarcian are indicated by coincident high FeHR/FeT and Fepy/FeHR values within the three Sulphur Bands (SB1 – 3), indicating three major episodes of bottom-water euxinia. Additionally, cm-scale sampling through the Sulphur Bands at Port Mulgrave and Hawsker Bottoms by Salem (Reference Salem2013) evidences shorter-term fluctuating palaeoredox conditions during their deposition, with cycling between euxinic and dysoxic conditions. The bioturbated mudstones below, between and above the Sulphur Bands show less enrichment of TOC, reactive iron and trace metals, with lower TOC/PT and DOPT values (Fig. 15), but nonetheless yield FeHR/FeT ratios that fluctuate around the Fe-proxy threshold characteristic of anoxia (0.38), equating to the dysoxic intervals of the TOC/PT and DOPT proxies. These observations demonstrate that short periods of anoxia – euxinia, beginning at the Pliensbachian – Toarcian boundary, preceded the T-OAE. This is consistent with thallium isotope data from western Canada that indicates that the onset of global deoxygenation of ocean water occurred at this time (Them et al., Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018).

Samples through T-OAE Unit III into the base of the Bituminous Shales Subunit IVa yield FeHR/FeT values of >0.38 and generally ∼0.6 (Salem, Reference Salem2013; Houben et al., Reference Houben, Goldberg and Slomp2021), which point to persisting anoxic bottom-water conditions (Fig. 15). However, short-lived oxic intervals on a seasonal or annual scale would not be resolved by the current sampling resolution. Brief oxygenation events are indicated by petrographic evidence of fine-scale bioturbation in the laminated mudstones. On the other hand, it is striking that there is a total absence of macroscopic trace fossils throughout the T-OAE in the Cleveland Basin. Elsewhere in the Toarcian of Europe, there are usually at least a few bioturbated intervals, reflecting geologically brief oxygenation events, but there is no ichnological evidence of this in the Yorkshire succession.

Fepy/FeHR ratios of 0.7 to 0.8, lying at the threshold to euxinic conditions, indicate that H2S was present in bottom waters, albeit not permanently.

12.d. FeEF, MnEF and PEF

Profiles for Fe, Mn and P for Dove’s Nest and the Yorkshire coast successions plotted as enrichment factors show excellent agreement between the sections (Fig. 14). FeEF and MnEF display highly coherent profiles that indicate common mineralogical and environmental controls on their distributions. The two elements are strongly associated by their co-occurrence in Fe–Mn oxyhydroxides, carbonates (calcite, dolomite and siderite) and berthierine (e.g. Aggett, Reference Aggett1990). Iron has an additional association with pyrite, and this is reflected by the relatively elevated values of the FeEF relative to the MnEF profile through Subunit IIIb, the interval of high DOPT values indicative of high pyrite contents.

Highest FeEF and MnEF values occur in (1) ferruginous bands of the Staithes Sandstone where Fe–Mn oxyhydroxides (e.g. goethite) are likely the dominant minerals; (2) ironstones in the Cleveland Ironstone associated principally with high siderite and low phyllosilicate contents; (3) the Sulphur Bands of the Grey Shale (siderite); (4) concretionary carbonates (calcite, dolomite) in the Jet Rock – the Whale Stones and Top Jet Dogger; (5) concretionary calcitic (Peak Stones, Ovatum Band) or sideritic (‘beds’ 46, 50) levels in the Bituminous Shales and Hard Shales.

With the exception of the beds noted above, FeEF values generally fluctuate around average shale values through Units I, II and IV, are consistently enriched relative to PASS in Unit III (average 1.5), reflecting the addition of authigenic pyrite and show a plateau of low FeEF values (average 0.75) through Unit V (Fig. 14). By contrast, Mn is significantly depleted throughout the succession (MnEF ∼0.2 – 0.3) and MnEF values of >1 occur only in ironstones and concretionary carbonates. The generally low Mn values are consistent with the presence of oxygen-depleted bottom waters that prevented the deposition, redox cycling and concentration of Mn in the sediments (Calvert & Pedersen, Reference Calvert and Pedersen1996) which retained Fe in carbonates, silicates and sulfides. The carbonaceous mudstones correspond to Thermodynamic Zone III (low Mn, high Fe, low V, sulfidic) facies of Quinby-Hunt & Wilde (Reference Quinby-Hunt and Wilde1994), typically associated with anoxic bottom waters.

The PEF profiles from Dove’s Nest and the Yorkshire coast show excellent agreement (Fig. 14). A cyclic pattern is particularly apparent in the ultra-high-resolution P data of Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) from the coast, with regular PEF peaks of >2 and intervening troughs with a typical minimum of <0.6. These cycles coincide with opposing shifts in the TOC/PT profile that correspond to a transition to oxic bottom waters driving P enrichment and periods of anoxic bottom water leading to P depletion. A matching pattern is observed in the lower resolution data from Dove’s Nest. This is consistent with the differing processes and rates of organic carbon and phosphorus remineralization under varying redox conditions (Section 12a) and evidences a cyclic pattern of bottom-water oxygenation in the Cleveland Basin during the early Toarcian.

Five PEF cycles are recognized between the bases of the D. clevelandicum († symbol in Fig. 14) and H. falciferum subzones dated at 183.94 Ma and 182.06 Ma (Gradstein et al., Reference Gradstein, Ogg, Schmitz and Ogg2020), an interval of 1.88 Ma. This yields a periodicity approaching the Milankovitch ∼405 ka long eccentricity cycle. Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) postulated a ∼405 ka eccentricity cycle for detrital input in the D. tenuicostatum – C. exaratum subzones (see Fig. 4) in the Cleveland Basin based on spectral analysis of Zr/Al data from the coast (Hawsker Bottoms), but these do not show a consistent relationship to the PEF cycles (compare Figs 4 and 14). By contrast, taking the base D. tenuicostatum and H. bifrons zone ages of 184.20 Ma and 181.17 Ma (3.03 Ma; Gradstein et al., Reference Gradstein, Ogg, Schmitz and Ogg2020) with 13 PEF cycles (Fig. 14) yields a 250 ka periodicity. Applying other subzone ages generates a similar range of values. There is, therefore, no evidence for a consistent periodicity with potential orbital forcing of the PEF cycles, but it is acknowledged that changes in sedimentation rate and/or hiatuses in the section (e.g. Suan et al., Reference Suan, Pittet, Bour, Mattioli, Duarte and Mailliot2008b; McArthur et al., Reference McArthur, Steuber, Page and Landman2016) make spectral analysis challenging (Boulila et al., Reference Boulila, Galbrun, Huret, Hinnov, Rouget, Gardin and Bartolini2014, Reference Boulila, Galbrun, Sadki, Gardin and Bartolini2019; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018).

13. Chemostratigraphic correlation, TOC and redox-sensitive trace metals

TOC, Mo, U and V profiles for the Dove’s Nest core are correlated to a compilation of data (Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; McArthur, Reference McArthur2019; Remírez & Algeo, Reference Remírez and Algeo2020) for the equivalent Yorkshire coastal succession in Figure 16. Molybdenum content (Mo), enrichment factors (MoEF) and Mo/TOC ratios, together with uranium (UEF) and vanadium (VEF) enrichment factors, clearly differentiate intervals of substantial redox-driven transition metal enrichment that complement interpretations based on Fe speciation (Fig. 15, Section 12.c). The lower resolution of the Dove’s Nest trace-metal profiles (Fig. 16) limits comparison to the ultra-high-resolution data available for the lower Toarcian of the coastal sections (Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) but there is good correspondence between the Mo, U and V curves of the two successions.

Figure 16. Correlation of Mo, TOC, V and U between the Dove’s Nest core and Yorkshire coastal outcrop sections. Dove’s Nest data this study. Yorkshire coast Mo and Mo/TOC and TOC plots from McArthur (Reference McArthur2019) with additional high-resolution TOC (thin black line, see Fig. 2), Mo and Mo/TOC curves (thin dark red lines; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018). Yorkshire VEF data were calculated from Ruvalcaba Baroni et al. (Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018) with MoEF, U and UEF data after Remírez & Algeo (Reference Remírez and Algeo2020). Stratigraphy as in Figures 2, 4. WS = Whale Stones (‘bed’ 35’); TJD = Top Jet Dogger (‘bed’ 39). Enrichment factors (Section 6.d), e.g. MoEF, are calculated relative to PASS. Vertical grey dotted lines are EF values of 1. Vertical green dotted lines indicate the ‘intermittent euxinia’ boundary of 25 ppm Mo (Scott & Lyons, Reference Scott and Lyons2012) and the ‘anoxic threshold’ of TOC = 2.5 wt% (Algeo & Maynard, Reference Algeo and Maynard2004). Consistent enrichment in authigenic uranium (UEF >1) characterizing Units III and IV is also well displayed in spectral gamma-ray logs of the coastal sections (Myers & Wignall, Reference Myers, Wignall, Leggett and Zuffa1987; Parkinson, Reference Parkinson1996).

The high-resolution structure and values of the Toarcian TOC curves are well correlated between the Dove’s Nest core and the coastal succession (Figs 2, 15, 16). The ‘anoxic threshold’ (oxic – anoxic non-sulfidic bottom-water boundary) of 2.5% TOC (Algeo & Maynard, Reference Algeo and Maynard2004) is initially exceeded briefly at the level of the three Sulphur Bands in the lower Grey Shale (Unit II). Peak TOC values then characterize T-OAE Unit III with an interval at the top of Subunit IIb documented in the higher resolution data from the coast that includes values of >10%, indicative of euxinic conditions (‘euxinic threshold’).

The interval of the TOC maximum (summit of Subunit IIIb) displays the highest UEF and VEF values in the lower Toarcian (Fig. 16), peaking in the Whale Stones ‘bed’ 35 at 5.4 and 2.8, respectively in the coast data and both reaching >7 at Dove’s Nest. A return to anoxia is indicated at the top of Unit III and through Unit IV with lower TOC, U and V contents, then generally dysoxic bottom waters are interpreted for Unit V with TOC values around or below the anoxic threshold. Pulses of renewed anoxia at the bottom and top of Subunit Vb are documented in the Hard Shales at the base of the middle Toarcian.

Factors controlling the TOC content of sediments are complex (Tyson, Reference Tyson and Harris2005) and TOC alone cannot be regarded as a definitive indicator of bottom-water redox, since similar TOC contents may occur in sediments deposited under fully oxygenated to euxinic water columns (e.g. Canfield, Reference Canfield1994; Hartnett et al., Reference Hartnett, Keil, Hedges and Devol1998; Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b). Nonetheless, the application of a 2.5% TOCWR cut-off to define the anoxic threshold coincides exactly with the disappearance of trace fossils in the Cleveland Basin succession (cf. Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023), which only reappear when bulk-rock contents fall consistently below that value (Section 19.a). Palaeontological evidence therefore fully supports the use of this value as an indicator of seafloor anoxia in the study sections.

Coincident TOC, Mo and U enrichments are generally considered to be strong indicators of deposition in oxygen-depleted bottom waters (e.g. Emerson & Huested, Reference Emerson and Huested1991; Calvert & Pedersen, Reference Calvert and Pedersen1993; Algeo & Tribovillard, Reference Algeo and Tribovillard2009; Scott & Lyons, Reference Scott and Lyons2012; Tribovillard et al., Reference Tribovillard, Algeo, Baudin and Riboulleau2012; Algeo & Liu, Reference Algeo and Liu2020; Fernández-Martínez et al., Reference Fernández-Martínez, Martínez Ruiz, Rodríquez-Tovar, Piñuela, García-Ramos and Algeo2023; Paul et al., Reference Paul, van Helmond, Slomp, Jokinen, Virtasalo, Filipsson and Jilbert2023) with individual enrichment factors providing information on bottom-water restriction and redox conditions, ranging from oxic through anoxic to euxinic. The current understanding of the geochemical behaviour of Mo and U in seawater and sediments has been reviewed by Paul et al. (Reference Paul, van Helmond, Slomp, Jokinen, Virtasalo, Filipsson and Jilbert2023).

Vanadium has also been advocated as a good redox proxy (e.g. Emerson & Huested, Reference Emerson and Huested1991; Quinby-Hunt & Wilde, Reference Quinby-Hunt and Wilde1994; Algeo & Maynard, Reference Algeo and Maynard2004; Piper & Calvert, Reference Piper and Calvert2009; Scholz, Reference Scholz2018; Algeo & Liu, Reference Algeo and Liu2020; Bennett & Canfield, Reference Bennett and Canfield2020). Unlike Mo and U, vanadium is present in high concentrations in the phyllosilicate fraction of sediments, as indicated by its high value in average marine shale (140 ppm vs 1 ppm Mo and 3 ppm U; Taylor & McLennan, Reference Taylor, McLennan and Meyers2001), so bulk-rock values are less sensitive to redox change.

Chemostratigraphic profiles for U and V show similar trends for the lower Toarcian that differ substantially from the Mo curves (Figs 5, 16). Despite their different stratigraphic trends, Mo and U are both positively correlated with TOC at Dove’s Nest (r ≥ 0.4, Table S3) and the three constituents show a close association on the PCA plot (Fig. 9) with negative PC1 and PC2 scores (Table S4) driven by samples from Units III and IV, the T-OAE and lower Bituminous Shales intervals. Some samples of the Hard Shales in Unit V fall in the same sector. Environmental factors responsible for the differing behaviour of Mo, U and V may be examined further using bivariate geochemical plots and knowledge of their distribution in sediments from modern ocean basins. TOC vs trace metal cross-plots of Yorkshire coast data have been used previously to interpret bottom-water redox conditions during the deposition of Cleveland Basin Toarcian mudstones (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Pearce et al., Reference Pearce, Cohen, Coe and Burton2008; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; McArthur, Reference McArthur2019; Remirez & Algeo, Reference Remirez and Algeo2020).

13.a. Molybdenum

A plot of Mo vs Al of upper Pliensbachian – middle Toarcian samples from Dove’s Nest and published data from the Yorkshire coastal succession (Fig. 17a) illustrates the significant enrichment in Mo (> 10 – 60 ppm) that characterizes the lower Bituminous Shales of Unit IV and to a lesser extent (> 3 ppm) the upper Grey Shale – Jet Rock interval of T-OAE Unit III, compared to other parts of the succession (≤ 2 ppm) and to average shale (1 ppm). Elevated Mo values also occur in the organic-rich intervals of the Sulphur Bands (SB1 – 3, 3 – 25 ppm) at the base of Subunit IIa, and at the bottom and top of the Hard Shales of Subunit Vb (up to 22 ppm). The Mo vs Al plot shows the lack of a clear correlation between the two elements, suggesting no significant association of Mo with the aluminosilicate fraction of the rocks. The Pearson correlation and PCA (Tables S3, S4) results both display a weak negative correlation (r = -0.3) between the two constituents.

Figure 17. Cross-plots for key redox-sensitive trace metals. (a) Mo vs Al. Mo shows no clear relationship with Al. (b) Mo vs TOC. The steep upper regression line (small grey dots) for Units IV and V (left) derived from the Dove’s Nest data, displaying a positive correlation between Mo and TOC, contrasts to the shallow lines (small grey dots) derived for T-OAE Unit III (right – upper line and statistics is for Dove’s Nest samples, lower line is for Yorkshire coast samples of McArthur et al. Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008). Regression lines from selected modern anoxic silled basins representing increasing deep water renewal times of <10 – 650 ka (Algeo & Rowe, Reference Algeo and Rowe2012) are Mo/TOC (ppm/%) ∼45 Saanich Inlet (purple); ∼25 Cariaco Basin (green); ∼9 Framvaren Fjord (red); ∼4.5 Black Sea (blue). Bottom water restriction trends after Algeo & Lyons (Reference Algeo and Lyons2006). (c) U vs Al. Lower regression line is for Pliensbachian Subunits Ia–c; upper line is for Toarcian Units IV and V (all samples ≥3 ppm U). (d) U vs TOC. Lower regression line (left) for Subunits Ia–c; upper regression line (right) for Units IV and V. (e) V vs Al. Regression line is for Pliensbachian Subunits Ia–c, excluding the three Fe-rich flyers. (f) V vs TOC. Regression lines are for Pliensbachian Subunits Ia–c (left, Dove’s Nest) and Toarcian Units II–V (right, coast samples; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018). TOC-based anoxic (2.5%, vertical dashed green line marking boundary between oxic and anoxic non-sulfidic conditions) and euxinic (10%, upper limit of x-axis) thresholds after Algeo & Maynard (Reference Algeo and Maynard2004). Solid symbols are from the Dove’s Nest core (this study), faded symbols are for Yorkshire coastal outcrop samples (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; McArthur, Reference McArthur2019; Remírez & Algeo, Reference Remírez and Algeo2020). Dove’s Nest TOC data are whole-rock values. Average shale (PASS) composition after Taylor & McLennan (Reference Taylor, McLennan and Meyers2001). Note that samples with 10 – 20% TOC reported from Unit III of the coastal outcrops fall outside the plot area of (B), (D) and (F) but lie on the trends of the regression lines shown.

The Mo vs TOC plot (Fig. 17b) illustrates that high Mo contents are associated with TOC enrichment, with Mo values increasing substantially in samples containing ≥2.5% TOC, the ‘anoxic threshold’ of Algeo & Maynard (Reference Algeo and Maynard2004), but with substantial scatter. Relatively few samples exceed the ‘intermittent euxinia’ threshold of 25 ppm Mo proposed by Scott & Lyons (Reference Scott and Lyons2012), all of which occur in Unit IV (Fig. 16). Positive correlations between Mo and TOC are observed for Unit III and IV sample suites but with very different slopes. In addition to the data plotted in Figure 17, samples with 10 – 20% TOC (exceeding the threshold for euxinic bottom waters) have been recorded from Unit III of the coastal outcrop (Fig. 2). These data fall outside the plot area of Figure 17a, d and f but lie on the trend of the Unit III lower regression line shown in Figure 17b.

The redox behaviour of Mo in marine environments has been reviewed by Tribovillard et al. (Reference Tribovillard, Algeo, Lyons and Riboulleau2006), Algeo & Tribovillard (Reference Algeo and Tribovillard2009), Scott & Lyons (Reference Scott and Lyons2012), Hardisty et al. (Reference Hardisty, Lyons, Riedinger, Isson, Owens, Aller, Rye, Planavsky, Reinhard, Gill, Masterson, Asael and Johnston2018), Bennett & Canfield (Reference Bennett and Canfield2020), Them et al. (Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022) and Paul et al. (Reference Paul, van Helmond, Slomp, Jokinen, Virtasalo, Filipsson and Jilbert2023), amongst others. Molybdenum is present in seawater principally in the form of molybdate (MoO42−). Molybdenum is not concentrated by plankton and displays little affinity for the surfaces of clay minerals, CaCO3 and, generally, Fe-oxyhydroxides (but see Scholz et al., Reference Scholz, Siebert, Dale and Frank2017) at marine pH values but may be captured by Mn-oxyhydroxides, typically at the sediment surface. Molybdenum accumulation in these deposits is very slow, but the overwhelming dominance of oxic environments in the modern oceans results in an estimated 35 – 50% of Mo being removed via adsorption onto Mn-oxyhydroxides.

Reduction of oxyhydroxides during diagenesis in organic-rich sediments liberates adsorbed Mo to pore waters. Molybdenum fixation is believed to take place principally by reduction to thiomolybdate complexes (MoOxS4−x , x = 0 – 3) in the presence of free H2S and subsequent sequestration in organic matter and/or iron sulfides. Molybdenum fixation may also occur directly in a euxinic water column. Here, Mo is scavenged from seawater by settling the organic matter and sulfide particles, leading to substantial Mo enrichment (reaching 100 ppm) in the underlying organic-rich sediment. It is estimated that 50 – 65% of Mo is removed in the oceans via these processes today because even though only a very small proportion (≪1%) of the modern seafloor is overlain by euxinic bottom waters, burial efficiency is very high.

McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008) argued that the development of weak hydrographic restriction in the Cleveland Basin during the early H. falciferum Subzone (lower Bituminous Shales of Unit IV) was responsible for substantial Mo enrichment from sulfidic bottom waters. However, samples from Unit III (T-OAE) also display a significant positive correlation with TOC yet these show up to two orders of magnitude less enrichment. By analogy to modern anoxic basin environments (Algeo & Lyons, Reference Algeo and Lyons2006; Algeo & Rowe, Reference Algeo and Rowe2012), the low Mo values of the T-OAE black shales (Unit III) relative to their high TOC contents have been previously interpreted to reflect the presence of an extremely restricted water mass in the Cleveland Basin during the T-OAE (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; McArthur, Reference McArthur2019; Remírez & Algeo, Reference Remírez and Algeo2020), with local Mo drawdown limiting uptake of the element into the sediment.

The impact of water mass restriction on the Mo/TOC (ppm/%) ratio of sediments in modern basins may be illustrated by data from weakly restricted (e.g. Saanich Inlet, ∼45) to very restricted (e.g. Black Sea, ∼4.5) silled basins (Fig. 17). It should be noted, however, that a low Mo/TOC ratio of ∼6 also characterizes sediments from strong upwelling systems, such as modern offshore Namibia (Algeo & Lyons, Reference Algeo and Lyons2006). Here, very high productivity induced by coastal upwelling leads to a high rate of TOC deposition that overwhelms the supply of Mo from upwelling water.

TOC whole-rock concentrations determined in the Dove’s Nest samples all fall below 10% (Fig. 17b) but values exceeding 15% have been recorded from the upper beds of Subunit IIIb on the Yorkshire coast (Figs 2, 7; Kemp et al., Reference Kemp, Coe, Cohen and Weedon2011; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; McArthur, Reference McArthur2019; Houben et al., Reference Houben, Goldberg and Slomp2021). The coast data form an array that continues the trend of the shallowest regression line plotted in Figure 17b, interpreted by McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008) to represent the most extreme water mass restriction associated with the highest TOC deposits. However, Cleveland Basin T-OAE samples define trends that show considerably greater Mo depletion (Mo/TOC 0.4 – 2, Fig. 17) than the modern Black Sea.

Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) argued that comparable trends in Mo across the T-OAE interval in both Yorkshire and the Paris Basin (Hermoso et al., Reference Hermoso, Minoletti and Pellenard2013) suggest a similar oceanic drawdown of this element accompanying widespread anoxia in both basins. Furthermore, clear evidence of a transgressive trend in the Cleveland Basin at the time of maximum Mo drawdown (Figs 10, 16) conflicts with the basin restriction model for the euxinic conditions that characterize the T-OAE interval. Data from a deep Panthalassic Ocean site in Japan (Inuyama area; Kemp et al., Reference Kemp, Chen, Cho, Algeo, Shen and Ikeda2022a) also evidences drawdown of Mo and U accompanying the T-OAE. Here also, a marked expansion of anoxic and possibly euxinic conditions to the seafloor occurred during both the PlToBE and the T-OAE, accompanied by increased organic carbon burial, but Mo drawdown cannot be related to hydrographic restriction.

Them et al. (Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022) documented comparable low Mo/TOC values (1 – 2) through the T-OAE intervals in a transect of 3 sites from the Western Canada Sedimentary Basin, an area that is interpreted to have maintained strong connections with Panthalassa open-ocean deep waters throughout the Early Jurassic. By contrast to the Cleveland Basin and Japan sites, this area experienced relatively stable anoxic and euxinic conditions throughout late Pliensbachian – middle Toarcian (Them et al., Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018), and there is no geological evidence to suggest that basin restriction played any role in the observed Mo/TOC trends. Sustained anoxic/euxinic conditions at a site are essential to successfully isolate global changes in the marine Mo reservoir from local factors.

The results of Them et al. (Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022) confirm previous work (Pearce et al., Reference Pearce, Cohen, Coe and Burton2008; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; Kemp et al., Reference Kemp, Chen, Cho, Algeo, Shen and Ikeda2022a) indicating that global drawdown of oceanic Mo accompanied the T-OAE. In the upper Pliensbachian, the average Mo/TOC value from the three Western Canada sites is ∼25 (range of ∼22 – 30), which is similar to the modern euxinic Cariaco Basin (∼25, Fig. 17; cf. Algeo & Lyons, Reference Algeo and Lyons2006), suggesting a large global marine Mo inventory similar to today. The average Mo/TOC value of ∼2 for anoxic/euxinic sediments deposited during the T-OAE in Western Canada is identical to that obtained for Dove’s Nest, with an even lower value of 0.4 being derived from the larger Yorkshire coast dataset (Fig. 17b). These very low values suggest additional local Mo drawdown in the Cleveland Basin from already highly depleted ocean water. Post-T-OAE Mo/TOC values in the Canadian sites typically range from 4 to 10, like those seen in the lower Bituminous Shales, Unit IV of Yorkshire, indicative of a return to higher Mo contents in seawater (and contraction in the global area of euxinic bottom waters), but at lower concentrations than those that preceded the T-OAE.

Local basin restriction with the development of highly euxinic bottom waters, coincident with a global drawdown of Mo, has been interpreted for the T-OAE interval in the Dutch Central Graben, West Netherlands Basin, SW German Basin and Paris Basin (Fernández-Martínez et al., Reference Fernández-Martínez, Martínez Ruiz, Rodríquez-Tovar, Piñuela, García-Ramos and Algeo2023). Other areas in northern Europe including the Cardigan Bay Basin, together with Tethyan and Panthalassa Ocean sites, remained unrestricted but with evidence of the presence of euxinic pore waters at the sediment/water interface at some localities. The small global marine Mo reservoir during the T-OAE in combination with variable basin restriction explains the large regional differences in sediment Mo (and inferred water mass) concentrations and isotopes documented from European basins.

13.b. Uranium

The U vs Al plot (Fig. 17c) illustrates a strong positive correlation (R2 = 0.73) between the two elements for the Pliensbachian samples of Subunits Ia – c and Toarcian Unit V that indicates an association between U and the aluminosilicate fraction (principally clay minerals and a heavy mineral suite) of the rocks, as generally seen in marine black shales (Swanson, Reference Swanson1961). Mudstones with the highest Al contents in the array, cluster around the average shale value of 3 ppm.

Low U concentrations of 2 – 4 ppm are generally found in mudstones (Swanson, Reference Swanson1961; Spirakis, Reference Spirakis1996) and U contents of 1 – 3 ppm are typical throughout the Earth’s crust (Hazen et al., Reference Hazen, Ewing and Sverjensky2009). Scatter towards lower values on the U vs Al plot (Fig. 17c) is driven principally by increasing proportions of U-poor carbonates (calcite, siderite) and quartz diluting the bulk-rock content. Anomalously high U contents, up to 6.5 ppm, characterize Units III and IV and these show a strong negative correlation (R2 = 0.73) with Al attributable to a positive correlation with TOC (Fig. 17d). The positive correlation between U vs Al and U vs TOC for Subunits Ia – c (Fig. 17c, d) contrasts to the absence of correlation displayed in the corresponding Mo plots. The low U and low TOC contents in the sideritic ironstone of the Avicula Seam are notable.

Uranium occurs in nature in two main redox states, U(IV) and U(VI). In oxygenated solutions, the hexavalent uranyl (UVIO22+) species forms highly soluble and non-reactive Ca and Mg complexes with UO2(CO3)34; the dominant aqueous U(VI) species in seawater is neutral [Ca2UO2(CO3)3](aq) (Endrizzi & Rao, Reference Endrizzi and Rao2014). Under dysoxic – anoxic conditions, U is reduced and forms insoluble tetravalent U compounds and is ultimately fixed in sediments principally as uraninite (UO2). This reduction, which is likely mediated by microorganisms, leads to preferential uptake of U by anoxic – euxinic sediments (Barnes & Cochran, Reference Barnes and Cochran1990; Klinkhammer & Palmer, Reference Klinkhammer and Palmer1991; Khaustova et al., Reference Khaustova, Tikhomirova, Korost, Poludetkina, Voropaev, Mironenko and Spasennykh2021). Unlike Mo, uranium reduction has not been observed in the marine water column but occurs predominantly in association with near-surface iron reduction in organic-rich sediments with diffusion of U into the sediment from overlying bottom waters, favoured by lowered sedimentation rates (Algeo & Maynard, Reference Algeo and Maynard2004).

Toarcian high-U samples plot consistently within the sulfidic field beyond the ‘anoxic threshold’ of Algeo & Maynard (Reference Algeo and Maynard2004) on the U vs TOC plot (Fig. 17d), evidencing the formation of authigenic U under anoxic conditions during sedimentation, as uranium was removed from solution. However, the mechanisms involved in this removal are complex (Cumberland et al., Reference Cumberland, Douglas, Grice and Moreau2016) and the phase partitioning in the sediment remains uncertain. Unit III samples of the T-OAE display substantial scatter due in part to variable dilution by carbonates in the concretion beds of the Jet Rock (Fig. 17d), while Unit IV samples from the lower Bituminous Shales define a tighter array with a steeper positive correlation between U and TOC than that defined by Subunits Ia – c.

The offset to lower U/TOC ratios for Unit III T-OAE samples compared to lower Bituminous Shales Unit IV is consistent with global drawdown of U accompanying the T-OAE, as discussed above for Mo (Section 13.a) and interpreted for other episodes of expanded anoxic seafloor area accompanying OAEs, based on uranium isotope studies (e.g. Montoya-Pino et al., Reference Montoya-Pino, Weyer, Anbar, Pross, Oschmann, van de Schootbrugge and Arz2010; Clarkson et al., Reference Clarkson, Stirling, Jenkyns, Dickson, Porcelli, Moy, Pogge von Strandman, Cooke and Lenton2018).

13.c. Vanadium

Pliensbachian Subunits Ia – c samples display a significant positive correlation between V and Al (Fig. 17e) with some Unit III and most Unit IV samples lying on the same trend. By contrast, Unit V samples form a distinct tight cluster that lies below the line. This offset reflects the lower V/Al ratio that differentiates Unit V from the older beds (Figs 5, 8). Samples of T-OAE Unit III from the Yorkshire coast form a linear array above the line with a negative trend between V and Al, attributable to the diluting effect of the high TOC content in V-rich samples.

Three Fe-rich samples display anomalous high V contents of up to 521 ppm (sample 195.18, Fig. 17e). Moderate enrichment of V in sideritic ironstones (average 114 ppm) from the Cleveland Ironstone was noted by Gad (Reference Gad1966), with distinctive higher values (average 560 ppm) recorded in berthierine (chamosite)-bearing samples. Vanadium enrichment in ironstones is well displayed in the V/Al stratigraphic profile for Dove’s Nest (Fig. 5). The highest V samples also display anomalous high Th contents (Fig. S2) as observed elsewhere in berthierine (Yang et al., Reference Yang, Shen, Qin, Jin, Zhang, Tong and Liu2021), consistent with the presence of the mineral in these ironstones. Thorium enrichment in these beds is responsible for their distinctive low U/Th ratio on the trace-element ratio profile (Fig. 5).

Positive correlations are observed between V and TOC with a steeper regression line generated by Subunit Ia – c samples compared to Units II – V (Fig. 17f). This steepening may be attributed to the additional presence of Fe-mineral (berthierine/chamosite, siderite and goethite) associated V and the greater influence of quartz and feldspar dilution effects in the coarser grained, TOC-poor, oxic facies of the Pliensbachian, compared to predominantly phyllosilicate and TOC-associated V in the anoxic – euxinic Toarcian mudstones. The Toarcian sample V vs TOC regression line, derived from the V data from the Yorkshire coast of Ruvalcaba Baroni et al. (Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018), has an intercept within error of average shale (PAAS = 140 ppm). Unit IV samples of the lower Bituminous Shales cluster above the Units II – V regression line indicate elevated V/TOC ratios compared to the full Toarcian sample suite, like, but less pronounced than, the offsets seen for Mo/TOC and U/TOC (Fig. 17b, d; also compare Figs 5 and S3).

Vanadium behaves as a conservative or near-conservative element in the modern oceans (Wu et al., Reference Wu, Owens, Huang, Sarafian, Huang, Sen, Horner, Blusztajn, Mprton and Nielsen2019). The early diagenetic behaviour of V in marine sedimentary environments is poorly studied. In oxic environments, it is considered that V is adsorbed on particulate organic matter and Fe–Mn oxyhydroxides, transferred from the water column to the sediment and released to pore waters during Mn and Fe reduction (e.g. Bennett & Canfield, Reference Bennett and Canfield2020). Dissolved V is reduced to the strongly adsorbed VO(OH)3 under anoxic (= suboxic) conditions and is retained by the sediment during burial. Free sulfide in close proximity to the Mn and Fe reduction zone likely favours the reduction and capture of V from the pore water as authigenic V(III) hydroxide. Under anoxic bottom waters, a hydrogenous fraction may be added to the sediment directly through inorganic redox reactions. The element is therefore commonly enriched in carbonaceous mudstones where it is initially bound to organic compounds but may also partition into clay minerals (Breit & Wanty, Reference Breit and Wanty1991; Tribovillard et al., Reference Tribovillard, Algeo, Lyons and Riboulleau2006).

The largest V enrichments are observed in modern sediments being deposited within the core of perennial oxygen-minimum zones on continental margins. Hydrographically restricted euxinic basins typically do not show high levels of V enrichment due to low rates of sedimentation limiting the particulate V flux and low rates of deep-water renewal limiting hydrogenous V uptake (Scholz et al., Reference Scholz, Siebert, Dale and Frank2017; Bennett & Canfield, Reference Bennett and Canfield2020). Importantly, V drawdown under reducing conditions does not require the presence of sulfide in the water column.

A study of Late Cretaceous (Cenomanian – Turonian) OAE2 indicates that drawdown of V accompanied the global expansion of oxygen-deficient but non-sulfidic waters, fingerprinting an expansion of anoxia prior to the development of euxinia evidence by subsequent drawdown of Mo (Owens et al., Reference Owens, Reinhard, Rohrssen, Love and Lyons2016). The offset to lower V/TOC ratios that characterize Unit III compared to higher values in Unit IV (Fig. 17f) is consistent with a drawdown and reduced inventory of V during the period of maximum oxygen depletion accompanying the T-OAE, although the additional impact of basin restriction cannot be excluded.

13.d. Molybdenum – Uranium covariation

Algeo & Tribovillard (Reference Algeo and Tribovillard2009) and Tribovillard et al. (Reference Tribovillard, Algeo, Baudin and Riboulleau2012) advocated the use of MoEF vs UEF covariation diagrams for marine sediments and rocks to interpret bottom-water redox conditions, the operation of metal-oxyhydroxide particulate shuttles in the water column, and the degree of water mass restriction driving the evolution of water chemistry.

Molybdenum and U exhibit conservative behaviour under oxic conditions but show progressive uptake by the sediment from seawater where the bottom water mass and/or surface sediments are anoxic (Algeo & Tribovillard, Reference Algeo and Tribovillard2009; Tribovillard et al., Reference Tribovillard, Algeo, Baudin and Riboulleau2012). Aqueous Mo and U are both removed from seawater across the sediment–water interface in reducing facies, but Mo requires a lower redox potential for sedimentary uptake than U: uptake of U begins at the Fe(II)–Fe(III) redox boundary and therefore exceeds that of Mo when redox conditions at the sediment/water interface are anoxic, while authigenic Mo enrichment requires the presence of H2S and is favoured over U when benthic redox conditions are euxinic.

In unrestricted open-ocean environments such as continental margin upwelling systems, typified by those of the modern eastern tropical Pacific, it is argued that a general shift from oxic through anoxic to sulfidic benthic redox conditions enhances sediment uptake of both Mo and U (Fig. 17) by diffusion downwards across the sediment/seawater interface, but the change in redox state drives a progressive increase in Mo/U ratios from 0.1 – 0.3 to 1 – 3 times the seawater molar ratio (Algeo & Tribovillard, Reference Algeo and Tribovillard2009).

In moderately restricted basins, such as the modern Cariaco Basin, with a deep or highly variable chemocline, the operation of a particulate Fe–Mn oxyhydroxide particulate shuttle may enhance the transfer of aqueous Mo to the sediment/water interface. Molybdenum is vigorously scavenged from seawater by Fe–Mn oxyhydroxide particles in the oxic water column. On reaching the sediment/water interface, these particles are reductively dissolved, releasing Mo to diffuse back into the bottom water or available to be scavenged by other phases within the sediment. Aqueous U is unaffected by this process, resulting in elevated Mo/U ratios that are typically 3 to 10 times those of seawater (Fig. 18).

Figure 18. MoEF vs UEF cross-plot for stratigraphic units comprising the upper Pliensbachian – middle Toarcian of Dove’s Nest core. The three diagonal lines represent multiples (0.3, 1, 3) of the Mo:U ratio of present-day seawater (SW) converted to an average weight ratio of 3.1 for the purpose of comparison with sediment Mo:U weight ratios (Tribovillard et al., Reference Tribovillard, Algeo, Baudin and Riboulleau2012). General patterns of MoEF vs UEF covariation in modern marine environments modified from Tribovillard et al. (Reference Tribovillard, Algeo, Baudin and Riboulleau2012) and Yano et al. (Reference Yano, Yasukawa, Nakamura, Ikehara and Kato2020): unrestricted open ocean field based on the eastern tropical Pacific; particulate shuttle field based on the Cariaco Basin and Saanich Inlet. Trend lines show deoxygenation trends in modern marine environments (Tribovillard et al., Reference Tribovillard, Algeo, Baudin and Riboulleau2012) with positions based on data from restricted basins and coastal settings (Paul et al., Reference Paul, van Helmond, Slomp, Jokinen, Virtasalo, Filipsson and Jilbert2023). The anomalous high EF values of Whale Stones sample 167.89, interpreted as sampling a carbonate concretion, are likely an artifact of the high carbonate content (81%).

In more restricted basins like the Black Sea, which are characterized by a shallow and stable chemocline, relative rates of authigenic Mo–U enrichment will depend on the degree of evolution of the water chemistry in the deep water mass. Initially, when the aqueous Mo/U molar ratio is close to that of seawater, sediment Mo/U ratios will be high (Fig. 18), but if aqueous Mo is depleted through sedimentation without compensatory resupply, the sediment Mo/U ratios will decline.

More recently, sequestration of Mo and U to particulate matter has been shown to occur at the seafloor in some open-ocean upwelling environments, exemplified by the Benguela upwelling system off Namibia, due to the intermittent presence of sulfide, either in bottom water or in pore water immediately at the sediment/water interface (He et al., Reference He, Clarkson, Andersen, Archer, Sweere, Kraal, Guthauser, Huang and Vance2021). Mo and U are sequestered to the sediment during temporarily more reducing conditions. They subsequently undergo oxidative dissolution under less reducing conditions and diffuse away from local Mo and U pore-water concentration peaks. Mo and U are taken up again into reduced authigenic phases (sulfide-related) at depth, setting the final Mo and U content and isotope characteristics of the sediment. In this way, early diagenetic enrichment of Mo and U may be governed by temporal redox fluctuations.

He et al. (Reference He, Clarkson, Andersen, Archer, Sweere, Kraal, Guthauser, Huang and Vance2021) have argued that early diagenetic processes in upwelling environments can produce patterns similar to those observed for coupled Mo–U covariation and isotope systematics in restricted and semi-restricted basins (see above) but via a different set of processes. Unambiguous identification of a restricted basin system therefore requires critical assessment of multiple proxies.

Samples from the upper Pliensbachian – lower Toarcian succession at Dove’s Nest fall within 3 different areas on a MoEF vs UEF cross-plot (Fig. 18). Samples from Units I (upper Pliensbachian), II (basal lower Toarcian) and V (upper lower to middle Toarcian) form a broad array with UEF values close to or less than PAAS (i.e., no U enrichment) but MoEF values range from ∼1 to ∼10 (moderate Mo enrichment). The pattern of U–Mo covariation does not fall along the established unrestricted open ocean trend but rises upward towards the particulate shuttle field. Significant outliers include Avicula Seam sample 200.39 which exhibits greater Mo and U enrichment than other samples from the Cleveland Ironstone (Fig. 18), possibly an artefact of a very low Al2O3 content (2.9%) and Sulphur Band 3 sample 180.91, which displays U enrichment typical of anoxic unrestricted water masses. Samples from the Hard Shales at the base of Unit V are characterized by variable high MoEF and UEF values.

Following the approach of Algeo & Tribovillard (Reference Algeo and Tribovillard2009) and Tribovillard et al. (Reference Tribovillard, Algeo, Baudin and Riboulleau2012), Yorkshire Pliensbachian samples lie within a poorly defined field of dominantly oxic – anoxic conditions (cf. Fig. 17) but with the levels of Mo enrichment potentially indicating periods of bottom water restriction and/or uptake under sulfidic conditions in the near-surface sediment. However, interpretation of MoEF values in these beds must be treated with caution since Mo concentrations in many samples were below the limit of quantification (2 ppm) for the Dove’s Nest study (Table S2) and were allocated a nominal value of 1 ppm (=PAAS) for plotting purposes, equivalent to the average Mo value (1.2 ± 1.0 ppm) obtained for samples from this part of the succession on the coast by McArthur (Reference McArthur2019).

It is observed that most Pliensbachian samples, together with those from Toarcian Units II and V, exhibit UEF values of <1 (Fig. 18). The Lower Jurassic in the central Cleveland Basin has been buried into the oil window (∼95° C) (Barnard & Cooper, Reference Barnard, Cooper and Brooks1983; Salem, Reference Salem2013; French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014; Song et al., Reference Song, Littke, Weniger, Ostertag-Henning and Nelskamp2015; Song et al., Reference Song, Littke and Weniger2017), and low UEF values have previously been attributed to U loss accompanying hydrocarbon generation and migration (Wignall in Tribovillard et al., Reference Tribovillard, Algeo, Baudin and Riboulleau2012). However, a study of lower Toarcian black shales from the Alum Shale Formation of Sweden indicates that U remains immobile during catagenesis (Lecomte et al., Reference Lecomte, Cathelineau, Michels, Peiffert and Brouand2017), while U/TOC ratios increase with maturity in the lower Toarcian Posidonia Shale of the Lower Saxony Basin, northern Germany (Dickson et al., Reference Dickson, Idiz, Porcelli, Murphy, Celestino, Jenkyns, Poulton, Hesselbo, Hooker, Ruhl and van der Boorn2022b), due to loss of organic matter with a low metal content during secondary migration. Low UEF values in the Cleveland Basin are therefore judged to be a primary feature.

Samples from T-OAE Unit III, the top Grey Shale and Jet Rock, form an extended array of high UEF (∼1 – ∼2) and high MoEF (∼4 – ∼13) values with a relatively constant U/Mo ratio that parallels the seawater trend (Fig. 18). A sample from the Top Jet Dogger (164.91 m) falls within the particulate shuttle field, another from the Whale Stones (167.89 m) displays very high MoEF and UEF values of 42 and 7.6, respectively, associated with the high carbonate content (81%) in this concretion sample. Unit II samples generally fall on the lower segment of the Black Sea trend of Yano et al. (Reference Yano, Yasukawa, Nakamura, Ikehara and Kato2020).

High MoEF values of ∼10 – ∼35 in lower Bituminous Shales Unit IV are accompanied by moderate U enrichment (UEF ∼1 – ∼2) and Mo/U ratios that exceed 3× modern seawater. These values imply the operation of a particulate shuttle (Fig. 18), although one that was less active than that in the modern Saanich Inlet and Cariaco Basin (compare Yano et al., Reference Yano, Yasukawa, Nakamura, Ikehara and Kato2020, fig. 6) where, in the latter, MoEF and UEF reach 200 and >8, respectively. Lower MoEF values paired with Mo/U ratios of ∼1 – 2× seawater in Unit III have been interpreted to reflect the drawdown of aqueous Mo as a consequence of transfer to the sediment in a highly restricted Cleveland Basin water mass during the T-OAE (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008), the ‘basin reservoir effect’ (Algeo, Reference Algeo2004; Algeo & Lyons, Reference Algeo and Lyons2006). Elevated UEF values are consistent with the development of more reducing conditions in bottom waters.

Restriction of Cleveland Basin bottom waters has been linked to the development of a ‘deep’ pycnocline (a total water depth of 50 – 100 m was proposed for the Mulgrave Shale by Hallam, Reference Hallam1997) caused by increased freshwater runoff generating low-salinity surface waters that limited mixing between surface and bottom waters and exchange with the adjacent seas and open ocean (Küspert, Reference Küspert, Einsele and Seilacher1982; Sælen et al., Reference Sælen, Doyle and Talbot1996, Reference Sælen, Tyson, Talbot and Telnaes1998, Reference Sælen, Tyson, Telnaes and Talbot2000; Bailey et al., Reference Bailey, Rosenthal, McArthur, van de Schootbrugge and Thirlwall2003; van de Schootbrugge et al., Reference van de Schootbrugge, McArthur, Bailey, Rosenthal, Wright and Miller2005; Wignall et al., Reference Wignall, Newton and Little2005; McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; McArthur, Reference McArthur2019; Remírez & Algeo, Reference Remírez and Algeo2020). However, the amplitude of any salinity decrease is hotly contested (Hesselbo et al., Reference Hesselbo, Little, Ruhl, Thibault and Ullmann2020a) and palaeontological evidence and salinity proxies derived from the organic geochemical data (Section 15.b) indicate normal marine salinity throughout the T-OAE in the Cleveland Basin (Song et al., Reference Song, Littke and Weniger2017).

Reinterpretation of the Mo proxy (Section 13.a) weakens the geochemical argument for extreme basin restriction during the T-OAE. Like Mo, the oceans would have been subject to drawdown of U accompanying the global expansion of anoxic-euxinic bottom waters and seafloor area during the T-OAE. However, the differing redox sensitivities of the two elements would impose a trend of changing global Mo/U seawater ratios that compromise a simplistic interpretation of water mass restriction based on observations from modern basins.

14. Other trace-element redox proxies: Yorkshire coast studies

Several other trace elements documented in the Yorkshire coastal succession, not determined at Dove’s Nest, have been considered as redox proxies, including As, Cu, Cd and Co (Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; McArthur, Reference McArthur2019).

14.a. Arsenic

Arsenic displays similarities to the distribution to Mo when plotted stratigraphically as As/TOC ratios but As contents of the Pliensbachian section (Unit I) overlap more strongly with those in the higher beds (Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018). The AsEF profile shows a very similar pattern to DOPT (Fig. 14) that likely reflects deposition of As with Fe–Mn oxyhydroxides and subsequent incorporation by organic matter and diagenetic pyrite (Maher & Butler, Reference Maher and Butler1988; Large et al., Reference Large, Bull and Maslennikov2011). Arsenic has been used as a redox proxy in some previous palaeoenvironmental studies (e.g. Bodin et al., Reference Bodin, Godet, Matera, Steinmann, Vermeulen, Gardin, Adatte, Coccioni and Föllmi2007; Atar et al., Reference Atar, Marz, Aplin, Dellwig, Herringshaw, Lamoureux-Var, Leng, Schnetger and Wagner2019a, Reference Atar, Marz, Schnetger, Wagner and Aplin2019b; Benamara et al., Reference Benamara, Charbonnier, Adatte, Spangenberg and Föllmi2020). However, critical analyses by Algeo & Liu (Reference Algeo and Liu2020) and Tribovillard (Reference Tribovillard2020, Reference Tribovillard2021) have concluded that arsenic is not a reliable proxy.

Arsenic shows marked enrichments in sediments from settings where dissolved reactive iron has been captured from seawater via shuttling by Fe–Mn oxyhydroxides and/or associated with cold seeps and is trapped in the sediment under reducing conditions (Tribovillard, Reference Tribovillard2020, Reference Tribovillard2021). Enrichment in arsenic indicates that reactive iron was present in sufficient abundance to keep free sulfide ions to low levels, thus preventing As solubilization, migration and loss from the sediment. However, when reactive iron availability is low, the development of high-sulfide and low-iron conditions prevent As capture and enrichment. In this case, protracted contact between free sulfide and organic matter induces sulfurization of the latter (Abubakar et al., Reference Abubakar, Taylor, Coker, Wogelius and van Dongen2022), producing a high-TOC, high-Mo and low-As association (Tribovillard, Reference Tribovillard2020).

The good correlation between As vs Mo and U in the Yorkshire successions is consistent with a high reactive-Fe environment and supports the operation of a particulate shuttle during the deposition of the anoxic facies of the lower Bituminous Shales, Unit IV, evidenced by Mo vs U covariation (Fig. 18). However, it should be noted that Kemp et al. (Reference Kemp, Chen, Cho, Algeo, Shen and Ikeda2022a) proposed a coincident global drawdown of Mo, U and As from seawater accompanying the severe deep-ocean deoxygenation accompanying the T-OAE.

14.b. Copper

Copper is regarded as having a weak euxinic affinity (Algeo & Maynard, Reference Algeo and Maynard2004) and correlates strongly with TOC in the Pliensbachian – Toarcian boundary succession of the Cleveland Basin. This association is demonstrated by the near-identical chemostratigraphic profiles for the two constituents illustrated by Ruvalcaba Baroni et al. (Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018, fig. 2) and Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018, fig. 2), with a high correlation coefficient (R2 = 0.7) in both studies, but Cu shows no additional trends that may be considered to have palaeoenvironmental significance.

CuEF values are generally <1 throughout the succession, rise to >1 in samples with >7% TOC in T-OAE Subunit IIIb, reaching a maximum of 2.5 in the highest-TOC (>12%) samples at the level of Whale Stones ’bed’ 35. Low CuEF values (<3) similarly characterize oxic – euxinic settings in the modern Cariaco Basin. High CuEF values of 3 – 7 recorded in the euxinic sediments of the modern Black Sea and upwelling zones of the Peru margin (Little et al., Reference Little, Vance, Lyons and McManus2015) have not been documented in the Yorkshire succession.

14.c. Cadmium

McArthur (Reference McArthur2019) presented Cd and Co data from the Yorkshire coastal succession and combined these with Mo and Mn results to apply the trace-metal palaeoenvironmental proxies proposed by Sweere et al. (Reference Sweere, van den Boorn, Dickson and Reichart2016), and assessed their concordance with the observations of Little et al. (Reference Little, Vance, Lyons and McManus2015).

Cadmium and Mo stand out as the two trace metals that typically show some of the highest enrichment factors in organic-rich sediments (Brumsack, Reference Brumsack2006, fig. 2). Cadmium is enriched in plankton (35 ppm) but has a low concentration in terrigenous detritus, reflected by an average shale content of 0.3 ppm (Brumsack, Reference Brumsack1989; Little et al., Reference Little, Vance, Lyons and McManus2015). Similar to C and P, Cd is rapidly regenerated in the ocean water column and shows a typical nutrient profile, but undergoes a dramatic solubility decrease at the O2/H2S boundary of anoxic basins. The Cd curve for the Pliensbachian – Toarcian boundary succession (McArthur, Reference McArthur2019, fig. 2) generally follows the S and DOPT trends (Fig. 14), consistent with an association between Cd and sulfides (Cd vs S, R2 = 0.40). However, low Cd contents (0.01 – 0.88 ppm) in the Yorkshire succession indicate an absence of significant CdS addition from euxinic bottom waters (cf. Sweere et al., Reference Sweere, Dickson, Jenkyns, Porcelli, Ruhl, Murphy, Idiz, van der Boorn, Eldrett and Henderson2020) or a restricted Cd supply to the basin.

Unlike Cd, Mo does not bioaccumulate in phytoplankton (2 ppm), which have a Cd/Mo mass ratio of ≫1. As a result, sediments in upwelling regions, which are subject to high export production of organic matter and Cd but not Mo, are characterized by high Cd/Mo ratios. In restricted environments, although anoxia/euxinia promote uptake of both Cd and Mo to the sediments, lower productivity limits plankton-derived export of Cd, leading to Cd/Mo ratios of ≪1 that can approach the seawater value of 0.008. The hydrographically restricted and upwelling fields on the Cd/Mo discrimination diagrams of Sweere et al. (Reference Sweere, van den Boorn, Dickson and Reichart2016) are separated by a Cd/Mo value of 0.1.

The Cd/Mo profile from the Yorkshire succession (McArthur, Reference McArthur2019, fig. 3) generally follows the TOC and TOC/P trends (Figs 14, 16), consistent with a close association between Cd and TOC. However, despite a positive correlation (R2 = 0.50) between Cd vs TOC (McArthur, Reference McArthur2019, fig. 4), low Cd/Mo ratios (generally <0.1) throughout the succession place the overwhelming majority of samples in the restricted water mass field of Sweere et al. (Reference Sweere, van den Boorn, Dickson and Reichart2016, fig. 4) and are indicative of a low productivity setting.

However, interpretation of the Cd/Mo proxy based on direct comparison to values from modern basins is again compromised by evidence of significant stratigraphic changes in the global Mo content of ocean water through the late Pliensbachian – middle Toarcian, with the global drawdown of Mo accompanying the T-OAE (Section 13.a). This would change the threshold values between the restricted vs open basin fields, potentially placing samples from the T-OAE interval into the unrestricted high-productivity field.

14.d. Cobalt

Downward-decreasing vertical profiles of Co and Mn in the ocean water column demonstrate that these elements are scavenged by settling particulates (terrigenous detritus and organic matter) and are taken up by bottom sediments, principally as oxyhydroxides (e.g. van Hulten et al., Reference van Hulten, Middag, Dutay, De Baar, Roy-Barman, Gehlen, Tagliabue and Sterl2017; Hawco et al., Reference Hawco, Lam, Lee, Ohnemus, Noble, Wyatt, Lohan and Saito2018). In upwelling regions, the hydrogenous supply is limited by the depletion of both elements in upwelled water. By contrast, restricted basins commonly have abundant Co and Mn supplied by rivers, providing a greater hydrogenous supply.

In oxic sediments, Co and Mn are buried as oxyhydroxides but may be remobilized by bacterial reduction driving anoxia with subsequent refixation, typically Co in pyrite and organic matter and Mn in rhodochrosite. However, where the oxic – anoxic boundary lies at or above the sediment/water interface, remobilized elements can escape from the sediment and be returned to bottom waters. In open-ocean (unrestricted) settings, such as coast upwelling systems, these elements can then be lost by lateral advection of bottom waters. Oxyhydroxide particulates settling from the surface waters into the underlying oxygen-minimum zone dissolve, releasing Mn and Co, which can be transported oceanward by a ‘Mn-conveyor belt’ (Brumsack, Reference Brumsack2006), leading to low metal enrichments in the sediments.

In restricted settings, remobilized elements cannot escape and are returned to the sediment via redox cycling, causing significant enrichment over the original particulate supply. Sweere et al. (Reference Sweere, van den Boorn, Dickson and Reichart2016) determined that modern restricted environments have Co (ppm) × Mn (%) values > 0.4 while unrestricted environment values are < 0.4. McArthur (Reference McArthur2019) proposed the same threshold value for a CoEF × MnEF proxy and demonstrated the use of shale-normalized data offered no advantage over a concentration-based ratio.

Cobalt and Mn display similar profiles through the Yorkshire upper Pliensbachian – middle Toarcian succession (McArthur, Reference McArthur2019, fig. 2), excluding large Mn peaks associated with sideritic mudstone and calcite concretion horizons in Unit V (‘beds’ 44, 46, 48 and 50; Fig. 14) which are Co depleted. Other levels containing significant siderite and/or calcite contents similarly show relatively lower Co contents reflecting the presence of carbonate-bound Mn and, in the upper beds, Fe–Mn oxyhydroxides at omission surfaces in condensed intervals (McArthur, Reference McArthur2019).

The shape of the Co × Mn profile (McArthur, Reference McArthur2019, fig. 3) is similar to those of FeEF and MnEF (Fig. 14), with high Co (ppm) × Mn (%) values of >0.4 in the Pliensbachian section (Unit I) interpreted by McArthur (Reference McArthur2019) to indicate restricted circulation that pre-dated the onset of black shale deposition. Maximum Co × Mn contents are recorded in the three Sulphur Bands and high values occur consistently through the T-OAE interval, Unit III, which features the highest average Mn content and Mn/Al ratio of the units in the succession (Fig. 8). A sharp decline in Mn and MnEF (Figs 8, 14), with Co (ppm) × Mn (wt%) values falling to <0.4, occurs in the upper Bituminous Shales at the base of Unit V, consistent with a change to more open conditions. The subsequent decreasing stratigraphically upward trend implies increasing open circulation accompanied by sea-level rise from the late early Toarcian (cf. Fig. 10).

McArthur (Reference McArthur2019) concluded that the Cd/Mo, Co × Mn and Mo/TOC proxies together confirm that hydrographic restriction was a defining feature of black shale deposition in the early Toarcian of the Cleveland Basin. However, evidence of global depletion of Mo and other trace metals in the oceans during the T-OAE requires the application of Mo-based ratios to be re-evaluated.

14.e. Rhenium

The principal focus of work on Re and Os in the lower Toarcian has been the generation of 187Os/188Os and 187Re/188Os ratios for isochron construction and Re-Os dating of black shales, and the generation of 187Os/188Osi seawater time series (Fig. 18, Section 16.b.4) to examine changes in the balance between continental weathering and hydrothermal input in the oceans (e.g. Cohen et al., Reference Cohen, Coe, Bartlett and Hawkesworth1999, Reference Cohen, Coe, Harding and Schwark2004; Percival et al., Reference Percival, Cohen, Davies, Dickson, Hesselbo, Jenkyns, Leng, Mather, Storm and Xu2016; Them et al., Reference Them, Gill, Selby, Gröcke, Friedman and Owens2017b; van Acken et al., Reference van Acken, Tütken, Daly, Schmid-Röhl and Orr2019; Kemp et al., Reference Kemp, Selby and Izumi2020). Rhenium abundances for the Yorkshire Toarcian have been presented by Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008).

Rhenium has the largest enrichment factor of all redox-sensitive metals in anoxic sediments relative to the detrital background (Helz & Adelson, Reference Helz and Adelson2013). The average Re concentration of 66 ppb in modern anoxic sediments compares to an upper crust value of 0.2 – 0.4 ppb, an enrichment factor of >200 (Sheen et al., Reference Sheen, Kendall, Reinhard, Creaser, Lyons, Bekker, Poulton and Anbar2018). Rhenium occurs as the soluble Re(VII)O4 oxyanion in oxic seawater and is a conservative element that occurs at low concentrations (average 7 parts per trillion in the modern ocean; Anbar et al., Reference Anbar, Creaser, Papanastassiou and Wasserburg1992; Colodner et al., Reference Colodner, Sachs, Ravizza, Turekian, Edmond and Boyle1993b; Morford et al., Reference Morford, Martin and Carney2012; Dickson et al., Reference Dickson, Hsieh and Bryan2020). The estimated residence time of Re in the ocean is 130 – 750 ka (Colodner et al., Reference Colodner, Boyle and Edmond1993a; Miller et al., Reference Miller, Peucker-Ehrenbrink, Walker and Marcantonio2011), likely marginally shorter than for Mo at 440 – 800 ka (Morford & Emerson, Reference Morford and Emerson1999; Miller et al., Reference Miller, Peucker-Ehrenbrink, Walker and Marcantonio2011) and U at 400 ka (Ku et al., Reference Ku, Knauss and Mathieu1977) but longer than V at 50 –100 ka (Wu et al., Reference Wu, Owens, Huang, Sarafian, Huang, Sen, Horner, Blusztajn, Mprton and Nielsen2019).

Under anoxic conditions, Re(VII) is reduced to Re(IV), becomes insoluble and is removed to the sediment either through complexation with organic matter and/or incorporated into sulfides. Unlike Mo, Re can be efficiently removed from anoxic sediments at low dissolved H2S levels when bottom waters are dysoxic or anoxic (Crusius et al., Reference Crusius, Calvert, Pedersen and Sage1996; Calvert & Pedersen, Reference Calvert, Pedersen, Hillaire-Marcel and de Vernal2007). The ratio of Re/Mo in anoxic sediments is as much as 2.5 times greater than the seawater value, while the ratio is close to that of seawater in euxinic (sulfidic) sediments. Under oxic bottom waters the ratio may be less than that in seawater (Re/Mo 0.74 × 10−3; Bruland & Lohan, Reference Bruland, Lohan and Elderfield2003) due to fixation of Mo by Mn-oxyhydroxides but no uptake of Re.

Sediment Re concentrations and Re/Mo ratios have been used as local redox proxies by Turgeon & Brumsack (Reference Turgeon and Brumsack2006), Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008) and Kunert & Kendall (Reference Kunert and Kendall2023), amongst others. Rhenium offers a potential proxy for tracking general ocean oxygen depletion (i.e. combined euxinic and non-euxinic anoxia) because the magnitude of authigenic Re enrichment in anoxic marine sediments is significantly higher than the detrital background compared with U, for example, as reflected by higher Re enrichment factors in organic-rich sediments.

A geochemical study of a lower Toarcian succession in western Canada by Kunert & Kendall (Reference Kunert and Kendall2023) indicates that Re and Mo oceanic mass balance models can be used to infer the extent of seafloor total anoxia and euxinia, respectively, using concentrations of these metals in locally anoxic or euxinic carbonaceous mudstones deposited in an unrestricted marine setting. They calculated an expansion of up to ∼7% of total global seafloor anoxia – euxinia, dominated by euxinia, during the early stages of the T-OAE, followed by a contraction before the end of the event.

A Re threshold of 5 ppb was used by Kunert & Kendall (Reference Kunert and Kendall2023) to identify samples where the sediment column was anoxic. Values rise consistently above this threshold from the base of the T-OAE interval Unit III in Yorkshire (Fig. 20), with a large step increase to ∼40 ppb at the base of the laminated black shale facies in ‘bed’ 31 of Subunit IIIa interpreted to reflect bottom water anoxia and decreasing basin restriction.

Figure 19. Stratigraphic variation in selected isotope geochemistry in the upper Pliensbachian – middle Toarcian of Yorkshire. δ13Corg profile for Dove’s Nest (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022, black) rescaled to coastal succession, with compiled high-resolution coast curve (grey, see Fig. 2; Cohen et al., Reference Cohen, Coe, Harding and Schwark2004; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005; Littler et al., Reference Littler, Hesselbo and Jenkyns2010). Stratigraphic framework as in Figure 2. Rescaled whole-rock TOC profile for Dove’s Nest (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022, dark green) with coast composite data of Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011; thin yellow-green high-resolution curve), McArthur (Reference McArthur2019; thin pale green low-resolution curve) and Ruvalcaba Baroni et al. (Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018; green-filled triangles). Yorkshire coast Mo profile of McArthur (Reference McArthur2019; thick pink line) with elemental results for Mo-isotope samples of Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008; red filled triangles) and high-resolution data of Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; thin dark red line, see Fig. 16). δ98/95Mo coast profile from Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008). Details of the Mo concentration and isotope curves within Unit III (T-OAE) are presented in Figure 20. Belemnite carbonate-associated sulfur isotope profile (δ34SCAS) from Gill et al. (Reference Gill, Lyons and Jenkyns2011; white filled squares) incorporating the data of Newton et al. (Reference Newton, Reeves, Kafousia, Wignall, Bottrell and Sha2011; orange squares). Belemnite 87Sr/86Sr coast curve after McArthur et al. (Reference McArthur, Donovan, Thirlwall, Fouke and Mattey2000). 187Os/188Osi profile of Cohen et al. (Reference Cohen, Coe, Bartlett and Hawkesworth1999; Reference Cohen, Coe, Harding and Schwark2004). Belemnite carbonate oxygen-isotope (δ18O, blue dots) and Mg/Ca ratios (blue circles) after McArthur et al. (Reference McArthur, Donovan, Thirlwall, Fouke and Mattey2000).

Figure 20. Stratigraphic profiles of δ13Corg, TOC and selected trace-metal isotopes within Unit III, the T-OAE interval of the Yorkshire coast. Stratigraphy as in Figures 6, 7. Bulk rock δ13Corg and TOC profiles of the composite section from Hesselbo et al. (Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000, pale coloured lines) and DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005, dark lines), Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011, dark lines) – see Figure 7. Shaded grey bands indicate Unit III Subunits a – c (see Section 8.c); shaded blue band is the interval of the carbonate maximum, Subunit IIId. A – D mark coincident sharp falls in δ13Corg (after DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005; Cohen et al., Reference Cohen, Coe and Kemp2007) and δ98Mo, with increased Mo, as noted by Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011). Biotic extinction levels ii and iii after Caswell et al. (Reference Caswell, Coe and Cohen2009); base of trace fossil absent interval follows Caswell & Herringshaw (Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023). Thallium isotopes (ε205Tl) from Nielsen et al. (Reference Nielsen, Goff, Hesselbo, Jenkyns, LaRowe and Lee2011). Rhenium, Mo, Re/Mo and δ98Mo profiles from Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008). Note that Nielsen et al.’s (Reference Nielsen, Goff, Hesselbo, Jenkyns, LaRowe and Lee2011, fig. 5) comparison figure of ε205Tl vs δ98Mo at Port Mulgrave incorrectly plotted the position of the Mo dataset relative to the stratigraphy, as presented by Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008, fig. 2). The replotted data in our figure do not support an anti-correlation between these two isotope systems, as proposed by Nielsen et al. (Reference Nielsen, Goff, Hesselbo, Jenkyns, LaRowe and Lee2011) and modelled by Owens et al. (Reference Owens, Nielsen, Horner, Ostrander and Peterson2017).

Euxinic environments are predicted to be characterized by low Re/Mo ratios approaching those of seawater. Normal oxic environments show very variable but generally higher Re/Mo values. Re/Mo (ppb/ppm) thresholds of 10 – 15 for the oxic/dysoxic – anoxic boundary and 2 – 4 for the anoxic – euxinic boundary have been proposed by Turgeon & Brumsack (Reference Turgeon and Brumsack2006) and Kunert & Kendall (Reference Kunert and Kendall2023). The latter authors reported Re/Mo ratios of 11 ± 5, 6 ± 2 and 2 ± 1 for Toarcian carbonaceous mudstones assigned to dysoxic, anoxic and euxinic redox conditions based on independent criteria, respectively.

The Re/Mo profile spanning the T-OAE interval of the Yorkshire coast (Fig. 20) indicates: (1) a progressive transition from oxic – dysoxic conditions immediately preceding the onset of the T-OAE (top Subunit IIb and lower IIIa); (2) sharply falling ratios through ‘bed’ 32 at the top of Subunit IIIa, indicative of increasing local anoxia – euxinia; (3) consistently low values indicative of euxinic conditions throughout the δ13Corg minimum of Subunit IIIb; and (4) variable higher values representing euxinic –anoxic conditions during the later stages and immediately following the T-OAE (Subunits IIIc – IVa). The Re/Mo ratio of modern euxinic sediments approaches that of seawater, so the values recorded in Yorkshire T-OAE Subunit IIIb likely reflect the composition of Toarcian seawater (Fig. 20). The higher Re/Mo ratio compared to modern seawater would reflect a higher proportion of euxinic : anoxic seafloor during the Toarcian compared to the modern ocean.

14.f. I/Ca ratios

Iodine/calcium ratios (I/Ca) in marine carbonate have been proposed as a geochemical proxy to constrain global seawater redox change (e.g. Lu et al., Reference Lu, Jenkyns and Rickaby2010; Zhou et al., Reference Zhou, Jenkyns, Owens, Junium, Zheng, Sageman, Hardisty, Lyons, Ridgwell and Lu2015), and I/Ca ratios have been reported for belemnites collected from the Jet Rock (T-OAE Subunits IIIb – d) of the Yorkshire coast (Lu et al., Reference Lu, Jenkyns and Rickaby2010). Iodine has a residence time of ∼300 ka in the modern oceans resulting in a uniform total iodine concentration of 0.45 μM in most places, although the distribution of iodine species is complex (Wadley et al., Reference Wadley, Stevens, Jickells, Hughes, Chance, Hepach, Tinerl and Carpenter2020; Luther, Reference Luther2023). Low I/Ca ratios in the T-OAE and OAE2 intervals of Italy and England, respectively, have been interpreted to reflect the reduction of iodate (IO3) to iodide (I) due to deoxygenation and global iodine drawdown by organic carbon burial (Lu et al., Reference Lu, Jenkyns and Rickaby2010).

A fall to very low I/Ca ratios in bulk carbonate has been documented in a Toarcian shallow-water carbonate platform section in southern Italy (Monte Sorgenza; Lu et al., Reference Lu, Jenkyns and Rickaby2010). The onset of falling I/Ca values begins at a peak of ∼8 μM/M at the Pliensbachian – Toarcian boundary interval at Monte Sorgenza based on its position in relation to the δ13Ccarb curve (cf. Woodfine et al., Reference Woodfine, Jenkyns, Sarti, Baroncini and Violante2008; Xu et al., Reference Xu, Ruhl, Jenkyns, Leng, Huggett, Minisini, Ullmann, Riding, Weijers, Storm, Percival, Tosca, Idiz, Tegelaar and Hesselbo2018). A sharp step fall from >2 μM/M to a minimum ∼0.5 μM/M occurs at the base of, and continues throughout, the T-OAE interval. Interpretation of bulk carbonate I/Ca values in shallow-marine sediments requires caution due to the high sensitivity of I/Ca ratios to diagenesis (Lau & Hardisty, Reference Lau and Hardisty2022), but it is significant that comparable low ratios were recorded from Yorkshire Jet Rock belemnites (Lu et al., Reference Lu, Jenkyns and Rickaby2010, fig. 3).

I/Ca ratio values remain low until some distance above the T-OAE interval at Monte Sorgenza and only return to pre-excursion levels in the middle Toarcian. This is consistent with the persistence of global anoxia indicated by other geochemical (e.g. δ98Mo, Section 16.b.1; ε205Tl, Section 16.b.2) and fossil proxies (Section 19).

15. Organic geochemistry and biomarkers

The organic geochemistry of the Yorkshire lower Toarcian has been studied by Salem (Reference Salem2013), French et al. (Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014) and Song et al. (Reference Song, Littke and Weniger2017).

15.a. Molecular carbon-isotope records

Compound-specific δ13C analyses were performed by French et al. (Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014) across the T-OAE interval at Hawsker Bottoms (‘beds’ 29 – 41) that provide information on changes in the isotopic composition of the marine vs atmospheric carbon reservoirs. All biomarkers record the negative CIE of the T-OAE through Unit III but have different amplitudes (Fig. 21), consistent with an earlier low-resolution study of Toarcian shales in SW Germany (Schouten et al., Reference Schouten, Van Kaam-Peters, Rijpstra, Schoell and Sinninghe Damsté2000). The n-C17, n-C18 and n-C19 alkanes display a negative excursion of ∼2 – 3‰, with a marginally smaller excursion (∼2‰) recorded by pristane and phytane; all of these compounds are considered to represent marine sources (French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014). Long-chain n-alkanes (n-C27, n-C28, n-C29), derived primarily from land plant wax lipids, provide a terrestrial record. These exhibit the largest compound-specific negative CIE of ∼5‰, which is lower than that derived from bulk organic matter (δ13Corg ∼ 6 – 7‰; Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005; Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022).

Figure 21. Stratigraphic profiles of δ13Corg, δ13C n-alkanes, isorenieratane, AOM, TOC and δ15Ntot within Unit III, the T-OAE interval of the Yorkshire coast. Bulk rock δ13Corg and TOC profiles of the composite section from Hesselbo et al. (Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000, pale coloured lines) and DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005, dark lines), Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011, dark lines) – see Figure 7; ‘anoxic threshold’ of TOCWR = 2.5 wt% follows Algeo & Maynard (Reference Algeo and Maynard2004). δ13C data for terrestrial wood (Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000, brown squares) are offset but track the bulk sediment δ13Corg curve. Carbon isotope values from Hawsker Bottoms of representative long-chain n-alkane biomarkers (δ13Cn-alkane) derived from terrestrial plants (n-C27, n-C29) also display the negative excursion of the T-OAE (French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014). Short-chain n-alkanes (n-C17n-C19) attributed to marine plants follow an identical δ13C trend (French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014, fig. 7). The isorenieratane profile, a biomarker for anaerobic phototrophic green sulfur bacteria, provides evidence of reducing conditions developing during the initial phase of the T-OAE, peaking at the time of deposition of ‘bed’ 40 (Millstones); chlorobactane and okenane (not shown) display identical patterns (French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014, fig. 5). Amorphous organic matter (AOM) and algal cysts dominate (50 – >90%) the palynological assemblages (after Slater et al., Reference Slater, Twitchett, Danise and Vajda2019) of the high-TOC anoxic – euxinic facies. Peak species richness of calcareous nannofossils, preserved as carbonate and external moulds in organic matter throughout the section (Slater et al., Reference Slater, Bown, Twitchett, Danise and Vajda2022), occurs in Whale Stones ‘bed’ 35, together with a pulse of prasinophyte algae and dense granular organic matter (Houben et al., Reference Houben, Goldberg and Slomp2021). Nitrogen isotopes (δ15Ntot) display low values attributable to enhanced N2 fixation by cyanobacteria in a strongly redox-stratified marine environment (Wang et al., Reference Wang, Ossa, Spangenberg and Schoenberg2021). Toarcian open ocean seawater field derived from Tethyan sections (Section 16.a.1). The δ15Ntot profile (Jenkyns et al., Reference Jenkyns, Gröcke and Hesselbo2001) broadly follows TOC. A – D mark coincident sharp falls in δ13Corg (after DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005; Cohen et al., Reference Cohen, Coe and Kemp2007) and δ98Mo, as noted by Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011) – see Figure 20. Biotic extinction levels ii and iii after Caswell et al. (Reference Caswell, Coe and Cohen2009); base of trace fossil absent interval follows Caswell & Herringshaw (Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023).

The molecular C-isotope records show a more marked step down to lower values within the T-OAE than the bulk organic records, and within the CIE compound-specific δ13C values are relatively stable compared to the δ13Corg curve, which has a marked concave profile (Fig. 21). By comparison, matched δ13C datasets from other Toarcian sections in Spain and China show good agreement in trends between n-alkane, bulk organic and carbonate carbon isotope records (e.g. Ruebsam et al., Reference Ruebsam, Reolid and Schwark2020d, fig. 2). Offsets in the Yorkshire profiles may perhaps be, in part, an artefact of the low sampling resolution of the organic geochemical data (French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014). Alternatively, short-distance migration of the bitumen fraction, which is mobile and can move in the pore space, may have modified the trend in the compound-specific δ13C data. Changing organic matter composition between successive bulk samples may also play a role.

Organic matter source mixing will affect the trend and amplitude of the bulk organic δ13C excursion. The Yorkshire Pliensbachian – middle Toarcian succession displays large variation in the proportions of marine to terrestrial organic matter (Bucefalo Palliani et al., Reference Bucefalo Palliani, Mattioli and Riding2002; French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014; Slater et al., Reference Slater, Twitchett, Danise and Vajda2019). Calculated percentages of terrigenous organic matter based on linear regression of petrographic measurements of terrigenous organic matter and Rock-Eval Hydrogen Index (HI; French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014) indicate a fall from ∼80% at the top of Subunit IIb, immediately below T-OAE Unit III, to <20% in the TOC maximum in Whale Stones ‘bed’ 35 at the top of T-OAE Subunit IIIb, rising to ∼40% at the base of Subunit IVa.

A similar trend is shown by the palynological count data of Slater et al. (Reference Slater, Twitchett, Danise and Vajda2019) with corresponding terrestrial palynomorph values of 80 – 50% in Subunit IIb, falling to 8% in IIIb, then rising to 10 – 20% in Unit IV. It is notable that the ‘marine’ fraction in Units III and IV is overwhelmingly dominated by amorphous organic matter, peaking at 90% towards the top of Subunit IIIb. The amorphous organic matter is assumed here to be of predominantly marine origin based on its close association with prasinophyte algae and high HI values, but may include a terrestrial component (cf. Tyson, Reference Tyson and Tyson1995)

The very large increase in the proportion of marine organic matter in the lower Toarcian and particularly within the T-OAE interval will increase the amplitude of the T-OAE CIE in the Yorkshire curve. Where matching data are available for samples (e.g. Suan et al., Reference Suan, van de Schootbrugge, Adatte, Fiebig and Oschmann2015; Ruebsam et al., Reference Ruebsam, Muller, Kovacs, Palfy and Schwark2018), the δ13Corg curve may be normalized using HI or carbon preference index (CPI; the odd over even predominance for long-chain n alkanes, n-C24 to n-C34) values. Suan et al. (Reference Suan, van de Schootbrugge, Adatte, Fiebig and Oschmann2015, fig. 5d) applied this method to the Yorkshire coast paired δ13Corg and HI data of Sælen et al. (Reference Sælen, Tyson, Telnaes and Talbot2000) and calculated an amplitude for the corrected negative excursion of 3 – 4‰ δ13Corg. This value is consistent with corrected δ13Corg, δ13Cphytane and δ13Ccarb curves from two sections in SW Germany that are dominated by marine organic matter throughout the lower Toarcian (Suan et al., Reference Suan, van de Schootbrugge, Adatte, Fiebig and Oschmann2015).

Insufficient data are available to fully assess the impact of changes in organic matter composition on the fine structure of the Yorkshire δ13Corg profile. Palynological data do not show large sample-to-sample variation within the T-OAE (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019) interval, and steps within the Yorkshire δ13Corg curve have been correlated to those in records from Bornholm (Denmark), Sancerre (Paris Basin), sections in the Polish Basin, and at Peniche (Lusitanian Basin) (Fig. 3; Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000; Hesselbo & Pieńkowski, Reference Hesselbo and Pieńkowski2011; Hermoso et al., Reference Hermoso, Minoletti, Rickaby, Hesselbo, Baudin and Jenkyns2012; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018, fig. 9; Fantasia et al., Reference Fantasia, Adatte, Spangenberg, Font, Duarte and Föllmi2019, fig. 12).

Most significantly, the presence of a large negative CIE in long-chain n-alkanes that parallels the terrestrial wood and bulk organic δ13C trends in the T-OAE interval of the Cleveland Basin (Figs 3, 21) and δ13Ccarb trends at Peniche (Littler et al., Reference Littler, Hesselbo and Jenkyns2010, fig. 3) and elsewhere, confirms the synchronous 13C-depletion of the entire exchangeable carbon reservoir, in both the oceans and the atmosphere.

15.b. Biomarkers as palaeoredox and salinity proxies

Specific organic molecules in the geological record provide biomarkers for green sulfur-reducing bacteria that are confined to environments with available sunlight and euxinic conditions (e.g. Koopmans et al., Reference Koopmans, Koster, vanKaamPeters, Kenig, Schouten, Hartgers, deLeeuw and Sissingh Damsté1996; Grice & Eiserbeck, Reference Grice, Eiserbeck, Holland and Turekian2014). Their presence provides a signature for photic zone euxinia (PZE). Isorenieratane is the only uniquely defining molecule of green sulfur-reducing bacteria Chlorobiaceae, but its derivatives and certain other molecular compounds such as chlorobactane can also be used as proxies for the presence of these bacteria in the water column.

There is organic geochemical evidence for PZE during the T-OAE from multiple locations throughout Europe, including the UK, France, Germany, Hungary and Italy (Schouten et al., Reference Schouten, Van Kaam-Peters, Rijpstra, Schoell and Sinninghe Damsté2000; Pancost et al., Reference Pancost, Crawford, Magness, Turner, Jenkyns and Maxwell2004; Schwark & Frimmel, Reference Schwark and Frimmel2004; Van Breugel et al., Reference Van Breugel, Baas, Schouten, Mattioli and Sinninghe Damsté2006; Ruebsam et al., Reference Ruebsam, Muller, Kovacs, Palfy and Schwark2018; Xu et al., Reference Xu, Ruhl, Jenkyns, Leng, Huggett, Minisini, Ullmann, Riding, Weijers, Storm, Percival, Tosca, Idiz, Tegelaar and Hesselbo2018; Ajuaba et al., Reference Ajuaba, Sachsenhofer, Bechtel, Galasso, Gross, Misch and Schneebeli-Hermann2022; Burnaz et al., Reference Burnaz, Littke, Grohmann, Erbacher, Strauss and Amann2024). Bowden et al. (Reference Bowden, Farrimond, Snape and Love2006) reported isorenieratane derivatives in four samples from the Whitby Mudstone Formation at Port Mulgrave (‘beds’ 35, 37, 38 and 41) with additional records from ‘beds’ 31, 32 and 34 provided by Salem (Reference Salem2013), indicating that the photic zone was intermittently euxinic. Isorenieratane is also present in the high TOC intervals in each of the three Sulphur Bands (Salem, Reference Salem2013). Samples from Hawsker Bottoms confirm the presence of aromatic carotenoid biomarkers including isorenieratane from the Whitby Mudstone (Fig. 21), but samples also yield okenane, a biomarker derived from the red-coloured aromatic carotenoid okenone of purple sulfur bacteria (French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014). The highest recorded isorenieratane/TOC values occur at the top of Unit III at Hawsker Bottoms (Fig. 21).

Iron-rich, anoxic and potentially non-sulfidic (ferruginous) bottom-water conditions have been interpreted for lower Toarcian shales from the Cleveland Basin (Runswick Bay; Song et al., Reference Song, Littke and Weniger2017), based on biomarkers [low dibenzothiophene/phenanthrene ratios (< 0.2) and gammacerane indices (< 0.1)] and elemental composition (high total S and high S/TOC ratio). The abundance of small framboidal pyrite in these rocks (> 95% of < 7 μm diameter) suggests formation in the water column which depleted the free H2S, leading to a low degree of sulfurization of organic matter. The reactive iron was supplied by a high terrigenous influx, which is also indicated by the biomarker data.

The aryl isoprenoid ratio (AIR) relates the proportions of low to high molecular weight aromatic compounds (C13–17/C18–22; Schwark & Frimmel, Reference Schwark and Frimmel2004). A high abundance of short-chain forms indicate intense aerobic degradation. AIR values of ≤0.5 are associated with persistent PZE (near year-round) and high values of 0.5 – 3 reflect short-term episodic PZE (possibly seasonal). Runswick Bay samples from the Jet Rock and Bituminous Shales have yielded AIR ratios of 2 – 3 indicative of episodic PZE (Song et al., Reference Song, Littke and Weniger2017), supporting the evidence provided by aromatic carotenoid biomarkers.

Caswell & Coe (Reference Caswell and Coe2014) plotted the better-characterized Toarcian AIR record from Dotternhausen (Schwark & Frimmel, Reference Schwark and Frimmel2004) against the Yorkshire succession as a palaeoredox proxy, but it is unwise to assume that redox changes were synchronous and of the same amplitude in the two basins. For example, levels with significant bioturbation, indicating temporary oxygenation of bottom waters, characterize the beds immediately below and immediately above the Oberer Stein at Dotternhausen (Röhl et al., Reference Röhl, Schmid-Röhl, Oschmann, Frimmel and Schwark2001). The equivalent interval in the Yorkshire coastal and core successions (Top Jet Dogger – Millstone interval) shows no evidence of oxygenation and biomarkers suggest peak PZE (Fig. 21; French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014).

Modern observations suggest that the presence of okenone would require an extremely shallow PZE and a highly restricted depositional environment. Based on the abundance of okenane and evidence of microbial wavy lamination in Yorkshire Toarcian shales (O’Brien, Reference O’Brien1990), French et al. (Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014) suggested that the okenane, and potentially chlorobactane and isorenieratane, may have formed in benthic microbial mats rather than a euxinic photic-zone water column. The biomarkers must, nonetheless, reflect the presence of photic zone anoxic – euxinic bottom waters.

Multiple salinity proxies derived from the organic geochemical data [e.g. gammacerane/(gammacerane + C30-hopane) index] were interpreted as indicating normal marine salinity throughout the T-OAE in the Cleveland Basin by Song et al. (Reference Song, Littke and Weniger2017). These contradict the interpretation of strong brackish to freshwater conditions promoted by Remírez & Algeo (Reference Remírez and Algeo2020) based on element-ratio data (Section 13.d). However, gammacerane is considered to originate from tetrahymanole produced by ciliates, whose habitat is in the anoxic part of the water column beneath the pycnocline, so gammacerane may be better regarded as a stratification parameter (cf. Sinninghe Damsté et al., Reference Sinninghe Damsté, Kenig, Koopmans, Köster, Schouten, Hayes and de Leeuw1995) rather than a salinity indicator. Evidence of bacterial sources of tetrahymanole (Banta et al., Reference Banta, Wei and Welander2015) further complicate the use of the gammacerane proxy.

Other organic geochemical proxies (e.g. low ratio of steranes/hopanes, < 1.0; abundant long-chain n-alkanes) were interpreted by Song et al. (Reference Song, Littke and Weniger2017) to indicate a significantly higher input of terrigenous organic matter with more bacteria and/or cyanobacteria in Yorkshire compared to Toarcian black shale sections in the Netherlands, Germany and Luxemburg. These likely constitute components of the amorphous organic matter that dominates the palynofacies within the T-OAE interval (Section 19.b.2).

A comprehensive study of biomarkers in a core from southern Luxemburg (NE Paris Basin) by Ruebsam et al. (Reference Ruebsam, Mattioli and Schwark2022) indicates that the marine phytoplankton community structure during the T-OAE was dominated by eukaryotic algae, with a decrease in the proportion of prokaryotic bacterial biomass compared to the pre- and post-excursion intervals. A dominance of algae in the phytoplankton community during the T-OAE in the Cleveland Basin is similarly evidenced by palynological data (Section 19.b.2.).

Five distinct palaeoecological events within the interval of the δ13Corg minimum of the Paris Basin core each show a proliferation of opportunistic green algae coincident with a decline in red algae groups and shallow- and deep-dwelling calcareous nannoplankton (Ruebsam et al., Reference Ruebsam, Mattioli and Schwark2022). These events correlate to biomarker evidence [increasing ratio of tri- /di-MTTCs (methylated 2-methyl-trimethyltridecylchromans)] for episodes of surface-water freshening. These were interpreted to have been caused by higher precipitation and surface run-off accompanying climate warming, paced by changes in the short eccentricity (100 kyr) orbital cycle (Ruebsam et al., Reference Ruebsam, Mattioli and Schwark2022). To our knowledge, comparable data are not currently available from the Yorkshire succession.

16. Isotope geochemistry: insights into global processes

Chemostratigraphic profiles of key isotope ratios (δ15N, δ34S, δ18O, 87Sr/86Sr, δ98Mo, 187Os/188Osi, ε205Tl) constructed using available data from the Yorkshire coast to complement the δ13C, TOC and elemental data from Dove’s Nest presented herein, are plotted against a common composite succession (Fig. 2) based on measured outcrop sections between Hawsker Bottoms and Port Mulgrave in Figures 1921.

16.a. Nitrogen and sulfur isotopes

16.a.1. Nitrogen isotopes – increased algal productivity and partial denitrification

Reviews of the oceanic N-cycle have been provided by Ader et al. (Reference Ader, Sansjofre, Halverson, Busigny, Trindade, Kunzmann and Nogueira2014) and Algeo et al. (Reference Algeo, Meyers, Robinson, Rowe and Jiang2014). The δ15N (δ15/14N ‰air) composition of seafloor sediments on modern continental margins and in anoxic basins generally records the isotopic composition of primary producers in the overlying surface waters. The δ15N of the primary producers reflects the mass and isotope balance between the two main sources of nitrogen supporting new primary production: fixation of atmospheric N2 by diazotrophs (NO3) such as cyanobacteria and upwelled nitrate (NO3) and/or ammonium (NH4+) reaching the photic zone, utilized by phytoplankton. Partial water-column denitrification (i.e. generation of 14N-enriched gaseous species, N2 and NO2) and/or anammox (anaerobic ammonium oxidation) in dysoxic – anoxic environments results in an average δ15N value of +5‰ for NO3 in the modern ocean. Assimilation of 15N-enriched NO3 by phytoplankton leads to ocean sediments having typical δ15N values of +5‰ – +6‰.

A δ15N profile spanning the T-OAE interval at Hawsker Bottoms (Jenkyns et al., Reference Jenkyns, Gröcke and Hesselbo2001) comprises erratic low (< −2‰) but rising values from the base that reach a maximum of ∼+2‰ around the TOC maximum in Whale Stones ‘bed’ 35 (Fig. 21). Values fall sharply to between −1‰ and −2‰ at the base of ‘bed’ 36 (base Subunit IIIc), coincident with a marked fall in TOC and the onset of rising δ13Corg. δ15N displays a minimum value of ∼−4‰ at the base of ‘bed’ 41, immediately above the T-OAE interval. Very similar stratigraphic profiles and values were shown by lower resolution data from other UK successions in the Winterborne Kingston and Mochras boreholes from the Wessex and Cardigan Bay basins (Jenkyns et al., Reference Jenkyns, Gröcke and Hesselbo2001).

The interval of enhanced carbonate deposition exemplified by Whale Stones ‘bed’ 35 in Yorkshire, which can be traced across Germany (Unterer Stein), France and Switzerland (e.g. Riegraf, Reference Riegraf, Einsele and Seilacher1982), contains the TOC and δ15N maxima (Fig. 21). Current evidence does not support the suggestion of van de Schootbrugge et al. (Reference van de Schootbrugge, McArthur, Bailey, Rosenthal, Wright and Miller2005) that this represents a regional pulse in calcareous nannofossil productivity. In Germany, the Unterer Stein contains only rare nannofossils and exhibits a unique granular laminated fabric of submicrometric – micrometric anhedral calcite grains (microcarbs) and δ13Ccarb values of ∼ −12‰ (Röhl et al., Reference Röhl, Schmid-Röhl, Oschmann, Frimmel and Schwark2001; Bour et al., Reference Bour, Mattioli and Pittet2007; Wang et al., Reference Wang, Ossa, Wille, Schurr, Saussele, Schmid-Rohl and Schoenberg2020). These indicate bacterially induced carbonate precipitation associated with organic matter remineralization enhanced by microbial sulfate reduction. However, the supraregional extent of this diagenetic limestone indicates a period of significant palaeoenvironmental change immediately preceding the onset of rising global δ13C values.

A massive influx of prasinophyte algae (Halosphaeropsis liassica Mädler) and acritarchs occurs at this level in Yorkshire (Houben et al., Reference Houben, Goldberg and Slomp2021), together with a dense clustered AOM-dominated palynofacies, that may be indicative of photo-autotrophic prokaryotes (e.g., cyanobacteria, green- and purple sulfur bacteria) during peak euxinic conditions, supported by the presence of isorenieratane (French et al., Reference French, Sepulveda, Trabucho-Alexandre, Gröcke and Summons2014; Fig. 21). Prasinophyte algae possess a strong preference for ammonium over nitrate (Cochlan & Harrison, Reference Cochlan and Harrison1991) and can use NH4+ released by nitrogen fixers as their own nitrogen source. Stratification and photic-zone anoxia driving denitrification/anammox would favour organisms with low requirements for nitrate.

A δ15N curve spanning the T-OAE interval at Dotternhausen (Wang et al., Reference Wang, Ossa, Spangenberg and Schoenberg2021, fig. 4), shows a similar profile shape, but higher and less variable values, predominantly between +0.3‰ and +2.5‰. There, a peak, and step down in bulk rock δ15N, is associated with the Unterer Stein limestone in the mid-exaratum Zone and occurs at an identical stratigraphic position to the Whale Stones in Yorkshire with respect to the δ13Corg curve. Elevated δ15N values positively correlated with intervals of high TOC are apparent in other T-OAE sections in Europe (Jenkyns et al., Reference Jenkyns, Gröcke and Hesselbo2001; Ruebsam et al., Reference Ruebsam, Muller, Kovacs, Palfy and Schwark2018; Wang et al., Reference Wang, Ossa, Spangenberg and Schoenberg2021). Enhanced export of organic matter and its subsequent preservation under reducing conditions was associated with elevated denitrification rates. Thus, increased primary productivity combined with the intensification of water mass denitrification are proposed as the main processes controlling the changes in δ15N values.

Positive near-0‰ δ15N values recorded from Dotternhausen and other north European basins within the T-OAE interval were attributed by Wang et al. (Reference Wang, Ossa, Spangenberg and Schoenberg2021) to an episode of enhanced N2 fixation by cyanobacteria using Mo-based nitrogenase to compensate for severe bioavailable N loss. This was explained by the near-complete consumption of NO3 in strongly restricted redox-stratified water masses following extensive denitrification and/or anammox accompanying the T-OAE (Wang et al., Reference Wang, Ossa, Spangenberg and Schoenberg2021). However, it is seen that the 3 UK sections studied by Jenkyns et al. (Reference Jenkyns, Gröcke and Hesselbo2001), located in 3 different basins, yielded δ15N values ∼ −1‰ to −2‰ in organic-lean (<1% TOC) upper Pliensbachian and post-OAE Toarcian sediments that lack evidence of basin restriction. These negative values imply long-term 15N depletion in a regional NW European water mass rather than being driven by increasing basin restriction and environmental change accompanying the T-OAE.

The low δ15N values from levels outside the T-OAE interval are surprising. Nitrogen isotopic compositions that are significantly less than 0‰ in marine settings are typically restricted to organic matter-rich sediments and mudstones from Mediterranean sapropels and Mesozoic black shales (Junium et al., Reference Junium, Dickson and Uveges2018). Such low values could potentially reflect highly oligotrophic conditions that lacked significant denitrification and had low levels of nitrate utilization, similar to the modern eastern Mediterranean which has seawater δ15N values ∼ −1‰ (Jenkyns et al., Reference Jenkyns, Gröcke and Hesselbo2001).

Low 15N values (−4‰ to −1‰) have been documented in Cenomanian – Turonian boundary OAE2 successions (Algeo et al., Reference Algeo, Meyers, Robinson, Rowe and Jiang2014, fig. 6; Zhai et al., Reference Zhai, Zeng, Zhang and Yao2023, fig. 3) and other Phanerozoic OAE sediments (Ader et al., Reference Ader, Sansjofre, Halverson, Busigny, Trindade, Kunzmann and Nogueira2014) but significant spatial variant exists. Higgins et al. (Reference Higgins, Robinson, Husson, Carter and Pearson2012) attributed these low values to new production being driven by direct NH4 assimilation from upwelled anoxic deep water, supplemented by diazotrophy, with chemocline denitrification and anammox quantitatively consuming NO3. A marine nitrogen reservoir dominated by NH4, in combination with kinetic isotope effects, would have led to eukaryotic biomass depleted in 15N (Higgins et al., Reference Higgins, Robinson, Husson, Carter and Pearson2012). Although this is a feasible explanation for the low δ15N values accompanying OAE2 it is difficult to apply to the long-term 15N depletion that characterizes Lower Jurassic records reported by Jenkyns et al. (Reference Jenkyns, Gröcke and Hesselbo2001) outside the T-OAE interval. Furthermore, by contrast to OAE2, trends within the T-OAE evidence increasing rather than decreasing δ15N values accompanied by intensified euxinia (Fig. 21).

It is evident that δ15N values reported at all sites in the Jenkyns et al. (Reference Jenkyns, Gröcke and Hesselbo2001) study consistently include a high proportion of values of ≪0‰. This contrasts with the N-isotope data reported in all later Toarcian studies, including those from Switzerland (Rietheim; Montero-Serrano et al., Reference Montero-Serrano, Föllmi, Adatte, Spangenberg, Tribovillard, Fantasia and Suan2015), Germany (Dotternhausen; Wang et al., Reference Wang, Ossa, Spangenberg and Schoenberg2021), Hungary (Réka Valley; Ruebsam et al., Reference Ruebsam, Muller, Kovacs, Palfy and Schwark2018) and Algeria (Ratnek El Kahla, Ruebsam et al., Reference Ruebsam, Reolid, Marok and Schwark2020b; Mellala, Baghli et al., Reference Baghli, Mattioli, Spangenberg, Ruebsam, Schwark, Bensalah, Sebane, Pittet, Pellenard and Suan2022), where δ15N values are consistently positive and generally lie between +1 – +3‰. These values are typical of greenhouse climate ocean sediments throughout the Phanerozoic (Algeo et al., Reference Algeo, Meyers, Robinson, Rowe and Jiang2014) and likely reflect the δ15N composition of early Toarcian open-ocean seawater (Fig. 21).

It should be noted that the incorrect use of HCl for decarbonation can generate to δ15N values as low as −12‰ from samples with positive values. Provisional results from the Dove’s Nest core have yielded δ15N values of +2 – +3‰ for selected samples throughout the succession. We recommend that additional nitrogen isotope studies of the Yorkshire and other T-OAE successions should be carried out to test the reliability of the Jenkyns et al. (Reference Jenkyns, Gröcke and Hesselbo2001) data. In the meantime, the δ15N values reported in that work should be treated with caution.

16.a.2. Sulfur isotopes – 32S drawdown and oceanic sulfate depletion

The sulfur isotope composition of ocean water is controlled principally by a balance between the input of relatively 34S-enriched sulfate from weathering of the continents and the output of 34S-depleted sulfur fixed and buried as pyrite and organic sulfur in sediments (Claypool et al., Reference Claypool, Holzer, Kaplan, Sakai and Zak1980). Modern rivers have variable δ34S (δ34/32S %‰VCDT) sulfate values of 0‰ to +10‰; sulfides in modern marine sediments are typically around −40‰ δ34S, although a large range is observed (Paytan et al., Reference Paytan, Yao, Gray, Gradstein, Ogg and Ogg2020). Sulfate is the second most abundant anion in the oceans and has a residence time of 10 – 20 Ma, which far exceeds the mixing time of the ocean ∼1 – 2 ka (Paytan et al., Reference Paytan, Kastner, Campbell and Thiemens2004) and seawater sulfate today has a constant δ34S value of +21.0 ± 0.2‰ (Rees et al., Reference Rees, Jenkins and Monster1978).

The sulfur isotope composition of carbonate-associated sulfate (δ34SCAS; i.e. sulfate substituted in the crystal lattice of carbonate minerals) provides a means of establishing past seawater values (Kampschulte & Strauss, Reference Kampschulte and Strauss2004). Data obtained from belemnite guards through the Pliensbachian – Toarcian of the Yorkshire coast (Kampschulte & Strauss, Reference Kampschulte and Strauss2004; Gill et al., Reference Gill, Lyons and Jenkyns2011; Newton et al., Reference Newton, Reeves, Kafousia, Wignall, Bottrell and Sha2011) display a marked 6‰ δ34SCAS increase from ∼+16‰ to +22‰ through the C. exaratum Subzone (T-OAE Subunits IIIb – d; Fig. 19). By comparison, Raiswell et al. (Reference Raiswell, Bottrell, Al-Biatty and Tan1993) analysed 11 pyrite samples from the same interval (i.e. Jet Rock) with an average δ34Spyrite value of −24 ± 2‰.

Samples from Dotternhausen (belemnites) and Monte Sorgenza, Italy (whole-rock carbonate) display similar δ34S values and increases across the T-OAE interval to those in Yorkshire (Gill et al., Reference Gill, Lyons and Jenkyns2011) indicating a consistent pattern of sulfur isotope variation throughout northern European basins and western Tethys. A compilation of global widely spaced δ34SCAS and δ34S evaporite data spanning the Jurassic suggests a return to pre-T-OAE seawater δ34S values of ∼+17‰ through the later Toarcian into the Aalenian (Gill et al., Reference Gill, Lyons and Jenkyns2011, fig. 6), representing an ∼8 Ma recovery period. This indicates a relatively long residence time for sulfate in the Jurassic ocean; estimates from modelling by Gill et al. (Reference Gill, Lyons and Jenkyns2011) put Toarcian marine sulfate concentrations at 4 to 8 mM compared to 28 mM today.

By contrast to the European data, sections in southern Tibet, representative of the southeastern open Tethys, display a much larger positive sulfur-isotope excursion, rising from ∼+20‰ around the Pl–To boundary to ∼+40‰ at the top of the T-OAE interval (Newton et al., Reference Newton, Reeves, Kafousia, Wignall, Bottrell and Sha2011; Han et al., Reference Han, Hu, He, Newton, Jenkyns, Jamieson and Franceschi2022, Reference Han, Hu, Newton, He, Mills, Jenkyns, Ruhl and Jamieson2023). The rate of isotopic change with time-based on the Tibetan sections indicates that early Jurassic seawater sulfate concentrations may have been as low as 1 – 5 mM.

The δ34S increase accompanying the T-OAE demonstrates a globally significant perturbation in the sulfur cycle driven by the drawdown of isotopically light 32S by pyrite and organic sulfur burial in euxinic environments. Modelling by Gill et al. (Reference Gill, Lyons and Jenkyns2011) indicated that pyrite deposition in the northern European epicontinental seaway was insufficient to cause the observed isotope excursion, accounting for at most 4% of the pyrite burial required. Substantial additional pyrite burial would be needed to drive the documented excursion, requiring a much wider extent of euxinic conditions in the world ocean during the T-OAE, although the impacts of an increase in 32S-enriched organic sulfur burial and a possible near cessation Ca-sulfate evaporite deposition during the T-OAE also need to be considered (Han et al., Reference Han, Hu, Newton, He, Mills, Jenkyns, Ruhl and Jamieson2023).

The large differences in δ34SCAS values and the amplitude of the excursions accompanying the T-OAE in Europe and Tibet imply substantial regional differences in local sulfur cycling (Gill et al., Reference Gill, Lyons and Jenkyns2011; Han et al., Reference Han, Hu, He, Newton, Jenkyns, Jamieson and Franceschi2022, Reference Han, Hu, Newton, He, Mills, Jenkyns, Ruhl and Jamieson2023). The contrasting records might be attributed to varying modifications of open-ocean sulfate isotopic compositions by water-mass isolation and the effects of changing regional weathering and pyrite burial fluxes. Significant spatial heterogeneity of δ34SCAS values and trends have also been observed over OAE2 (Zhai et al., Reference Zhai, Zeng, Zhang and Yao2023) indicating that sulfate in the Cretaceous ocean was also not as well mixed as today and that δ34SCAS was influenced by regional oceanic redox conditions.

Modelling by Han et al. (Reference Han, Hu, He, Newton, Jenkyns, Jamieson and Franceschi2022) indicates that the initial formation of an isotopically heterogeneous ocean for seawater sulfate began in the Pliensbachian as concentrations began to fall. At the time of maximum early Toarcian euxinia, during the T-OAE, this may have culminated in a large difference in the scale of response between the European epicontinental sea and western Tethyan continental margin, where coeval isotopic values are comparable, and the easterly Tibetan shelf of southern Tethys. The greater amplitude of the sulfur isotope excursion in Tibet has been ascribed to the upwelling of 34S-enriched equatorial deep water, while the lower magnitude of the Toarcian sulfur excursion in Europe reflects a smaller, local reduced-sulfur sink (Han et al., Reference Han, Hu, Newton, He, Mills, Jenkyns, Ruhl and Jamieson2023).

A two-phase pattern of ocean deoxygenation during the early Toarcian is indicated by the Tibetan δ34SCAS data: the first began around the Pl–To boundary and continued up to the onset of the T-OAE; the second coincided with the onset of rising δ13C values and terminated around the end of the T-OAE. The termination of deoxygenation also marks the point where the sulfur cycle reaches a new steady state with a negligible Ca-sulfate burial flux. Higher positive δ34S values were maintained after the T-OAE by a long-term global reduction in Ca-sulfate (gypsum and anhydrite) burial, driven by falling and then continuous low seawater-sulfate concentrations during and after the T-OAE.

The depletion of marine sulfate during the T-OAE has implications for other geochemical processes. Adams et al. (Reference Adams, Hurtgen and Sageman2010) have argued that sulfate levels regulate the recycling of phosphorus during remineralization of organic matter by microbial sulfate reduction. Efficient phosphorus recycling is considered to be essential to maintaining high levels of primary productivity and anoxic conditions during OAEs (Van Cappellen & Ingall, Reference Van Cappellen and Ingall1994). Thus, sulfate depletion accompanying and following the T-OAE, driven by pyrite burial, provides a significant negative feedback mechanism that would favour the termination of widespread anoxia and the T-OAE.

16.b. Trace-metal isotopes – Mo, Tl, U, Os

The relatively long modern ocean residence times of Mo (440 – 800 ka), U (∼450 ka), Tl (∼20 ka) and Os (∼10 – 40 ka) compared to the ocean mixing time of ∼1 – 2 ka, combined with distinctive stable-isotope fractionation association with different marine sinks, means that the isotopic composition of these metals has potential to serve as global oceanic tracers (Kendall et al., Reference Kendall, Andersen, Owens, Ernst, Dickson and Bekker2021). Uranium and Mo isotope systems offer established tracers of global oceanic redox conditions, particularly for the extent of anoxic and euxinic seafloor, and Tl isotopes provide a proxy for the extent of well-oxygenated seafloor characterized by Mn oxyhydroxide burial (Kendall et al., Reference Kendall, Andersen, Owens, Ernst, Dickson and Bekker2021). Marine carbonaceous mudstones provide the principal archive for reconstructing palaeoredox histories from these isotope records. Osmium isotopes enable the identification of changes in weathering versus hydrothermal fluxes (Peucker-Ehrenbrink & Ravizza, Reference Peucker-Ehrenbrink, Ravizza, Gradstein, Ogg and Ogg2020) that might impact global productivity.

Variations in trace-metal isotope records from the Toarcian of Yorkshire have previously been attributed principally to local changes in water chemistry in the Cleveland Basin (Cohen et al., Reference Cohen, Coe, Harding and Schwark2004; McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Pearce et al., Reference Pearce, Cohen, Coe and Burton2008; Nielsen et al., Reference Nielsen, Goff, Hesselbo, Jenkyns, LaRowe and Lee2011; Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017), but there is now increasing evidence that they incorporate a significant component of global ocean change (e.g. Them et al., Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022).

16.b.1. Molybdenum isotopes – a proxy for global euxinic seafloor area

Molybdenum isotopes, reported as δ98Mo (δ98/95Mo ‰SRM3143), have become established as a valuable proxy to assess temporal variation in the global extent of seafloor euxinia and organic burial in the oceans. The redox sensitivity of Mo isotopes has been reviewed by Kendall et al. (Reference Kendall, Andersen, Owens, Ernst, Dickson and Bekker2021). Chemically inert molybdate (MoO42–) is the principal species in oxygenated rivers and seawater, but this is converted to particle-reactive thiomolybdates (MoO4−xSx2−) and polysulfides in sulfidic bottom waters and pore waters.

Modern seawater has a δ98Mo value of 2.3 ± 0.1‰. The dominant Mo input (95%) to the oceans, from rivers and groundwater, has an isotopic composition similar to the average upper crust value of ∼0.6‰ δ98Mo. The influence of hydrothermal fluids (∼1.5‰ δ98Mo) is negligible (5%). Molybdenum isotope fractionation in marine environments results in preferential removal of light Mo isotopes to sediments. The largest fractionation (−3‰) occurs in well-oxygenated settings, where Mo is adsorbed to Mn oxyhydroxides with average values of around −0.7‰ δ98Mo, leading to 98Mo enrichment of seawater.

Under dysoxic to anoxic bottom waters, where dissolved sulfide occurs in sediment pore waters, Mo isotope fractionation is variable. Here, it depends on the type of oxyhydroxides delivering Mo to sediments and pore-water H2S concentration, with Mo burial efficiency increasing with the level of dissolved sulfide. The average fractionation is −0.8‰ and an output value of ∼1.5‰ δ98Mo. Larger fractionations reflect weakly to intermittently euxinic conditions and/or a strong Fe–Mn oxyhydroxide particulate flux that delivers isotopically light Mo to sediments.

The δ98Mo of euxinic sediments approaches that of seawater when Mo removal from bottom waters is near quantitative, particularly in significantly restricted, but not isolated, basins with water stratification and highly sulfidic bottom waters like the modern Black Sea (Kendall et al., Reference Kendall, Andersen, Owens, Ernst, Dickson and Bekker2021). The δ98Mo of seawater sequestered into euxinic mudstones is therefore hypothesized to reflect the δ98Mo of seawater, and variations in δ98Mo with stratigraphic level through a black shale section have been interpreted to reflect a changing balance of sequestration globally into oxic and euxinic sediments (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Pearce et al., Reference Pearce, Cohen, Coe and Burton2008). The oceanic Mo mass balance is highly sensitive to the extent of sulfidic (especially euxinic) environments. Hence, seawater Mo concentrations and residence times will be lower in ancient oceans with a greater extent of euxinic seafloor than today (Reinhard et al., Reference Reinhard, Planavsky, Robbins, Partin, Gill, Lalonde, Bekker, Konhauser and Lyons2013).

A Mo isotope profile for the Yorkshire coast spanning the lower Toarcian, sampled at Port Mulgrave, Saltwick Bay and Hawsker Bottoms, was published by Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008). Their Mo and δ98Mo data are plotted in Figures 19 and 20 and compared to companion curves for ε205Tl, Re and Re/Mo spanning the T-OAE interval in Figure 20. Light δ98Mo values of ∼0‰ and low Mo concentrations of ∼0.5 ppm below the T-OAE interval in basal Toarcian Unit II (Figs 18, 20) are consistent with adsorption of Mo to Mn oxyhydroxides in sediments underlying an oxic water column.

Molybdenum isotope values increase markedly to >1‰ δ98Mo and Mo concentrations rise to ≥4 ppm at the base of T-OAE Unit III indicating the establishment of dysoxic and then anoxic bottom waters (Fig. 20; Pearce et al., Reference Pearce, Cohen, Coe and Burton2008). δ98Mo oscillates with an amplitude of ∼0.7‰ through the T-OAE interval. Mo contents show small sharp increases (reflecting a higher availability of Mo in the basin seawater) at stratigraphic levels where δ98Mo becomes lighter and δ13Corg falls (events A – D of Pearce et al., Reference Pearce, Cohen, Coe and Burton2008), highlighted by the pale-coloured curves in Figure 20. Associated Re/Mo ratios remain low, consistent with persistent euxinia and show no consistent pattern. The four δ98Mo maxima are ∼ 1.5‰. Evidence of significant environmental change at these levels is provided by the biota: the abundance and average size of Pseudomytiloides dubius (Sowerby) shells vary within the same stratigraphic levels as the δ98Mo record (Caswell et al., Reference Caswell, Coe and Cohen2009; Caswell & Coe, Reference Caswell and Coe2013). Periods of higher primary productivity and greater oxygenation accompanying more open basin conditions, indicated by increased δ98Mo values, supported larger bivalves.

Molybdenum isotope values increase from the top of the T-OAE interval into the Bituminous Shales (H. falciferum Subzone, Fig. 19) with δ98Mo values of 1.8 ± 0.2‰ in Unit IV. The increase begins immediately above the top of the laminated black shale facies of the Jet Rock at the level of the Upper Pseudovertebrae concretions (cf. Howarth, Reference Howarth1962) in ‘bed’ 38 and coincides with the onset of increasing Re and Mo contents, falling Re/Mo ratio and increasing ε205Tl (Fig. 20). Subunits IVa and b each display a broad peak in δ98Mo with a maximum of up to 2.1‰, associated with high Mo and Re values, with the two subunits being separated by a trough in all three constituents. δ98Mo values fall back to ∼ 1.5‰ in uppermost lower and basal middle Toarcian Unit V, based on 4 data points (Fig. 19).

Global expansion of oceanic euxinia, evidenced by low Mo and Mo/TOC values in T-OAE black shales from open marine basins imply a significant drawdown of the marine Mo inventory during the early Toarcian (Them et al., Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022; Section 13.a); this would be expected to cause a decrease in seawater δ98Mo. The observed opposing shift to higher δ98Mo accompanying the onset of the T-OAE in the Cleveland Basin, therefore, is interpreted to reflect the local change to more intensely reducing conditions at the onset of the T-OAE, as evidenced by Fe-speciation and other redox proxy data (Fig. 14). Changes in local bottom-water redox exerted a greater control on stratigraphic changes in δ98Mo than the global expansion of oceanic euxinia (Fig. 20).

Drawdown of Mo during the T-OAE (Section 13.a) potentially decreased the oceanic residence time of Mo sufficiently to cause seawater δ98Mo values to fluctuate rapidly in response to redox variations or even become spatially heterogeneous between different ocean basins. Sections of carbonaceous mudstones in Germany (Dotternhausen) and the Netherlands (Rijswijk core) do not show high-amplitude oscillations in δ98Mo within the T-OAE interval like those observed in Yorkshire (Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017), so the Yorkshire cycles are unlikely to reflect global changes in the Mo continental weathering flux or euxinic seafloor area as proposed by Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008).

A small global marine-Mo reservoir in conjunction with basin restriction potentially resulted in the large regional variability of water mass (and therefore sedimentary) Mo concentrations and δ98Mo observed in different European basins during the early Toarcian. However, it is significant that Mo/TOC ratios, which may also reflect the inventory of dissolved Mo in euxinic marine basins (Algeo & Lyons, Reference Algeo and Lyons2006) display similar minimum values of ∼1 within the T-OAE intervals in Yorkshire (Fig. 16), Germany and the Netherlands (Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017, fig. 6).

Dickson et al. (Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017) and Kendall et al. (Reference Kendall, Andersen, Owens, Ernst, Dickson and Bekker2021) suggested that the highest δ98Mo values of ∼1.45‰ in the T-OAE interval in Yorkshire may represent a close estimate of global seawater δ98Mo during the early Toarcian. Episodes of decreased restriction of the Cleveland Basin, however, might allow an increased inflow of Mo from open ocean seawater at a rate faster than the removal of Mo to sediments (Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017). This would lead to a euxinic basin with incomplete Mo drawdown (i.e. non-quantitative Mo removal) and sediments fractionated by ∼−0.7‰ from open ocean seawater, as seen in the modern Cariaco Basin (Arnold et al., Reference Arnold, Anbar, Barling and Lyons2004). T-OAE sections in SW Germany (Dotternhausen) and the Netherlands (Rijswijk core) also have lower δ98Mo than the inferred seawater δ98Mo value of ∼1.45‰ within the T-OAE interval (Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017), consistent with less restricted and/or less reducing conditions at those localities.

High δ98Mo values of up to 2.1‰ characterize Unit IV (Fig. 19). Authigenic phases in modern marine sediments do not record δ98Mo values higher than coeval seawater, so this higher value implies rising seawater δ98Mo accompanying a global decrease in euxinic seafloor area during the later early Toarcian.

16.b.2. Thallium isotopes – assessing the extent of oxic seafloor area

Thallium isotopes offer further potential to constrain the geological evolution of seawater oxygenation on a basin to a global scale (Nielsen et al., Reference Nielsen, Rehkämper and Prytulak2017; Ostrander et al., Reference Ostrander, Owens and Nielsen2017; Owens et al., Reference Owens, Nielsen, Horner, Ostrander and Peterson2017; Owens, Reference Owens2019). Thallium isotopes are reported relative to Tl metal standard SRM997 using the epsilon notation where:

$${{\rm{\varepsilon }}^{{\rm{205}}}}{\rm{Tl }} = {\rm{ [}}{\left( {^{{\rm{205}}}{\rm{Tl}}{{\rm{/}}^{{\rm{203}}}}{\rm{Tl}}} \right)_{{\rm{sample}}}}{\rm{/}}{\left( {^{{\rm{205}}}{\rm{Tl}}{{\rm{/}}^{{\rm{203}}}}{\rm{Tl}}} \right)_{{\rm{SRM9971}}}} - {\rm{1]}} \times {10^4}$$

Thallium is supplied to the oceans by rivers, mineral aerosols (dust), volcanic gases and particles and high-temperature hydrothermal fluids, all of which have similar ε205Tl values of around −2 and pore-water fluxes from continental-margin sediments ε205Tl with a value of around 0. The only two important marine Tl sinks are adsorption by authigenic phases, principally Mn oxyhydroxides in pelagic clays, with ε205Tl values of ∼+8 to +12 and uptake of Tl during low-temperature alteration of oceanic crust with an ε205Tl value of −7.2. For short-term ≤1 Ma redox events it is considered that changes in the Tl isotopic composition of seawater are driven principally by changing rates of Mn-oxyhydroxide burial (Owens et al., Reference Owens, Nielsen, Horner, Ostrander and Peterson2017), since changes in the rate of oceanic crust alteration will operate on longer time scales.

The relatively short modern ocean residence time for Tl of ∼20 ka, which is nonetheless one order of magnitude longer than the ocean mixing time, ensures a uniform ε205Tl value of −6.0 ± 0.6 in modern open-ocean seawater (Nielsen et al., Reference Nielsen, Rehkämper and Prytulak2017; Owens et al., Reference Owens, Nielsen, Horner, Ostrander and Peterson2017; Owens, Reference Owens2019). Mn-oxyhydroxides that are ubiquitous in sediments deposited from oxygenated bottom waters, combined with the detrital mineral fraction, generate ε205Tl signatures in bulk sediments and partial extracts that are generally isotopically heavier than ambient seawater (e.g. Wang et al., Reference Wang, Lu, Costa and Nielsen2022, but see discussion therein). Oxyhydroxides are absent under anoxic or euxinic conditions because the reduction of insoluble Mn (IV) to more soluble Mn (II) and Mn (III) occurs when oxygen is removed from the water in contact with the sediment. In this case, the ε205Tl composition from the oxic part of the water column has been shown to transfer quantitatively to the authigenic fraction (principally Fe sulfides) in sediments underlying euxinic (Owens et al., Reference Owens, Nielsen, Horner, Ostrander and Peterson2017) or highly reducing anoxic bottom waters (Fan et al., Reference Fan, Nielsen, Owens, Auro, Shu, Hardisty, Horner, Bowman, Young and Wen2020), which may then provide a record of the isotopic evolution of the surface ocean. In partially and predominantly restricted basins ε205Tl values vary between open-ocean (−6) and continental material (−2), scaling with the degree of restriction, as exemplified by the Cariaco Basin (−5.4 – weak restriction) and the Black Sea (−2.3 – strong restriction) (Owens et al., Reference Owens, Nielsen, Horner, Ostrander and Peterson2017).

Thallium isotope data in early diagenetic pyrite from the upper beds of the T-OAE (top ‘bed’ 32 – 39) at Port Mulgrave were published by Nielsen et al. (Reference Nielsen, Goff, Hesselbo, Jenkyns, LaRowe and Lee2011). ε205Tl values show a large range from −8 to −2 (Fig. 20). Thallium isotope data for authigenic Tl, based on bulk rock extracts from the upper Pliensbachian – middle Toarcian of 3 sections from the Western Canada Sedimentary Basin, Alberta, Canada and SW German Basin (Dotternhausen) were presented by Them et al. (Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018). The Canada profiles, obtained from sections that were subject to consistent anoxic – euxinic conditions throughout the study interval (based on FeHR/FeT and FePy/FeHR Fe-speciation proxies) yielded ε205Tl values close to modern seawater of −6 through the upper Pliensbachian. They then rise upward from around the bottom of the base Toarcian D. tenuicostatum Zone equivalent, to −2 at the onset of the T-OAE. ε205Tl falls back to a broad trough of −4 mid-T-OAE, with a second peak of −2 at the top of the T-OAE and then falls upward towards −4 in the H. bifrons equivalent Zone. A similar trend and values were observed in the SW German Basin Toarcian profile (Them et al., Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018, fig. 3).

The increase in ε205Tl at the Pliensbachian – Toarcian boundary in the Western Canada Sedimentary Basin was interpreted by Them et al. (Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018) to mark the onset of increasing global bottom-water anoxia that preceded the T-OAE by 600 ka and was sustained into the middle Toarcian H. bifrons Zone, representing an interval of ∼2 Ma. Rising values were attributed to a global reduction in Fe–Mn oxyhydroxide precipitation accompanying oxygen depletion, with an accompanying decrease in 205Tl output flux leading to rising ε205Tl seawater values. Authigenic Tl concentrations show no systematic variation throughout the section (Them et al., Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018), which suggests the marine inventory was not depleted due to increased anoxic conditions.

Thallium isotopes from the T-OAE interval in Yorkshire show a wider range and extend to lower values (ε205Tl −8 to −2; Nielsen et al., Reference Nielsen, Goff, Hesselbo, Jenkyns, LaRowe and Lee2011) compared to Toarcian seawater values interpreted from the successions in western Canada and Germany (ε205Tl −4 to −2; Them et al., Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018). Shifts to values below those of open ocean water (Fig. 20) might be attributed to short-term basin-scale Mn-oxyhydroxide burial events in oxic areas of the Cleveland Basin margin driving local episodes of 205Tl drawdown from the local water mass.

Interestingly, the increase of ε205Tl in the lower Toarcian profile in Canada and to some extent in Yorkshire (Fig. 20), is matched by an increasing δ34SCAS trend in the Cleveland Basin (Fig. 19) but also in Italy (Monte Sorgena) and southern Tibet (Han et al., Reference Han, Hu, He, Newton, Jenkyns, Jamieson and Franceschi2022, Reference Han, Hu, Newton, He, Mills, Jenkyns, Ruhl and Jamieson2023). Broadly parallel trends for thallium and sulfur isotopes are a characteristic feature of the long-term Cenozoic record (Nielsen et al., Reference Nielsen, Rehkämper and Prytulak2017).

Finally, it should be noted that the comparison figure of ε205Tl vs δ98Mo at Port Mulgrave by Nielsen et al. (Reference Nielsen, Goff, Hesselbo, Jenkyns, LaRowe and Lee2011, fig. 5) incorrectly plots the position of the Mo dataset relative to the stratigraphy, as presented by Pearce et al. (Reference Pearce, Cohen, Coe and Burton2008, fig. 2). The correctly aligned stratigraphic profiles are shown in Figure 20. These do not support an anti-correlation between these two isotope systems, as proposed by Nielsen et al. (Reference Nielsen, Goff, Hesselbo, Jenkyns, LaRowe and Lee2011) and modelled by Owens et al. (Reference Owens, Nielsen, Horner, Ostrander and Peterson2017).

16.b.3 Uranium isotopes – potential insights into global anoxia

Uranium is insoluble in its reduced form and, in contrast to Mo, does not require dissolved sulfide for its incorporation and burial in sediments (Sections 13.b, d). U-isotope data (δ238/235U ‰SRM145) from organic-rich mudstones and carbonates are increasingly employed as a tool for tracing global oceanic redox changes during Phanerozoic anoxic events (e.g. Clarkson et al., Reference Clarkson, Stirling, Jenkyns, Dickson, Porcelli, Moy, Pogge von Strandman, Cooke and Lenton2018; Lau et al., Reference Lau, Romaniello and Zhang2019; Cheng et al., Reference Cheng, Elrick and Romaniello2020; McDonald et al., Reference McDonald, Partin, Sageman and Holmden2022; Kulenguski et al., Reference Kulenguski, Gilleaudeau, Kaufman, Kipp, Tissot, Goepfert, Pitts, Pierantoni, Evans and Elrick2023). However, isotope fractionation factors and U removal mechanisms in anoxic marine environments need to be better constrained (Kendall et al., Reference Kendall, Andersen, Owens, Ernst, Dickson and Bekker2021), and the basis for global oceanic redox interpretations has been questioned (Cole et al., Reference Cole, Planavsky, Longley, Böning, Wilkes, Wang, Swanner, Wittkop, Loydell, Busigny, Knudsen and Sperling2020; Zhang et al., Reference Zhang, Lenton, del Rey, Romaniello, Chen, Planavsky, Clarkson, Dahl, Lau, Wang, Li, Zhao, Isson, Algeo and Anbar2020; Gangl et al., Reference Gangl, Stirling, Jenkyns, Preston, Clarkson, Moy, Dickson and Porcelli2023). Surprisingly, to our knowledge, no U-isotope study has been published for the T-OAE.

16.b.4. Osmium isotopes – radiometric age and constraining continental weathering

The rhenium–osmium (Re–Os) isotope system can be used to obtain absolute ages for the deposition of marine mudstone successions, although the precision of Re–Os dates generally remains poorer than that currently achievable using U–Pb and 40Ar/39Ar geochronometers (Schmitz et al., Reference Schmitz, Singer, Rooney, Gradstein, Ogg and Ogg2020). By contrast, the 187Os/188Os composition of ancient marine sediments, when age corrected for the contribution of 187Re decay (187Os/188Osi), records temporal variation in seawater that provides valuable evidence of palaeoenvironmental change (e.g. Cohen, Reference Cohen2004; Katchinoff et al., Reference Katchinoff, Syverson, Planavsky, Evans and Rooney2021).

Samples from the C. exaratum and H. falciferum subzones of the Yorkshire coast have yielded imprecise Re–Os isochron ages of 181 ± 13 Ma (Cohen et al., Reference Cohen, Coe, Bartlett and Hawkesworth1999) and 178 ± 5 Ma (Cohen et al., Reference Cohen, Coe, Harding and Schwark2004), respectively. Subsequently, an older, more precise, Re–Os age of 183.0 ± 2.0 Ma has been obtained for the equivalent Toarcian H. falciferum Subzone Posidonia Shale of SW Germany (van Acken et al., Reference van Acken, Tütken, Daly, Schmid-Röhl and Orr2019), consistent with the base Toarcian age of the 184.20 Ma assigned by GTS2020 (Hesselbo et al., Reference Hesselbo, Ogg, Ruhl, Hinnov, Huang, Gradstein, Ogg, Schmitz and Ogg2020b) and 183.73 + 0.35/−0.50 Ma calculated using U–Pb zircon ages from Argentina (Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022).

Modern deep ocean water has a globally uniform 187Os/188Os value of 1.06 ± 0.1. The 187Os/188Os composition is predominantly controlled by the mass balance of two end-member Os isotope components: weathered continental crust (∼1.4) and mantle inputs (0.13) derived from submarine volcanism and basalt weathering (Peucker-Ehrenbrink & Ravizza, Reference Peucker-Ehrenbrink and Ravizza2002, Reference Peucker-Ehrenbrink, Ravizza, Gradstein, Ogg and Ogg2020). This, coupled with the short residence time of Os in seawater of ∼10 – 40 ka (Levasseur et al., Reference Levasseur, Birck and Allègre1999; Oxburgh, Reference Oxburgh2001) makes the 187Os/188Osi composition of marine sediments an excellent monitor of palaeoenvironmental change in the geological record (Cohen, Reference Cohen2004; Peucker-Ehrenbrink & Ravizza, Reference Peucker-Ehrenbrink, Ravizza, Gradstein, Ogg and Ogg2020).

The first 187Os/188Os data spanning the lower Toarcian and the T-OAE interval, published by Cohen et al. (Reference Cohen, Coe, Harding and Schwark2004), were derived from the Yorkshire coastal succession. 187Os/188Osi values of ∼0.3 in the lowest Toarcian rise sharply to a maximum of ∼1.0 within the CIE interval and fall back to ∼0.4 immediately above (Fig. 19). Comparable data have been generated subsequently for Mochras, Wales (Percival et al., Reference Percival, Cohen, Davies, Dickson, Hesselbo, Jenkyns, Leng, Mather, Storm and Xu2016); East Tributary, Alberta, Canada (Them et al., Reference Them, Gill, Selby, Gröcke, Friedman and Owens2017b); Dormettingen, SW Germany (van Acken et al., Reference van Acken, Tütken, Daly, Schmid-Röhl and Orr2019); and Sakuraguchi-dani, Japan (Kemp et al., Reference Kemp, Selby and Izumi2020) with similar trends observed in all sections. This peak in 187Os/188Osi values suggests that the flux of radiogenic 187Os derived from continental crust weathering increased significantly during the T-OAE but declined rapidly after the event.

It has been suggested that weathering of juvenile basalt from LIPs (Jenkyns et al., Reference Jenkyns, Dickson, Ruhl and Van den Boorn2017), including the Karoo–Ferrar (Cohen & Coe, Reference Cohen and Coe2007), would have significantly increased the input of non-radiogenic Os into seawater during and after their emplacement. As a result, the increase in continental weathering of ancient crust must have been of much greater magnitude than prior to LIP emplacement to cause the excursion to more radiogenic Os values developed during the T-OAE. Evidence from the Karoo, however, indicates the lavas were emplaced very rapidly as a succession of single eruptive events each lasting less than a century, and with the total duration of the main eruptive phase perhaps as short as 250 kyr (Moulin et al., Reference Moulin, Fluteau, Courtillot, Marsh, Delpech, Quidelleur and Gérard2017). Individual flows generally lack macroscopically distinct weathering products, one of the lines of evidence supporting their rapid outpouring, which precludes basalt weathering from being a major source of non-radiogenic Os to seawater during the T-OAE.

Previously, McArthur et al. (Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008) have argued that the increase in 187Os/188Osi in the Yorkshire succession was a local artefact caused by marine Os drawdown in a restricted local water mass sensitising the system to small changes in riverine input that drove large excursions in 187Os/188Osi. These, it was suggested, also compromised the Re–Os dating, generating the large errors associated with the Yorkshire ages (Cohen et al., Reference Cohen, Coe, Bartlett and Hawkesworth1999, Reference Cohen, Coe, Harding and Schwark2004). However, synchronous positive 187Os/188Osi excursions on a global scale, including sites in western North America and Japan that show no evidence for water-mass restriction (Them et al., Reference Them, Gill, Selby, Gröcke, Friedman and Owens2017b; Kemp et al., Reference Kemp, Selby and Izumi2020) refute this argument.

The pattern and magnitude of 187Os/188Osi changes in Japan and North America at the onset of the δ13Corg negative excursion (chemostratigraphic Subunit IIIa of the Yorkshire stratigraphy) match closely, rising from 0.2 to 0.6 in both sections, before falling back to a plateau of ∼0.4 in the post-T-OAE section. Matching patterns in sections on two sides of the Panthalassa Ocean (Fig. 1) strongly suggest that these variations reflect changes in the composition of global seawater Os at the onset of the T-OAE. However, the drawdown of redox-sensitive trace metals accompanying the T-OAE, combined with a pulse of radiogenic Os driven by enhanced weathering, might adversely impact age modelling, particularly for the C. exaratum Zone.

The inverse relationship between δ13Corg and 187Os/188Osi at the onset of the T-OAE demonstrates that the flux of radiogenic osmium to the oceans increased synchronously with the decrease in carbon-isotope values, evidence of coupling between a massive release of isotopically light carbon and enhanced global continental crust weathering (Kemp et al., Reference Kemp, Selby and Izumi2020). However, the greater amplitude of the 187Os/188Osi peak in Yorkshire compared to Canada and Japan may reflect the influence of regional weathering and basin restriction enhancing the global seawater signal. Deviations from open ocean 187Os/188Osi values have been observed in marginal seas (Marquez et al., Reference Marquez, Tejada, Suzuki, Peleo-Alampay, Goto, Hyun and Senda2017) and restricted ocean basins (Dickson et al., Reference Dickson, Davies, Bagard and Cohen2022a) of various ages.

The 187Os/188Osi excursion in Yorkshire was interpreted by Cohen et al. (Reference Cohen, Coe, Harding and Schwark2004) as resulting from a 400 – 800% increase in continental chemical weathering rates due to an accelerated hydrological cycle. Using data from western North America and numerical modelling Them et al. (Reference Them, Gill, Selby, Gröcke, Friedman and Owens2017b) calculated that weathering may have increased by 215% and potentially up to 530% during the T-OAE compared to the pre-event baseline. Kemp et al. (Reference Kemp, Selby and Izumi2020) compared 187Os/188Osi and δ13Corg data from Japan with the European and North American profiles and concluded that abrupt negative shifts in carbon isotopes in the Japan profile were coeval with rapid increases in weathering by >40% across each of these intervals. Overall, they concluded that global weathering rates may have increased up to 600% through the entire T-OAE.

Values derived from osmium isotope data are consistent with an independent estimate of a ∼500% increase in global weathering rate during the T-OAE based on Ca isotopes (δ44/40Ca) in brachiopods and bulk carbonate from Peniche (Brazier et al., Reference Brazier, Suan, Tacail, Simon, Martin, Mattioli and Balter2015). However, δ44/40Ca belemnite values from Peniche and Yorkshire (Q Li et al., Reference Li, McArthur, Thirlwall, Turchyn, Page, Bradbury, Weis and Lowry2021), when corrected for temperature and normalized to a constant Mg/Ca ratio, indicate that no significant change in the Ca isotope composition of seawater accompanied the T-OAE, so the comparable weathering estimates may be fortuitous. Better constraints on the rates and magnitudes of change in the radiogenic (continental weathering) vs non-radiogenic (hydrothermal) Os fluxes are required to further reduce uncertainty.

16.c. Strontium isotopes

The strontium (Sr) isotope composition of seawater essentially represents a mixture of unradiogenic Sr from the hydrothermal alteration of mid-ocean ridge basalt (87Sr/86Sr ≅ 0.703) and radiogenic Sr in rivers (≅ 0.711) from the weathering of ancient continental crust (McArthur et al., Reference McArthur, Howarth, Shields, Zhou, Gradstein, Ogg, Schmitz and Ogg2020). Modern seawater has an 87Sr/86Sr value of 0.709174 ± 0.000002. A record of ancient seawater 87Sr/86Sr ratios may be obtained from the analysis of biogenic calcite in marine fossils including belemnites, brachiopods and foraminifera.

87Sr/86Sr ratio data obtained from Yorkshire coast belemnites (Jones et al., Reference Jones, Jenkyns and Hesselbo1994; McArthur et al., Reference McArthur, Donovan, Thirlwall, Fouke and Mattey2000) underpin the construction of the upper Pliensbachian – lower Toarcian global Sr-isotope reference curve of McArthur et al. (Reference McArthur, Howarth, Shields, Zhou, Gradstein, Ogg, Schmitz and Ogg2020). The stratigraphic profile (Fig. 19) shows falling values through the A. margaritatus Zone (Staithes Sandstone and Cleveland Ironstone Penny Nab Member; Subunits Ia – b), the continuation of a long-term trend beginning in the early Sinemurian (∼200 Ma). A step fall in 87Sr/86Sr ratios at the base of the P. spinatum Zone (Kettleness Member; Subunit Ic), together with possible steps in δ18Obel and Mg/Cabel ratio (Fig. 19), support the presence of a large stratigraphic gap at the disconformity. The Sr-isotope trend through the P. spinatum Zone is less well defined but shows generally falling-upward values, with a reversal to gently rising values at the base of the PlToBE, towards the top of the zone. This pattern represents one of the most rapid reversals in 87Sr/86Sr of the Phanerozoic record (McArthur et al., Reference McArthur, Howarth, Shields, Zhou, Gradstein, Ogg, Schmitz and Ogg2020, fig. 7.2).

A marked change to steeply rising 87Sr/86Sr ratios is observed through the C. exaratum Subzone (Jet Rock; Subunits IIIb – d) of the T-OAE interval (Fig. 19), prior to a return to more gently increasing values through the remainder of the Toarcian. The change at the base of the C. exaratum Subzone has been ascribed to an increase in the flux of radiogenic Sr from the chemical weathering of continental crust because it coincides with a marked excursion in the seawater osmium (Cohen & Coe, Reference Cohen and Coe2007; Section 16.b.4) and other weathering proxies. The increase in chemical weathering has been linked to an increase in average temperatures across the T-OAE (Fig. 19, Section 21.d.7) and an accelerated hydrological cycle (Cohen et al., Reference Cohen, Coe, Harding and Schwark2004; Them et al., Reference Them, Gill, Selby, Gröcke, Friedman and Owens2017b; Izumi et al., Reference Izumi, Kemp, Itamiya and Inui2018).

McArthur et al. (Reference McArthur, Donovan, Thirlwall, Fouke and Mattey2000, Reference McArthur, Steuber, Page and Landman2016) have suggested that the increase in the Yorkshire profile (Fig. 19) is attributable to a decrease in the sedimentation rate driven by condensation and hiatuses in the C. exaratum Subzone (Subunits IIIb – d). Condensation at this level based on changes in the 87Sr/86Sr profile was also indicated by Jenkyns et al. (Reference Jenkyns, Jones, Gröcke, Hesselbo and Parkinson2002, fig. 13). Indeed, the sharp break in slope of the 87Sr/86Sr at the base of the P. spinatum Zone (Fig. 19) coincides with the disconformity at the base Pecten Seam in coastal sections (Section 8.a.2). However, as noted by Cohen et al. (Reference Cohen, Coe, Harding and Schwark2004), an a priori assumption that 87Sr/86Sr ratios ratio must always change linearly with time is hard to justify in view of the reversal of the seawater Sr-isotope curve in the latest Pliensbachian ∼1 Ma earlier.

Nonetheless, the presence of prominent carbonate concretion horizons throughout the laminated black carbonaceous mudstones of the C. exaratum Subzone supports the presence of significant hiatuses (cf. Raiswell, Reference Raiswell and Marshall1987, Reference Raiswell1988; Marshall & Pirrie, Reference Marshall and Pirrie2013) and an overall reduction in bulk sedimentation rates. The interval showing an increased slope in the Sr-isotope profile is bounded by the Cannon Ball Doggers at the bottom and the Millstones at the top of the subzone (Fig. 19).

17. Geochronology and cyclostratigraphy – age and duration of the T-OAE

The current international chronostratigraphic chart of the ICS (https://stratigraphy.org/ICSchart/ChronostratChart2023-09.pdf) retains the GTS2020 age of 184.2 ± 0.3 Ma for the base of the Toarcian Stage (Gradstein et al., Reference Gradstein, Ogg, Schmitz and Ogg2020). New radiometric dates (U–Pb zircon ages) from the T-OAE interval of the Neuquén Basin, Argentina (Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022) have provided an age of 183.73 + 0.35/−0.50 Ma for the Pliensbachian – Toarcian boundary, which is consistent with an astronomical age of 183.70 ± 0.50 Ma derived from the Mochras succession (Storm et al., Reference Storm, Hesselbo, Jenkyns, Ruhl, Ullmann, Xu, Leng, Riding and Gorbanenko2020).

An age of 182.77 + 0.11/−0.15 Ma was calculated for the Tethyan D. tenuicostatumH. serpentinum zonal boundary by Al-Suwaidi et al. (Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022), but this was judged to be coincident with onset of the δ13C fall at the base of the T-OAE. The onset of the δ13C fall lies in the upper D. tenuicostatum Zone in Yorkshire (mid-D. semicelatum Subzone) and not at the base of the H. serpentinum Zone (Fig. 2). By comparison, a base H. serpentinum Zone age of 182.8 Ma was calculated by Ruebsam et al. (Reference Ruebsam, Mayer and Schwark2019) and 183.16 Ma was assigned by GTS2020 (Hesselbo et al., Reference Hesselbo, Ogg, Ruhl, Hinnov, Huang, Gradstein, Ogg, Schmitz and Ogg2020b). The correlation between South America and Europe requires additional constraints. Other U–Pb ages from Peru, USA and Canada lack sufficient stratigraphic control with respect to European ammonite zones for high-resolution correlation to Yorkshire or the Toarcian GSSP at Peniche (Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022). Available radiometric dates are insufficient to rigorously assess changes in sedimentation rate within the Pliensbachian – Toarcian succession.

Cohen et al. (Reference Cohen, Coe and Kemp2007), following DB Kemp et al. (Reference Kemp, Coe, Cohen and Schwark2005), offered an estimate of ∼350 ka for the duration of the T-OAE as defined here in Yorkshire. Cyclostratigraphically inferred durations of the T-OAE, mainly derived from the Paris and Lusitanian basins, remain controversial, with two notably different estimates of 300 – 500 ka (e.g. Boulila et al., Reference Boulila, Galbrun, Huret, Hinnov, Rouget, Gardin and Bartolini2014) and ∼900 ka (Suan et al., Reference Suan, Pittet, Bour, Mattioli, Duarte and Mailliot2008b; Huang & Hesselbo, Reference Huang and Hesselbo2014). Differences relate principally to the allocation of obliquity (35 or 37.5 ka) vs short eccentricity (100 ka) as periodicities to the dominant cyclicity.

Spectral analysis of Yorkshire coast carbonate and δ13Corg data by Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011) clearly identified cyclicity within the upper D. tenuicostatum – lower H. exaratum subzones, and lower H. falciferum Subzone, but cycles were absent in the upper C. exaratum Subzone. Spectral analysis of high-resolution elemental time series from Yorkshire by Thibault et al. (Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018) provided a continuous cycle record spanning the T-OAE (Fig. 7), including the complete C. exaratum Subzone, that enabled a cycle-to-cycle correlation between Yorkshire, Peniche and Sancerre. A suggested duration of either 560 ka (obliquity-based) or 1500 ka (short-eccentricity based) was assigned to the T-OAE.

Orbital tuning of uppermost Pliensbachian – lower Toarcian δ13C and CaCO3 time series, based on the short and long, stable 405 ka (g2–g5) eccentricity cycles, from the Talghemt section in Morocco provided a duration of ∼400 – 500 ka for the T-OAE (Boulila et al., Reference Boulila, Galbrun, Sadki, Gardin and Bartolini2019). This time interval is very close to that inferred by Boulila et al. (Reference Boulila, Galbrun, Huret, Hinnov, Rouget, Gardin and Bartolini2014) from the Sancerre core in the Paris Basin (300 – 500 ka) and to a revised estimate from the Peniche section of ∼472 ka (Boulila & Hinnov, Reference Boulila and Hinnov2017). By contrast, a cyclostratigraphic study of high-resolution CaCO3 and TOC data from the Posidonia Shale of Dotternhausen (SW German Basin) tuned to the 100 ka and 405 ka eccentricity cycles, yielded ∼1.2 Ma for the event (Ruebsam et al., Reference Ruebsam, Schmid-Röhl and Al-Husseini2023) based on the recognition of 3 long-eccentricity cycles spanning the interval of the negative CIE.

A duration of ∼1.2 Ma is consistent with the assignment of a 405 ka periodicity to the Zr/Rb and other geochemical detrital-proxy cycles observed through the T-OAE in Yorkshire (Figs 4, 7; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018), as illustrated by Ruebsam et al. (Reference Ruebsam, Schmid-Röhl and Al-Husseini2023, fig. 7). Incorporating our lower placement of the base of the T-OAE (Fig. 7, 174.5 – 164.1 m in the Dove’s Nest core) compared to Ruebsam et al. (Reference Ruebsam, Schmid-Röhl and Al-Husseini2023) yields ∼1.4 Ma for Unit III and the full extent of the CIE in Yorkshire. Fendley et al. (Reference Fendley, Frieling, Mather, Ruhl, Hesselbo and Jenkyns2024), incorporating recent U–Pb dates, derived a duration of ∼ 1 Ma for the negative CIE interval in their Mochras study. By comparison, McArthur et al. (Reference McArthur, Steuber, Page and Landman2016) derived an estimated duration for the C. exaratum Subzone alone (i.e. the upper part of the T-OAE) of 1.1 – 2.4 Ma based on global rates of change in Sr isotopes.

Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011) found no evidence for changing sedimentation rate before, during or after the T-OAE in Yorkshire, although a small hiatus at the large negative δ13Corg step within ‘bed’ 34 has been proposed (Kemp et al., Reference Kemp, Coe, Cohen and Weedon2011; Boulila & Hinnov, Reference Boulila and Hinnov2017; Boulila et al., Reference Boulila, Galbrun, Sadki, Gardin and Bartolini2019; Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019). Nonetheless, the presence of prominent concretion horizons indicates that multiple levels of significant condensation and likely hiatuses are present in the succession (cf. Raiswell, Reference Raiswell and Marshall1987, Reference Raiswell1988; Marshall & Pirrie, Reference Marshall and Pirrie2013).

Significant differences exist between authors in the placement of boundaries for the T-OAE, the recognition and duration of cycles, the equivalence of biostratigraphic marker taxa and correlation of the lower Toarcian between Yorkshire, Mochras, Sancerre, Peniche and other sections (e.g. compare Boulila & Hinnov, Reference Boulila and Hinnov2017; Thibault et al., Reference Thibault, Ruhl, Ullmann, Korte, Kemp, Gröcke and Hesselbo2018; Boulila et al., Reference Boulila, Galbrun, Sadki, Gardin and Bartolini2019; Ruebsam et al., Reference Ruebsam, Schmid-Röhl and Al-Husseini2023). The carbon isotope correlation (Fig. 3) reveals very large differences in absolute and relative thicknesses of biozones between sites in different basins. Further work is required to obtain higher resolution, stratigraphically longer, radiometrically dated time series for rigorous cyclostratigraphic analysis.

18. The Toarcian hyperthermal – temperature and salinity records

Carbonate oxygen-isotope (δ18Obel) and Mg/Ca ratios obtained from belemnites (Mg/Cabel) collected from multiple Pliensbachian – Toarcian sections along the Yorkshire coast have been reported by Sælen et al. (Reference Sælen, Doyle and Talbot1996), McArthur et al. (Reference McArthur, Donovan, Thirlwall, Fouke and Mattey2000), Bailey et al. (Reference Bailey, Rosenthal, McArthur, van de Schootbrugge and Thirlwall2003) and Ullmann et al. (Reference Ullmann, Thibault, Ruhl, Hesselbo and Korte2014, Reference Ullmann, Frei, Korte and Hesselbo2015). Based on the Mg/Ca temperature dependency observed in modern biogenic calcites, an increase in Mg/Cabel values from ∼ 8 to 16 mmol/mol (Fig. 19), coincident with decreasing δ13Corg values in the lower T-OAE interval (Subunits IIIa – b), was interpreted by Bailey et al. (Reference Bailey, Rosenthal, McArthur, van de Schootbrugge and Thirlwall2003) to indicate a warming of 6 – 7° C.

However, Ullmann et al.’s (Reference Ullmann, Frei, Korte and Hesselbo2015) study of the chemical and isotopic architecture of a belemnite rostrum from the Grey Shale Member (‘bed’ 21, D. tenuicostatum Subzone) at Hawsker Bottoms, showed no correlation between δ18O and Mg/Ca ratios in the rostrum, arguing against a significant temperature dependence in Mg/Ca ratios. Instead, the data were interpreted to indicate a strong dependency of Mg/Ca ratios on metabolic effects on element ratios within the (internal) mineralizing fluids, alongside growth rates and crystal orientation effects. Nonetheless, the strong coherence between stratigraphic trends of the Mg/Cabel and δ18Obel palaeotemperature proxies (Fig. 19) remains compelling.

A concomitant δ18Ocarb negative shift of 3‰ at the onset of the T-OAE [from −0.5‰ to −3.5‰; Fig. 19], if caused by temperature alone, would imply an increase of 13° C, but was interpreted to be the product of 6 – 7° C warming accompanied by a small salinity decrease of ∼2.5 psu by Bailey et al. (Reference Bailey, Rosenthal, McArthur, van de Schootbrugge and Thirlwall2003). Additionally, Ullmann et al. (Reference Ullmann, Thibault, Ruhl, Hesselbo and Korte2014) proposed that a change in belemnite life habits from cool bottom-water to warm surface-water dwelling taxa accompanied the development of anoxic bottom waters, leading to an overestimation of the true increase in average water temperature. More generally, δ18Obel (and arguably Mg/Cabel) values from the upper Pliensbachian indicate the presence of cooler and relatively more saline waters, with a cooling trend prior to early Toarcian warming and hyperthermal conditions accompanying the T-OAE (Fig. 19).

A salinity decrease of ∼5 psu for surface waters in the Cleveland Basin during the T-OAE was calculated by Sælen et al. (Reference Sælen, Doyle and Talbot1996) based on salinity modelling of their Yorkshire δ13Cbel and δ18Obel data. Dera & Donnadieu (Reference Dera and Donnadieu2012) used a fully coupled ocean–atmosphere model (FOAM) to study the palaeoclimatic and palaeoceanographic consequences of increases in atmospheric pCO2 levels at a multiscale resolution. They interpreted a regional reduction in salinity (∼1 – 5 psu) across NW European basins during early Toarcian warming due to an input of brackish water from the Arctic Ocean through the Viking Corridor. Modelling by Ruvalcaba Baroni et al. (Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018) indicated a near mean-ocean salinity of ∼35 psu in southern European epicontinental basins, falling to ∼ 33 psu in more northerly areas like the Cleveland Basin.

Extreme brackish (10 – 25 psu) waters during the T-OAE in the Cleveland Basin, with freshwater conditions during late C. exaratum Subzone time (top Subunit IIIb and IIIc), were proposed by Remírez & Algeo (Reference Remírez and Algeo2020) based principally on the use of a Ba/Ga proxy. However, as noted by Hesselbo et al. (Reference Hesselbo, Little, Ruhl, Thibault and Ullmann2020a), the presence of belemnites throughout the study interval (Doyle, Reference Doyle1990; Caswell & Coe, Reference Caswell and Coe2014, fig. 2; Ullmann et al., Reference Ullmann, Thibault, Ruhl, Hesselbo and Korte2014, fig. 2), particularly during the supposed ‘freshwater’ interval and their incorporated geochemical data (Fig. 19), together with abundant ammonites (Howarth, Reference Howarth1962) and pseudoplanktonic crinoids, contradict such low salinity conditions. They are also arguably inconsistent with organic biomarker data (Section 15.b).

Belemnite O-isotope and Mg/Ca records from the Pliensbachian – Toarcian of Mochras (Fig 1a) follow the same trends but show lower isotopic variability than in the Cleveland Basin (Ullmann et al., Reference Ullmann, Szȕcs, Jiang, Hudson and Hesselbo2022), but data are lacking for the T-OAE interval. δ18Obel and Mg/Cabel values from the Junction Bed in Dorset indicate significant warming during the Toarcian hyperthermal (Jenkyns & Macfarlane, Reference Jenkyns and Macfarlane2021). δ18O data from marine calcite fossils from a range of other European basins show a consistent shift to very light values accompanying the T-OAE (e.g. Jenkyns et al., Reference Jenkyns, Jones, Gröcke, Hesselbo and Parkinson2002; Dera et al., Reference Dera, Brigaud, Monna, Laffont, Puceat, Deconinck, Pellenard, Joachimski and Durlet2011; Korte et al., Reference Korte, Hesselbo, Ullmann, Dietl, Ruhl, Schweigert and Thibault2015, fig. 2). Significantly, δ18Obel and Mg/Cabel data from open basin settings in central and northern Spain with no evidence of salinity changes, indicate a 5 – 8° C temperature rise in the western Tethys during the T-OAE (Rosales et al., Reference Rosales, Robles and Quesada2004; Gómez et al., Reference Gómez, Goy and Canales2008; Comas-Rengifo et al., Reference Comas-Rengifo, Arias, Gómez, Goy, Herrero, Osete and Palencia2010).

Oxygen isotope data from calcite shells of benthic fauna (rhynchonellid brachiopods and Gryphaea bivalves) from Spain and Portugal suggest that bottom-water temperatures were elevated by ∼3.5° C through the entire T-OAE (Ullmann et al., Reference Ullmann, Boyle, Duarte, Hesselbo, Kasemann, Klein, Lenton, Piazza and Aberhan2020) compared to the underlying Toarcian. Based on results from Peniche (Fig. 1a), temperatures increased by ∼7° C between the latest Pliensbachian to the T-OAE maximum with a brief episode of temporary cooling of 4 – 5° C accompanying the PlToBE (Suan et al., Reference Suan, Mattioli, Pittet, Mailliot and Lecuyer2008a).

Marine fossil-based δ18O water temperature estimates from Europe are consistent with global carbon-isotope mass-balance modelling that indicates a 3 – 5° C increase in global land surface temperature during the T-OAE (Beerling et al., Reference Beerling, Lomas and Gröcke2002). Cooling followed by rapid global warming of ∼6.5° C during the T-OAE has also been inferred from stomatal index analysis of fossil leaves from Bornholm, Denmark (McElwain et al., Reference McElwain, Wade-Murphy and Hesselbo2005). The early Toarcian hyperthermal event appears synchronous throughout Europe within the limitations of sampling, when calibrated against δ13C profiles.

The TEX86 proxy, derived from fossilized archaeal lipids, was employed by Ruebsam et al. (Reference Ruebsam, Reolid, Sabatino, Masetti and Schwark2020c) to reconstruct absolute sea surface temperatures (SSTs) for the NW Tethys Shelf during the late Pliensbachian to early Toarcian based on the concatenation of data from Spain (southern Iberian margin) and Italy (Umbria-Marche Basin). The overall trend of the molecular palaeothermometry data covaries with δ18O signatures derived from brachiopods and belemnites, evidencing coupled warming of surface and bottom waters. Reconstructed late Pliensbachian to early Toarcian SSTs of 22° – 32° C in NW Tethys are comparable to current tropical SSTs (Locarnini et al., Reference Locarnini, Mishonov, Baranova, Boyer, Zweng, Garcia, Reagan, Seidov, Weathers, Paver and Smolyar2019). A long-term rise of ∼10° C was documented through the earliest Toarcian, with maximum SSTs attained during the T-OAE, but with the time series incorporating transient temperature excursions of >5° C magnitude with lapse rates of ∼0.1° C/kyr. These transient events were interpreted as evidencing the existence of a labile cryosphere driving rapid climate fluctuations and providing a reservoir that facilitated pronounced δ13C excursions (Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019, Reference Ruebsam, Reolid, Sabatino, Masetti and Schwark2020c).

19. The fossil record – biotic response to the T-OAE

The Pliensbachian–Toarcian boundary succession in the Cleveland Basin records a major extinction event that is observable throughout Western Europe at species and genus levels affecting benthic foraminifera, ostracods, bivalves, rhynchonellid brachiopods and corals, and marks a turnover in ammonites and belemnites (Hesselbo et al., Reference Hesselbo, Ogg, Ruhl, Hinnov, Huang, Gradstein, Ogg, Schmitz and Ogg2020b; Vasseur et al., Reference Vasseur, Lathuiliére, Lazăr, Martindale, Bodin and Durlet2021). The multi-phase extinctions in the earliest Toarcian are a global event that extended across the Boreal, Tethyan and Panthalassa oceans (Caruthers et al., Reference Caruthers, Smith and Gröcke2013); this is an expression of the second-order ‘Pliensbachian’ global extinction event recognized by Raup & Sepkoski (Reference Raup and Sepkoski1984; Bambach, Reference Bambach2006).

The base of the Toarcian in Yorkshire is marked by the extinction of Boreal amaltheid family ammonites and a surge of dactylioceratids and hildoceratids of Tethyan derivation (e.g. Dera et al., Reference Dera, Neige, Dommergues, Fara, Laffont and Pellenard2010). The base-Toarcian extinction was followed by a second pronounced extinction among benthic invertebrates, particularly bivalves and brachiopods, that coincides with the onset of the T-OAE and black shale deposition in northern Europe but is not restricted to anoxic facies (Wignall & Bond, Reference Wignall and Bond2008). Among the bivalves, 84% of species became extinct in Europe with the onset of black shale deposition, which represents by far the most important extinction event of the whole Jurassic (Hallam, Reference Hallam1986). Comparable extinctions of bivalves occurred during the early Toarcian in the Andean basins of South America (Aberhan & Baumiller, Reference Aberhan and Baumiller2003).

19.a. Macrofauna and trace fossils

The Yorkshire coastal succession provides one of the most detailed published palaeontological records of the upper Pliensbachian – middle Toarcian. Howarth (Reference Howarth1955, Reference Howarth1962, Reference Howarth1973, Reference Howarth1992) published comprehensive ammonite and associated faunal records based on bed-by-bed collecting of the coastal sections. Further macrofossil records and a species-level range chart for the Cleveland Basin were produced by Little (Reference Little1995) and presented by Little (Reference Little, Ryder, Fastovsky and Gartner1996), Little & Benton (Reference Little and Benton1995), Harries & Little (Reference Harries and Little1999) and, in part, by Wignall et al. (Reference Wignall, Newton and Little2005). Additional records and a revised range chart were published by Caswell et al. (Reference Caswell, Coe and Cohen2009), with further collection reported by Danise et al. (Reference Danise, Twitchett, Little and Clémence2013) and Atkinson et al. (Reference Atkinson, Little and Dunhill2023). The trace-fossil record has been described by Morris (Reference Morris1979), Martin (Reference Martin and McIlroy2004), Caswell & Frid (Reference Caswell and Frid2017), Caswell & Dawn (Reference Caswell and Dawn2019) and Caswell & Herringshaw (Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023). The last of these studies included data from the Dove’s Nest core.

The upper Pliensbachian of the Cleveland Basin (Unit I - Staithes Sandstone and Cleveland Ironstone formations) yields rich trace-fossil and high-diversity benthic body-fossil assemblages, with most species being evenly represented and including a wide range of ecological groups that occupy all levels of benthic tiering (Little, Reference Little1995; Caswell et al., Reference Caswell, Coe and Cohen2009; Danise et al., Reference Danise, Twitchett, Little and Clémence2013; Caswell & Frid, Reference Caswell and Frid2017; Atkinson et al., Reference Atkinson, Little and Dunhill2023; Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023), indicative of well-oxygenated conditions.

The extinction of Amaltheidae family ammonites marking the base of the Toarcian occurs at Sulphur Band 1 (Littler et al., Reference Littler, Hesselbo and Jenkyns2010), a short distance above the base of the Whitby Mudstone Formation (Sections 7.b, 8.b). This is followed by a sharp drop in the taxonomic richness of trace fossils (Fig. 22; Caswell & Frid, Reference Caswell and Frid2017; Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023) but with no clear decline in benthic macrofossil diversity (Danise et al., Reference Danise, Twitchett, Little and Clémence2013; Caswell & Frid, Reference Caswell and Frid2017; Atkinson et al., Reference Atkinson, Little and Dunhill2023). A first extinction horizon (level i, Fig. 22; extinction step 2 of Harries & Little, Reference Harries and Little1999; ii of Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023), which is followed by an influx of the low-oxygen specialist bivalve Pseudomytiloides dubius (Sowerby), was placed at the base of Sulphur Band 2 (Hawsker Bottoms ‘bed’ 2) by Caswell et al. (Reference Caswell, Coe and Cohen2009).

Figure 22. TOC, biotic trends, extinction levels and palaeoredox change through the upper Pliensbachian – middle Toarcian of the Cleveland Basin. Yorkshire coast stratigraphy as in Figure 2. Rescaled whole-rock TOC profile for Dove’s Nest (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022, dark green) with coast composite data of Kemp et al. (Reference Kemp, Coe, Cohen and Weedon2011; thin yellow-green high-resolution curve) and trend of the low-resolution coast datasets of Ruvalcaba Baroni et al. (Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018) and McArthur (Reference McArthur2019) (thin pale green low-resolution curve; see Fig. 2). Ranges of low-oxygen specialist bivalve taxa compiled from Little (Reference Little1995) and Caswell et al. (Reference Caswell, Coe and Cohen2009). Trace-fossil taxonomic richness from Caswell & Frid (Reference Caswell and Frid2017), Caswell & Dawn (Reference Caswell and Dawn2019) and Caswell & Herringshaw (Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023). Macrofaunal diversity after Danise et al. (Reference Danise, Twitchett, Little and Clémence2013, Reference Danise, Twitchett and Little2015). Profiles rescaled to reference stratigraphy (Fig. 2) using subzone and marker bed datum levels. Extinction levels (i) – (iii) from Caswell et al. (Reference Caswell, Coe and Cohen2009). The boundary between pre-extinction and post-extinction survival intervals of Atkinson et al. (Reference Atkinson, Little and Dunhill2023) lies at the base of the C. exaratum Subzone (level iii). The top of the ‘survival interval’, the base of recovery phase 1, occurs in the lower H. bifrons Zone above the top of our study interval. Geochemical palaeoredox interpretations from this study (see text). Climate interpretation incorporates palynological interpretation of Slater et al. (Reference Slater, Twitchett, Danise and Vajda2019) with oxygen isotope and Mg/Ca trends from belemnites (Fig. 21).

The lower Toarcian Grey Shale (Subunit IIa) initially yields a sparse normal marine benthic and nektonic fauna reflecting continuing sedimentation in a relatively well-oxygenated environment but with increasingly dysoxic, intermittently anoxic (Sulphur Bands 1 – 3) seafloor conditions indicated by reduced species richness and the changing characteristics of trace-fossil assemblages. An interval of greater oxygen depletion is evidenced in the lower part of Subunit IIb, following step increases in TOC and PEF, by an interval barren of trace fossils and characterized by low body-fossil diversity (Fig. 22; see also Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023). Geochemical redox proxies (e.g. Figs 15, 16) generally fail to capture evidence of this event. Trace-fossil and benthic diversity temporarily recover towards the top of Subunit IIb. Disappearances of typical upper Pliensbachian benthic macrofossil taxa then occur progressively, beginning towards the top of the Grey Shale within T-OAE Subunit IIIa through ‘beds’ 31 – 32 (Caswell et al., Reference Caswell, Coe and Cohen2009, fig. 2).

The onset of a major phase of extinction is placed at the base of the consistently laminated black carbonaceous mudstones (black shale facies) in mid-‘bed’ 31 (level ii of Caswell et al., Reference Caswell, Coe and Cohen2009), with the main extinction horizon (level iii) occurring around 2 m higher, coincident with the first large step-fall in the δ13Corg profile (upper ‘bed’ 32, A in Fig. 18; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005). Trace fossils disappear at level ii and, with the sole exception of a record of pyritized Trichichnus in ‘bed’ 48 at Saltwick Bay, remain unrecorded from all Yorkshire sections until the reappearance of an impoverished assemblage in the middle Toarcian Main Alum Shale Beds (Subunit Vc), towards the base of ‘bed’ 51 (Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023). This stratigraphically extended absence of macroscopic trace fossils indicates a longer period of continuous deoxygenation than elsewhere in NW Tethys and the adjacent European shelf basins.

A steady decline in macrofaunal diversity is observed between datum levels ii and iii (Fig. 22; Danise et al., Reference Danise, Twitchett, Little and Clémence2013). The bivalve Bositra radiata (Goldfuss) dominates the benthic fauna of the extinction interval, with shell size doubling as δ13Corg decreases within Subunit IIIa. Pseudomytiloides dubius is also common at some levels. Pseudomytiloides dubius and B. radiata occur as parautochthonous – autochthonous assemblages, respectively, forming 1 – 2-shell-thick monospecific shell pavements (Little, Reference Little1995; Caswell et al., Reference Caswell, Coe and Cohen2009; Caswell & Coe, Reference Caswell and Coe2013). The two low-oxygen specialist taxa rarely co-occur.

Pseudomytiloides dubius is the only abundant and consistently occurring benthic macrofossil in the laminated black shales of Subunits IIIb – Va inclusive (C. exaratum – H. falciferum subzones), evidencing the development of strongly oxygen-depleted bottom waters and anoxic – euxinic sediments within the lower part of the ‘survival interval’, following the extinctions (Fig. 22). Sporadic occurrences of the scallop Meleagrinella substriata (Munster) is the only other taxon to occur consistently through the Jet Rock (Caswell et al., Reference Caswell, Coe and Cohen2009).

Bositra radiata temporarily reappears in T-OAE Subunit IIId immediately following the reappearance of Oxytoma inequivalve (J. Sowerby) along with the lowest occurrences and temporary appearances of new bivalve species and inarticulate brachiopods towards the top of Subunit IIIc (Caswell et al., Reference Caswell, Coe and Cohen2009). These define a short interval of increased benthic diversity (Fig. 22) in the upper C. exaratum Subzone at the termination of the T-OAE. The temporary influx of a more diverse faunal assemblage coincides with a shift to intermittent dysoxic bottom waters based on the TOC/PT and DOPT proxies (Fig. 15) and decreasing basin restriction based on Re and Mo isotopes (Fig. 20).

Remarkably, records of P. dubius almost perfectly track the TOC curve, only occurring where values exceed the 2.5% anoxic threshold (Fig. 22). The species is largely absent from the upper Bituminous Shales Subunit Va but reappears in association with the higher-TOC beds that characterize the Hard Shales Subunit Vb, before disappearing again above. Benthic diversity increases significantly in the lower part of Subunit Va and generally remains high in Subunits Vb – c, above. Increases in benthic diversity and the re-establishment of a Toarcian infauna after the T-OAE were gradual. Diversity does not return to the level observed in the pre-T-OAE section until the upper Toarcian Phlyseogrammoceras dispansum Zone (Atkinson et al., Reference Atkinson, Little and Dunhill2023). The absence and then low abundance and taxonomic richness of trace fossils in the D. commune Subzone suggest that dissolved oxygen remained below 0.5 mL L−1 into the middle Toarcian (Caswell & Dawn, Reference Caswell and Dawn2019; Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023), consistent with evidence of persisting dysoxia provided by the TOC/PT and DOPT redox proxies (Fig. 14).

Ammonites and belemnites occur throughout the upper Pliensbachian – middle Toarcian succession in Yorkshire, including the T-OAE interval. A ‘belemnite gap’ was proposed in T-OAE Subunits IIIa – IIIb from mid-‘bed’ 32 to lower ‘bed’ 34 in Yorkshire coastal sections by Xu et al. (Reference Xu, Ruhl, Jenkyns, Leng, Huggett, Minisini, Ullmann, Riding, Weijers, Storm, Percival, Tosca, Idiz, Tegelaar and Hesselbo2018) based on literature data. However, McArthur et al. (Reference McArthur, Donovan, Thirlwall, Fouke and Mattey2000) recorded belemnites from every ‘bed’ in T-OAE Unit III except ‘beds’ 33 (Cannon Ball Doggers) and 40 (Millstones). The ‘belemnite gap’ in the Cleveland Basin coincides with a turnover in belemnite taxa characterized by the disappearance of Passaloteuthis and the appearance of Acrocoelites in the Cleveland Basin (Ullmann et al., Reference Ullmann, Thibault, Ruhl, Hesselbo and Korte2014), attributed as a response to the progressive expansion and shoaling of euxinic bottom waters during the initiation of the T-OAE. Remarkably, Acrocoelites are relatively common in the upper beds of the T-OAE interval (mid- to upper C. exaratum Subzone, e.g. Ullman et al., Reference Ullmann, Thibault, Ruhl, Hesselbo and Korte2014, fig. 2) despite evidence of increasing episodes of PZE (Section 15b). It has been inferred that this turnover represents belemnites adapting to environmental change by shifting their habitat from cold bottom waters to warm surface waters in response to expanded seafloor anoxia.

Stratigraphically more extended ‘belemnite gaps’ but continuous ammonite records have been documented in the lower Toarcian at Mochras and Peniche, which were attributed to unfavourable redox conditions in the deeper water column (Hesselbo et al., Reference Hesselbo, Jenkyns, Duarte and Oliveira2007; Xu et al., Reference Xu, Ruhl, Jenkyns, Leng, Huggett, Minisini, Ullmann, Riding, Weijers, Storm, Percival, Tosca, Idiz, Tegelaar and Hesselbo2018, fig. 4), reflecting the nektobenthic lifestyle of the former taxon (Hoffmann & Stevens, Reference Hoffmann and Stevens2020). However, neither of these sections show palaeontological or geochemical evidence of extended anoxic – euxinic bottom-water conditions during the T-OAE. Bioturbation and low-diversity benthonic macrofauna are present throughout the sections, sediment TOC contents generally remain significantly below the 2.5% anoxic threshold and redox-sensitive trace metals show no enrichment (e.g. Fantasia et al., Reference Fantasia, Adatte, Spangenberg, Font, Duarte and Föllmi2019). Rare thin laminated beds with biomarkers that suggest transient PZE and episodic anoxia at the seafloor occurred (Xu et al., Reference Xu, Ruhl, Jenkyns, Leng, Huggett, Minisini, Ullmann, Riding, Weijers, Storm, Percival, Tosca, Idiz, Tegelaar and Hesselbo2018), like those associated with the lower Toarcian Sulphur Bands of Yorkshire, but the evidence supports predominantly oxic – dysoxic bottom-water conditions throughout the T-OAE. It is unlikely, therefore, that water column oxygenation provided the main control on belemnite occurrences at Mochras and Peniche, given that belemnites remain present in association with bottom-water euxinia in the Cleveland Basin.

Water temperature and salinity may have played a role. Belemnites are considered to favour temperatures of 10 – 30° C and salinities of 27 – 37 psu (Hoffmann & Stevens, Reference Hoffmann and Stevens2020), and have been postulated to have migrated into warm surface waters during the T-OAE in response to deep water anoxia (Ullmann et al., Reference Ullmann, Thibault, Ruhl, Hesselbo and Korte2014). Ruebsam et al. (Reference Ruebsam, Pienkowski and Schwark2020) reported tropical SSTs of 32° C on the NW Tethyan shelf during peak hyperthermal conditions at the height of the T-OAE, with southern areas dominated by a warm equatorial Tethyan current. More northerly areas were influenced by cool lower salinity Boreal arctic surface waters (Section 18).

If water temperatures, and perhaps salinity, in southern areas, exceeded the tolerance of belemnites, they would have been temporarily excluded until temperatures declined following the climate optimum. It is notable that brachiopods are also temporarily absent from the T-OAE interval of many Tethyan sites, despite evidence for the continual presence of oxygenated bottom waters (García Joral et al., Reference García Joral, Gómez and Goy2011; Danise et al., Reference Danise, Clemence, Price, Murphy, Gomez and Twitchett2019; Ullmann et al., Reference Ullmann, Boyle, Duarte, Hesselbo, Kasemann, Klein, Lenton, Piazza and Aberhan2020). The northerly location of the Cleveland Basin would have seen a greater influence of cool currents flowing through the Viking Corridor (Fig. 1), enabling belemnites to maintain their presence in the area’s surface waters throughout the early Toarcian hyperthermal, while being excluded from the anoxic – euxinic bottom waters that developed during the T-OAE.

It is significant that a marked increase in nekton diversity occurs immediately above the T-OAE interval in Yorkshire (Fig. 22, Unit IVa) with an influx of ammonites indicative of a better oxygenated upper water column and/or improved connection with open ocean water masses.

19.b. Plankton records

19.b.1. Calcareous nannofossils

In stark contrast to the macrofauna, the late Pliensbachian – early Toarcian, including the T-OAE, was not marked by a major extinction event in phytoplankton but was, nonetheless, a significant time for calcareous nannofossil evolution. Major speciation took place and some of the most common Jurassic and Cretaceous genera (Biscutum, Lotharingius, Discorhabdus, Watznaueria) appeared and rapidly evolved (Bown et al., Reference Bown, Lees, Young, Thierstein and Young2004).

Increased biological productivity with increased calcareous nannofossil speciation, preceding the T-OAE, have been documented in north Tethyan and north European successions in the upper Pliensbachian – basal Toarcian (e.g. Erba, Reference Erba2004; Erba et al., Reference Erba, Bottini, Faucher, Gambacorta and Visentin2019; Menini et al., Reference Menini, Mattioli, Hesselbo, Ruhl, Suan, Reolid, Duarte, Mattioli and Ruebsam2021). A decrease in the abundance of Schizosphaerella and Mitrolithus jansae (Wiegand) together with a fall in the size of Schizosphaerella occurs around the Pliensbachian – Toarcian boundary in Portugal, Spain, France and Italy (Suan et al., Reference Suan, Mattioli, Pittet, Mailliot and Lecuyer2008a, 2010; Reolid et al., Reference Reolid, Emanuela, Nieto and Rodríguez-Tovar2014; Clémence et al., Reference Clémence, Gardin and Bartolini2015; Peti & Thibault, Reference Peti and Thibault2017; Erba et al., Reference Erba, Bottini, Faucher, Gambacorta and Visentin2019; Fraguas et al., Reference Fraguas, Gómez, Goy, Comas-Rengifo, Reolid, Duarte, Mattioli and Ruebsam2021; Visentin & Erba, Reference Visentin and Erba2021; Faucher et al., Reference Faucher, Visentin, Gambacorta and Erba2022). A more pronounced decline in Schizosphaerella size accompanies the onset of the T-OAE, the ‘Schizosphaerella crisis’ (Visentin & Erba, Reference Visentin and Erba2021; Faucher et al., Reference Faucher, Visentin, Gambacorta and Erba2022), with a small size persisting up towards the top of the T-OAE interval and highest occurrence of M. jansae. These changes have variously been attributed to global warming, ocean fertilization and possibly acidification leading to a dominance of opportunistic taxa.

A decline in calcareous nannofossil abundance and a sharp size decrease of Schizosphaerella spp. have been observed at the onset of the T-OAE CIE in the Mochras borehole of the Cardigan Bay Basin (Menini et al., Reference Menini, Mattioli, Hesselbo, Ruhl, Suan, Reolid, Duarte, Mattioli and Ruebsam2021), with a size peak in the recovery phase of the CIE. Samples of T-OAE interval mudstones from the Brown Moor borehole in the southern Cleveland Basin studied by Bucefalo Palliani et al. (Reference Bucefalo Palliani, Mattioli and Riding2002) were recorded as being barren of calcareous nannofossils from the high-TOC interval of the upper D. semicelatum – lower C. exaratum subzones. Calcareous nannofossils reappeared and increased in absolute abundance and species richness with falling TOC values through the top C. exaratum and H. falciferum subzones. Rare occurrences of Schizosphaerella were recorded in these upper beds.

Abundant coccoliths were reported in the laminated limestone of the Top Jet Dogger ‘bed’ 39 from the Yorkshire coast by Sælen et al. (Reference Sælen, Doyle and Talbot1996), who noted low abundances in the adjacent beds. The presence of coccolith-rich pellets in the Jet Rock was illustrated by Macquaker et al. (Reference Macquaker, Keller and Davies2010). However, more generally, samples taken through the upper Pliensbachian – middle Toarcian of the coastal succession have been reported to be barren or yield very rare to rare coccolith body fossils (e.g. Slater et al., Reference Slater, Bown, Twitchett, Danise and Vajda2022).

New work has demonstrated that although Toarcian black shale facies generally yield few calcareous nannofossils, or are barren, this is a preservation artefact. ‘Ghost’ nannofossils, including Schizosphaerella, preserved as external moulds in amorphous organic matter are abundant throughout the lower Toarcian of the Yorkshire coast and Japan (Slater et al., Reference Slater, Bown, Twitchett, Danise and Vajda2022). Near monospecific assemblages of nannofossil moulds or prasinophytes (principally sphaeromorphs, cf. Slater et al., Reference Slater, Twitchett, Danise and Vajda2019) during the T-OAE in Yorkshire likely represent persistent algal blooms. The absence of carbonate is a product of early diagenetic dissolution by acidic pore waters or, for coastal exposures, weathering (cf. Littke et al., Reference Littke, Klussmann, Krooss and Leythaeuser1991; Brantley et al., Reference Brantley, Holleran, Jin and Bazilevskaya2013). Future study of calcareous nannofossils in the Dove’s Nest core would enable the impact of weathering to be assessed.

At a community level, the ‘ghost’ records show that contrary to the conclusions of a previous study in the Cleveland Basin (Bucefalo Palliani et al., Reference Bucefalo Palliani, Mattioli and Riding2002), calcareous nannoplankton flourished throughout the T-OAE with increased speciation rates and no evidence of increased extinctions. This challenges the view that ocean acidification (e.g. Hönisch et al., Reference Hönisch, Ridgwell, Schmidt, Thomas, Gibbs, Sluijs, Zeebe, Kump, Martindale, Greene, Kiessling, Ries, Zachos, Royer, Barker, Marchitto, Moyer, Pelejero, Ziveri, Foster and Williams2012; Erba et al., Reference Erba, Bottini, Faucher, Gambacorta and Visentin2019) played a dominant role in driving the biotic changes associated with the T-OAE. Boron-isotope (δ11B) results from Peniche reported as showing declining ocean pH in the early Toarcian (Müller et al., Reference Müller, Jurikova, Gutjahr, Tomasovych, Schlogl, Liebetrau, Duarte, Milovsky, Suan, Mattioli, Pittet and Eisenhauer2020) is not supported by Ca and Sr isotope data (Q Li et al., Reference Li, McArthur, Thirlwall, Turchyn, Page, Bradbury, Weis and Lowry2021).

19.b.2. Organic-walled phytoplankton

The palynology of the upper Pliensbachian – middle Toarcian of the Yorkshire coast was described by Slater et al. (Reference Slater, Twitchett, Danise and Vajda2019). Abundances of organic-walled dinoflagellate cysts (dinocysts) and spiny acritarchs decrease significantly at the base of the Toarcian, then decline markedly at the base of the T-OAE interval. Here, they are replaced by a substantial increase in sphaeromorphs, interpreted as the prasinophyte algae Halosphaeropsis liassica, a marker taxon for the T-OAE. These co-occur with abundant amorphous organic matter considered to represent euxinic marine conditions. The prasinophyte Tasmanites also increases in abundance at the base of the T-OAE section. Similar trends were documented in the Brown Moor borehole by Bucefalo Palliani et al. (Reference Bucefalo Palliani, Mattioli and Riding2002). Short-term increases in sphaeromorphs with large spikes and amorphous organic matter content occur also in the Sulphur Bands (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019).

The algal dominance and associated low abundance of dinocysts continue upwards through the lower Bituminous Shales (Unit IV), decline at the top of the Bituminous Shales (Subunit Va) and increase again in the Hard Shales (Subunit Vb) before declining above. Dinocysts only return to pre-event abundances in the main Alum Shale beds (Subunit Vc) demonstrating the persistence of algal-dominated marine plankton in the Cleveland Basin from the onset of the T-OAE in the D. semicelatum Subzone to the D. commune Subzone of early middle Toarcian.

A marine phytoplankton community structure dominated by eukaryotic algae within the interval of the δ13Corg minimum (our Subunit IIIb) in the Paris Basin is evidenced by biomarker data (Section 15.b.; Ruebsam et al., Reference Ruebsam, Mattioli and Schwark2022). Here, a proliferous of green algae and a decline in red algae groups accompanied five surface-water freshening events during the T-OAE. These were interpreted to represent recurrent warming phases, paced by changes in the short eccentricity (100 kyr) orbital cycle, that initiated higher precipitation and increased surface run-off (Ruebsam et al., Reference Ruebsam, Mattioli and Schwark2022).

19.b.3. Geochemical and biotic interaction

Redox-sensitive trace metals, including Mo, V and Cu, are important micronutrients that are necessary for several metabolic processes, including photosynthesis. Metals modulate the growth of organisms and their cycling of major nutrients, including C and N (Morel & Price, Reference Morel and Price2003). Trace metal availability can significantly influence ecosystem structure: a healthy and fully functioning phytoplankton-driven ecosystem is directly controlled by nutrient availability in the oxygenated upper ocean.

The drawdown of trace metals, especially Mo, during the T-OAE would potentially impact the basal ecology of the oceans (Owens et al., Reference Owens, Reinhard, Rohrssen, Love and Lyons2016; Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017; Them et al., Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022). A larger impact on eukaryotes would be expected due to their higher trace metal metabolic requirements (Liu et al., Reference Liu, Ji, Hu, Xia, Yi, Them, Sun and Chen2021). It is notable that organic geochemical studies of the T-OAE intervals on the NE western Tethyan shelf (Hungary; Ruebsam et al., Reference Ruebsam, Muller, Kovacs, Palfy and Schwark2018) and in eastern Tethys (Tibet; Liu et al., Reference Liu, Ji, Hu, Xia, Yi, Them, Sun and Chen2021) indicate a transition to bacterial-dominated microbial ecosystems coincident with the change to low-diversity macrofaunal assemblages.

The interval of rising ε205Tl in western Canada and the SW German Basin during the early Toarcian with peak values at the onset of the T-OAE (Them et al., Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018), considered to represent a minimum in global oxic seafloor area, corresponds to the interval of trace-fossil decline and macrofossil stepped extinctions recorded in the Cleveland Basin and elsewhere (Fig. 22). The extinctions may potentially, therefore, be related to the global spread of anoxic bottom waters and more general oxygen depletion of the oceans. The slow biotic recovery that followed in the middle Toarcian, which did not accelerate until the late Toarcian (Atkinson et al., Reference Atkinson, Little and Dunhill2023), might have been a product of persisting poor ocean oxygenation, as indicated by continuing high ε205Tl values.

20. T-OAE – a Jurassic global anoxic event?

The concept of an oceanic anoxic event was introduced by Schlanger & Jenkyns (Reference Schlanger and Jenkyns1976) to describe the global development of Cretaceous organic-rich pelagic sediments formed in a variety of palaeobathymetric settings, including oceanic plateaus and basins sampled by ocean drilling, continental margins and shelf seas. This was attributed to a coincidence of high sea level and warm climate that favoured the formation of an expanded oxygen-minimum layer and where this intersected the sediment-water interface, the deposition of organic carbon-rich deposits. It was recognized that although an OAE was global in nature, organic-rich sediments would show a widespread but irregular distribution.

Despite organic carbon being a key sedimentary signature of an OAE, even the most intense Cretaceous event, OAE2 at the Cenomanian – Turonian boundary, the best documented and perhaps the most widespread marine organic carbon burial event in Earth history (e.g. Reershemius & Planavsky, Reference Reershemius and Planavsky2021), is not represented by organic enrichment in all sedimentary settings (Jenkyns, Reference Jenkyns1980; Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Tuenter, Henstra, van der Zwan, van de Wal, Dijkstra and de Boer2010; Owens et al., Reference Owens, Lyons and Lowery2018, fig. 2; Zhai et al., Reference Zhai, Zeng, Zhang and Yao2023). The associated positive carbon-isotope excursion (cf. Scholle & Arthur, Reference Scholle and Arthur1980; Schlanger et al., Reference Schlanger, Arthur, Jenkyns, Scholle, Brooks and Fleet1987), however, provides a reliable stratigraphic marker observable in all carbon phases (e.g. Jenkyns, Reference Jenkyns2010).

In contrast to Cretaceous OAEs, the general absence of Lower Jurassic oceanic crust and pelagic sediments prevents a physical assessment of the global extent of anoxic waters and black shale deposition associated with the T-OAE in the deep sea. Marine records are mostly limited to shallow epicontinental seas, shelves and marginal basins. Extensive black shale deposits are a defining feature of most north European Boreal lower Toarcian sections but the majority of southern European Tethyan sites are organic lean (e.g. Remirez & Algeo, Reference Remirez and Algeo2020a, fig. 3; Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b, fig. 2). Based on the lack of evidence for globally extensive anoxia in the early Toarcian and evidence of hydrographic control on anoxia in northern Europe, McArthur (Reference McArthur2019) proposed abandoning the term ‘OAE’ for the Toarcian event.

Most sites globally show an increase in TOC with the T-OAE, defined by the interval of the negative CIE, relative to pre-event lower Toarcian values (Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b), even on shallow-water carbonate platforms (e.g. Ettinger et al., Reference Ettinger, Larson, Kerans, Thibodeau, Hattori, Kacur and Martindale2021). TOC enrichment is best displayed in sections dominated by marine rather than terrestrial organic matter. Furthermore, although stratigraphic control is commonly poor, there is evidence for lower Toarcian organic-rich facies in many areas outside Europe – North Africa, including western and Arctic Canada, Arctic Siberia, Madagascar and Australia (Jenkyns, Reference Jenkyns1988; Nikitenko & Mickey, Reference Nikitenko and Mickey2004; Them et al., Reference Them, Gill, Caruthers, Gröcke, Tulsky, Martindale, Poulton and Smith2017a, Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018, Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022).

Deep-water black shales and cherts attributed to the T-OAE are recorded from Panthalassa Ocean sites including Japan and Kamchatka Russia (Gröcke et al., Reference Gröcke, Hori, Trabucho-Alexandre, Kemp and Schwark2011; Ikeda & Hori, Reference Ikeda and Hori2014; Ikeda et al., Reference Ikeda, Hori, Ikehara, Miyashita, Chino and Yamada2018; Filatova et al., Reference Filatova, Konstantinovskaya and Vishnevskaya2022; Kemp et al., Reference Kemp, Chen, Cho, Algeo, Shen and Ikeda2022a) although, again, not all Panthalassa sites display anoxia (e.g. Kemp & Izumi, Reference Kemp and Izumi2014; Fantasia et al., Reference Fantasia, Föllmi, Adatte, Bernárdez, Spangenberg and Mattioli2018a). Furthermore, deep-water black shales deposited during Cretaceous OAEs have been shown to be redeposited shallow-water deposits (Dean et al., Reference Dean, Arthur and Stow1984; Trabucho-Alexandre et al., Reference Trabucho-Alexandre, van Gilst, Rodríguez-López and de Boer2011). Nonetheless, there is strong evidence that the T-OAE was globally distributed and widely associated with enhanced organic matter deposition.

Geochemical proxies offer unique insights into global-scale processes and their impact during the T-OAE. In oxic environments, organic-lean sediments (TOC <1%) show a positive correlation between TOC and burial rate (e.g. Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b and references therein). This relationship is lost in dysoxic – anoxic settings where oxygen depletion promotes enhanced preservation of organic matter on the seafloor and during early burial, producing higher sediment TOC contents (1 – 2.5%). However, the very high TOC contents (>5%) observed for sediments deposited in the basins of NW Europe and elsewhere during the early Toarcian can be attributed principally to low sedimentation rates (Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b) caused by reduced siliciclastic and biogenic (carbonate and silica) sediment fluxes rather than solely bottom-water anoxia – euxinia.

Based on a review of global TOC data for the lower Toarcian, Kemp et al. (Reference Kemp, Suan, Fantasia, Jin and Chen2022b) offered a conservative estimate that the amount of extra marine organic carbon buried in shallow seas during the T-OAE relative to the preceding early Toarcian was ∼9000 Gt. Surprisingly, calculated rates of organic carbon burial during the T-OAE are low relative to shallow-water settings at the present day. Kemp et al. (Reference Kemp, Suan, Fantasia, Jin and Chen2022b) used a duration of 900 ka for the T-OAE interval compared with ∼450 ka estimated by others (Section 17) but the shorter duration does not affect the conclusions. The modern global median continental shelf organic carbon burial rate of 19.6 g C m−2 a−1 (Wilkinson et al., Reference Wilkinson, Besterman, Buelo, Gephart and Pace2018) is more than an order of magnitude higher than the calculated mean T-OAE rates (0.6 – 1.6 g C m−2 a−1) and the average of 0.89 g C m−2 a−1 derived for northern Europe (Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b).

Documented TOC increases are generally highest where deoxygenation was most severe: carbon burial rates in anoxic – euxinic basins may have increased ∼500% on average during the T-OAE, potentially sequestering an extra ∼800 Gt relative to the same time interval immediately preceding the event (Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b). Nonetheless, less than 10% of the excess organic carbon in shallow-marine settings was buried in the anoxic – euxinic basins of northern Europe.

Deep open-ocean carbon burial fluxes are very poorly constrained, and it can be speculated that the deep ocean may have accounted for a significant fraction of the total amount of carbon buried during the T-OAE. Similar issues arise when considering the carbon burial budget during OAE2 where a small but significant increase in deep ocean OC burial offers one potential explanation for the ‘missing’ carbon (Owens et al., Reference Owens, Lyons and Lowery2018). It is worth noting that Ikeda et al. (Reference Ikeda, Hori, Ikehara, Miyashita, Chino and Yamada2018) reported TOC contents of ≥30% in lower Toarcian black bedded cherts from Inuyama, Japan, considered to represent a deep-sea central Panthalassa Ocean setting during the T-OAE.

Molybdenum and TOC data from a deep-water Panthalassa chert succession in Japan evidence a globally significant drawdown of seawater Mo accompanying the T-OAE (Section 13.a; Kemp et al., Reference Kemp, Chen, Cho, Algeo, Shen and Ikeda2022a), consistent with and Mo- and Tl-isotope data that demonstrate a substantially increased global extent of anoxia (Sections 16.b.1, 16.b.2; Dickson, Reference Dickson2017; Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017; Them et al., Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018). A global increase in pyrite burial under oxygen-depleted conditions is also indicated by sulfur isotopes (Section 16.a.2; Gill et al., Reference Gill, Lyons and Jenkyns2011; Newton et al., Reference Newton, Reeves, Kafousia, Wignall, Bottrell and Sha2011).

Them et al. (Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022) used open-ocean Mo records from the Western Canada Sedimentary Basin, with comparisons to a global compilation of Pliensbachian – Toarcian Mo data, to estimate the amount of Mo buried globally, incorporating values for local sedimentation rates and global weathering rates during the T-OAE (Them et al., Reference Them, Gill, Selby, Gröcke, Friedman and Owens2017b, Reference Them, Gill, Caruthers, Gerhardt, Gröcke, Lyons, Marroquin, Nielsen, Alexandre and Owens2018). These data indicate that ∼41 Gt of Mo was buried during the T-OAE which, using Mo/TOC values derived from modern basins, derived a value of ∼244,000 Gt organic carbon buried globally at that time. This is more than one order of magnitude larger than that estimated to be buried in shallow-marine sediments by Kemp et al. (Reference Kemp, Suan, Fantasia, Jin and Chen2022b).

The global value of ∼244,000 Gt C burial derived by Them et al. (Reference Them, Owens, Marroquín, Caruthers, Trabucho-Alexandre and Gill2022) would require a minimum of 3% of the ocean floor to be covered by euxinic waters compared to <0.3% today (Algeo, Reference Algeo2004; Helly & Levin, Reference Helly and Levin2004; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Palastanga and Stomp2020). This implies that, despite their prominent black shale sections, European basins, with an estimated < 2000 Gt C buried during the T-OAE (Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b), made a very small contribution to driving the observed changes in the carbon cycle. Sulfur isotope data and modelling similarly indicate that European basins played only a minor role in sulfide and therefore organic carbon, sequestration during the T-OAE interval (Gill et al., Reference Gill, Lyons and Jenkyns2011). For comparison, increased lacustrine organic productivity during the T-OAE attributed to elevated fluvial nutrient supply, resulted in the burial of ∼460 Gt of organic carbon in the Sichuan Basin alone (Xu et al., Reference Xu, Ruhl, Jenkyns, Hesselbo, Riding, Selby, Naafs, Weijers, Pancost, Tegelaar and Idiz2017, Reference Xu, Weijers, Ruhl, Idiz, Jenkyns, Riding, Gorbanenko, Hesselbo, Reolid, Duarte, Mattioli and Ruebsam2021), one of three major Early Jurassic lake systems in China.

Based on a negative CIE of about −3‰, the total amount of carbon released during the T-OAE was estimated to be ∼10,000 Gt by Ruebsam et al. (Reference Ruebsam, Mayer and Schwark2019) assuming a biogenic methane source of carbon (δ13C = −60‰ to −80‰), but >50,000 Gt would be required for a purely volcanic CO2 source (δ13C = −7‰ to −10‰). The release of 10,000 Gt C over a period of about 800 ka (see Section 17 for a discussion of the duration of the T-OAE, ∼1400 ka is an alternative estimate) was calculated to increase global temperatures by about 5° C and cause a two- to three-fold increase in pCO2 (Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019). For comparison, an increase in atmospheric pCO2 from ∼500 ppmv to ∼1000 ppmv during the T-OAE has been derived from reconstructions using the stomatal index and the difference in the magnitude of the CIE in marine and terrestrial environments (McElwain et al., Reference McElwain, Wade-Murphy and Hesselbo2005; Ruebsam et al., Reference Ruebsam, Reolid and Schwark2020d). A seawater temperature rise of 6 – 7° C is indicated for the Cleveland Basin (Section 18).

Ruebsam et al. (Reference Ruebsam, Mayer and Schwark2019) argued that a volcanic-driven modest rise in temperature in the early Toarcian triggered a melt-down of Earth’s cryosphere that had expanded during the icehouse episode of the late Pliensbachian, and the rapid release of greenhouse gases, mainly as 13C-depleted CH4, driving the negative δ13C excursion. However, the contribution of volcanic CO2 and the mass of carbon released during the T-OAE may have been significantly underestimated since a globally rapid silicate weathering response, surface ocean uptake and increased organic carbon burial would have consumed atmospheric CO2, preventing an atmospheric build-up and its climate impact.

21. Regional to global warming and anoxia

The Cleveland Basin provides a global reference for the T-OAE. An unprecedented range of existing high-quality, high-resolution sedimentological, palaeontological and geochemical data, including multiple elemental and isotopic proxies, available from the upper Pliensbachian – middle Toarcian of the Yorkshire coast, supplemented by results from the Dove’s Nest core, provide a unique framework for interpreting regional palaeoenvironmental change and the global impact of the T-OAE.

21.a. Late Pliensbachian – a well-oxygenated shallow-marine basin

The Staithes Sandstone and Cleveland Ironstone formations (A. margaritatusP. spinatum zones) display the greatest lithological and geochemical variation in the study interval with short-term mineralogical and grain-size changes reflecting a shallow-water and high-energy epeiric setting. Deposition of the Staithes Sandstone (Subunit Ia) occurred above the storm wave-base, deepening to largely sub-storm wave-base conditions for the Cleveland Ironstone (Subunits Ib – c). The high silt – fine sand fraction of the Staithes Sandstone is reflected in high and variable Si/Al, Na/Al, Ti/Al and Zr/Al ratios (quartz, feldspar and heavy mineral proxies) with an overall decreasing upward (fining) trend. Low CIA values point to a dominance of physical weathering processes supplying sediment to the basin and cool climate conditions.

The Cleveland Ironstone Formation presents more stable geochemical profiles interspersed with large Fe, Mn and Mg peaks associated with sideritic ironstone horizons. Five stacked CU cycles are evidenced within the Cleveland Ironstone Penny Nab and Kettleness members (cf. Rawson et al., Reference Rawson, Greensmith and Shalaby1983; Powell, Reference Powell2010) from detrital proxy element ratios and CIA values. Cycle 2 terminates with the Avicula Seam, a prominent siderite ironstone. Subunit Ib in coarsening-upward cycle 3, spanning the Amaltheus gibbosus Subzone, incorporates high δ13Corg values corresponding to the late Pliensbachian CIE and terminates at a regional disconformity below the Pecten Seam, a series of beds with 3 ironstones in the core. This cryptic disconformity, at the base of the Kettleness Member (P. spinatum Zone), is represented by an unremarkable bedding plane in the core but is a prominent feature of its chemostratigraphic profiles. The coarsening-upward cycles are interpreted as shallowing-upwards parasequences (’sequences’ of Macquaker & Taylor, Reference Macquaker and Taylor1996); medium-term cycle stacking patterns correspond well to the sequences and inferred relative sea-level curve for the Cleveland Basin proposed by Hesselbo (Reference Hesselbo2008). The base P. spinatum Zone disconformity is interpreted to represent a sequence boundary and superimposed transgressive surface.

Persistent oxic bottom-water conditions during the late Pliensbachian throughout the deposition of Unit I are evidenced by low TOC, TOC/PT and DOPT values and low redox-sensitive trace metal (e.g. U, Mo) contents. This is supported by a high trace-fossil taxonomic richness and a high benthic macrofossil diversity that includes bivalves, gastropods, brachiopods and benthic crinoids indicating a well-oxygenated open-marine environment. A proximal basin setting is indicated by palynofacies comprising 70 – 90 % terrestrial palynomorphs. Assemblages are dominated by fern spores and bisaccate pollen indicative of a cool moist climate, with subordinate dinocysts (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019), and low-abundance, low-diversity nannofossil assemblages representing marine plankton (Slater et al., Reference Slater, Bown, Twitchett, Danise and Vajda2022).

General cooling through the latest Pliensbachian is evidenced by belemnite calcite geochemistry: rising δ18Obel and falling Mg/Cabel ratios. The presence of glendonites in the P. spinatum Zone of southern Germany (Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019; Merkel & Munnecke, Reference Merkel and Munnecke2023) points to the episodic presence of cold bottom water masses in Europe. These have been associated with the establishment of a late Pliensbachian icehouse climate (Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019; Ruebsam & Schwark, Reference Ruebsam, Schwark, Reolid, Duarte, Mattioli and Ruebsam2021; Nordt et al., Reference Nordt, Breecker and White2022), including a persistent northern-hemisphere cryosphere with the likely presence of a NE Siberia icecap and extensive permafrost. Evidence provided in support of this includes regionally distributed glaciogenic sediments (pebbly argillites, tillites and matrix-supported conglomerates), potential periglacial sediments, dropstones and glendonites in Arctic Russia (Ruebsam & Schwark, Reference Ruebsam, Schwark, Reolid, Duarte, Mattioli and Ruebsam2021, fig. 8).

21.b. Pliensbachian – Toarcian boundary event – precursor to the T-OAE

The interval of the Pliensbachian – Toarcian Boundary CIE (PlToBE) marks a major shift in the chemostratigraphic profiles, reflecting the transition from the Cleveland Ironstone Formation to finer-grained mudstones of the Grey Shale Member (D. tenuicostatum Zone) at the base of the Whitby Mudstone Formation. This represents a significant long-term change in depositional conditions in the Cleveland Basin accompanying sea-level rise and coincides with the onset of a cascade of global environmental changes during the early Toarcian.

Hallam (Reference Hallam1997) estimated a relative sea-level rise of ∼35 – 85 m during the early Toarcian based on the facies change from the Staithes Sandstone to the Mulgrave Shale. A second-order eustatic sea-level rise of up to 100 m beginning in the earliest Toarcian D. tenuicostatum Zone and reaching a maximum in the middle Toarcian upper H. bifrons Zone has been proposed (Hardenbol et al., Reference Hardenbol, Thierry, Farley, Jacquin, Graciansky, Vail, Graciansky, Hardenbol, Jacquin and Vail1998; Haq, Reference Haq2018) that flooded large parts of the European shelf and other epicontinental areas globally. A transition from shallow-water limestones and/or sandstones to deeper-water mudstones is observed in many regions including the Paris (Hermoso et al., Reference Hermoso, Minoletti and Pellenard2013) and SE France basins (Bodin et al., Reference Bodin, Fantasia, Krencker, Nebsbjerg, Christiansen and Andrieu2023), Middle and High Atlas Morocco (Ait-Itto et al., Reference Ait-Itto, Price, Ait Addi, Chafiki and Mannani2017), NE Siberia Russia (Zakharov et al., Reference Zakharov, Shurygin, Il’ina and Nikitenko2006) and Neuquén Basin Argentina (Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022). Combined with the rising eustatic sea level, the collapse of the neritic carbonate factory in the earliest Toarcian halted carbonate mud export and caused sediment starvation in many basins previously dominated by carbonate-rich facies, as seen in the Central High Atlas Basin and elsewhere (Krencker et al., Reference Krencker, Fantasia, Danisch, Martindale, Kabiri, El Ouali and Bodin2020, Reference Krencker, Fantasia, El Ouali, Kabiri and Bodin2022; Bodin et al., Reference Bodin, Fantasia, Krencker, Nebsbjerg, Christiansen and Andrieu2023).

Step falls in detrital silicate and heavy mineral proxy element ratios at the facies change in the Cleveland Basin are matched by step increases in phyllosilicate proxies (K2O, K/Al, Cs/Al and Rb/Al) and CIA values indicative of decreased sand and silt fractions and the change to fine mudstones with an illite-rich clay-mineral assemblage. A coincident step rise in TOC may in-part reflect enhanced organic matter preservation via adsorption to the increased number of clay mineral surfaces (cf. Kennedy et al., Reference Kennedy, Pevear and Hill2002).

21.b.1. Sulphur Band events and onset of anoxia

The PlToBE comprises an interval with a pair of negative δ13Corg excursions of ∼2.5‰, the minima of which coincide with thin laminated mudstones containing up to 6% TOC (Littler et al., Reference Littler, Hesselbo and Jenkyns2010). These represent the first episodes of seawater anoxia and carbonaceous mudstone deposition in the Cleveland Basin, precursors to the thick black shales of the Jet Rock and Bituminous Shales, above. The Sulphur Band (SB1; Kettleness ‘bed’ 26, Hawsker Bottoms mid-’bed’ 43, Howarth, Reference Howarth1955) located a few decimetres above the base of the mudstones marks the base Toarcian. The laminated beds display geochemical evidence (DOPT, iron speciation, MoEF and isorenieratane proxies) for the intermittent development of an anoxic – euxinic water column.

High-resolution (cm-scale) geochemical records from the 15-cm thick Sulphur Band (SB1) (Salem, Reference Salem2013) demonstrate 2 – 3 short-term anoxia/euxinia to oxic cycles within the laminated interval over a period of < 10 ka (cf. McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008). Trace fossils immediately above the Sulphur Band likely represent a single, very brief oxygenation event. A return to more well-oxygenated bottom waters during the PlToBE is demonstrated by the presence of a diverse benthic fauna in the 1 m of sediment overlying SB1. However, geochemical proxies indicate predominantly dysoxic – anoxic conditions prior to the deposition of a second 50-cm interval of laminated mudstone, Sulphur Band 2 (SB2), at the top of the PlToBE interval. Palynofacies of the two ‘sulphur’ bands include high proportions (≥60%) of amorphous organic matter with small increases in marine algal cysts but spores and pollen continue to dominate (≥80%) the palynomorph assemblages (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019).

21.b.2. Relationship to Karoo–Ferrar LIPs magmatism

Sediments within and immediately below SB1 display significant enrichment in mercury; values of up to 0.5 ppm Hg at Hawsker Bottoms are one order of magnitude higher than in the immediately underlying and overlying strata (Percival et al., Reference Percival, Witt, Mather, Hermoso, Jenkyns, Hesselbo, Al-Suwaidi, Storm, Xu and Ruhl2015). Mercury enrichment is not confined to the black shale horizon. The highest Hg/TOC (ppm/%) ratios of up to 0.4 occur at the stage boundary immediately below the main black shale of SB1. Other levels within the PlToBE, including SB2, show no Hg/TOC enrichment. Mercury and Hg/TOC peaks at the level of the stage boundary, which also occur at Mochras, Peniche and in the Neuquén Basin of Argentina but are not seen in all Toarcian boundary sections (Them et al., Reference Them, Jagoe, Caruthers, Gill, Grasby, Gröcke, Yin and Owens2019), have been attributed to released volatiles from the onset of Karoo–Ferrar LIPs volcanism.

Robust 40Ar/39Ar and U–Pb geochronology indicates that peak Karoo-LIP basaltic magmatism postdated the PlToBE at 183.73 Ma (Greber et al., Reference Greber, Davies, Gaynor, Jourdan, Bertrand and Schaltegger2020; Jiang et al., Reference Jiang, Jourdan, Olierook and Merle2023). Nonetheless, the PlToBE has been equated to the onset of Karoo magmatism (Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022), based on records of older volcanic ages spanning the stage boundary (e.g. Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019, fig. S4; Luttinen et al., Reference Luttinen, Kurhila, Puttonen, Whitehouse and Andersen2022) and the presence of Hg/TOC anomalies in many sections. Furthermore, new high-precision plagioclase 40Ar/39Ar dates from Ferrar intrusions suggest earlier peak volcanism coincident with the PlToBE (Ware et al., Reference Ware, Jourdan and Timms2023), followed by waning and cessation ∼182 Ma at the termination of the T-OAE. An increase in 187Os/188Osi ratio from ∼0.4 to ∼0.5 within and above the PlToBE in Yorkshire and at Mochras evidence enhanced weathering accompanied the event (Percival et al., Reference Percival, Cohen, Davies, Dickson, Hesselbo, Jenkyns, Leng, Mather, Storm and Xu2016) which was potentially triggered by increasing atmospheric CO2 levels sourced from volcanic emissions.

Ruebsam et al. (Reference Ruebsam, Mayer and Schwark2019) argued that rising CO2 initiated a period of global warming that accelerated in D. semicelatum Subzone times and drove a transition from late Pliensbachian icehouse to early Toarcian greenhouse climate that caused ice sheet melting, thermal expansion of the oceans and glacioeustatic sea-level rise.

21.b.3. A Toarcian global reference section

Bodin et al. (Reference Bodin, Fantasia, Krencker, Nebsbjerg, Christiansen and Andrieu2023) proposed that the top Pliensbachian uppermost P. spinatum Zone is missing at Hawsker Bottoms and more generally in NW European sections and that a significant hiatus occurs at the base of the Sulphur Band (Bodin et al., Reference Bodin, Fantasia, Krencker, Nebsbjerg, Christiansen and Andrieu2023, fig. 10). However, their interpretation does not incorporate the onset of the PlToBE negative CIE preceding the stage boundary (Fig. 2). The double negative δ13C excursion spanning the stage boundary can be correlated between Yorkshire, Peniche, Mochras and the Lorraine Sub-Basin (Fig. 3; Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019, fig. 5), and there is a close association between a well-defined negative δ13Corg excursion and the faunal turnover of ammonite taxa associated with the stage boundary in the Neuquén Basin, Argentina (Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022).

There is no clear biostratigraphic or sedimentological evidence of a hiatus at the stage boundary at Hawsker Bottoms (cf. Kemp & Sadler, Reference Kemp, Sadler and Coe2022) although some condensation at the level of the uppermost P. spinatum Zone ironstone (‘bed’ 42, top of the Cleveland Ironstone Formation), below, is likely. Nonetheless, we consider that the Yorkshire expression of the PlToBE remains a viable reference for global correlation.

21.b.4. Biotic change and Toarcian extinctions

The PlToBE marks a period of major biotic change with the disappearance of many bivalve taxa and a major decline in trace-fossil taxonomic richness indicating increasingly dysoxic bottom conditions. Marine palynomorphs display a decline in dinocysts and increase in spiny acritarchs and algal cysts. Caswell et al. (Reference Caswell, Coe and Cohen2009) considered that macrofossil extinctions in the P. paltum Subzone at the top of the PlToBE level (their extinction horizon i), described in detail from Yorkshire, were of global significance. Major changes in pollen assemblages across the stage boundary in the Cleveland Basin, including a decline in bisaccates and substantial increases in Classopollis spp. and Chasmatosporites spp., are indicative of a transition to a strongly seasonal warmer climate (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019).

21.c. Earliest Toarcian – sea-level rise, oxygen depletion and climate warming

Following the PlToBE, the Grey Shale of the P. paltum to lower D. semicelatum subzones (Unit II) shows more uniform mineralogy and geochemistry compared to the underlying Pliensbachian reflecting a deeper water environment, likely close to the limit of storm wave-base (typically 15 – 40 m depth), established following sea-level rise. Detrital proxies (Ti/Al ratio, quartz and silt content) display variation but no long-term trend following their marked fall within the interval of the PlToBE. A third thin interval of laminated black shales in the lower D. clevelandicum Subzone, Sulphur Band 3 (SB3), evidences a further short-lived episode of temporary bottom-water anoxia. This occurs above a minor CU cycle with rising K/Al ratios pointing to an increase in the proportion of illitic clays. SB3 marks a further significant environment shift, reflected by the Subunit IIa – b boundary, with a step change to higher TOC and amorphous organic matter contents, geochemical evidence of increasingly dysoxic bottom waters (e.g. rising DOPT values), and the temporary disappearance of trace fossils.

A climate shift to a warming trend through the earliest Toarcian is indicated by falling δ18Obel and rising Mg/Cabel ratios from the Yorkshire coast (Section 18), consistent with brachiopod oxygen-isotope data from Portugal (Suan et al., Reference Suan, Mattioli, Pittet, Mailliot and Lecuyer2008a; Müller et al., Reference Müller, Jurikova, Gutjahr, Tomasovych, Schlogl, Liebetrau, Duarte, Milovsky, Suan, Mattioli, Pittet and Eisenhauer2020) and TEX86 results from Spain and Italy (Ruebsam et al., Reference Ruebsam, Reolid, Sabatino, Masetti and Schwark2020c). Palynological trends include a rising upward proportion of Chasmatosporites spp. cycad pollen, a warm/dry climate indicator, and amorphous organic matter, but the proportion of terrestrial palynomorphs remains generally ≥80% (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019). The impact of early Toarcian climate warming on terrestrial vegetation was not limited to the Cleveland Basin, with palynological evidence of significant warming recorded from many regions, including continental settings in Central Asia (Schnyder et al., Reference Schnyder, Pons, Yans, Tramoy, Abdulanova, Brunet, McCann and Sobel2017).

21.d. T-OAE (Unit III) – expansion of anoxia and a hyperthermal event

The interval of the negative δ13C excursion defining the T-OAE (Unit III), spanning the upper D. semicelatum and C. exaratum subzones, represents the period of most extensive palaeoenvironmental change in the Cleveland Basin and globally. The base of T-OAE Unit III is defined by the onset of falling δ13C values recorded in bulk sediment organic matter, terrestrial wood, individual marine and terrestrial biomarkers, and rising TOC content, followed, a short distance above, by the onset of laminated black shale deposition and the disappearance of trace fossils and most benthic fauna. This reflects the establishment of more persistent anoxia and then, from the base of the Jet Rock (C. exaratum Subzone), euxinic bottom waters that dominated through the interval of the δ13Corg minimum that characterizes Subunit IIIb. Anoxia persisted after the T-OAE, throughout Unit IV, into the later early Toarcian, upper H. falciferum Subzone.

21.d.1. Global mass extinction, ocean anoxia and a Toarcian hyperthermal

Biotic turnover with numerous highest occurrences in the upper D. semicelatum Subzone (Subunit IIIa) is well documented in the Cleveland Basin and has been widely observed in other Boreal and Tethyan sections, with the recognition of two extinction levels (ii, iii) by Caswell et al. (Reference Caswell, Coe and Cohen2009). The highest of these, at the base of the C. exaratum Subzone and the fall to a δ13Corg minimum (Subunit IIIb), marks the level of a second-order global mass extinction which, based on geochemical proxies (e.g. Re/Mo, δ98Mo, δ18Obel), coincided with the maximum spread of anoxic and euxinic seafloor area in the oceans and peak hyperthermal conditions. Terrestrial palynomorphs evidence an increasingly hot climate with extreme wet/dry seasons in the Cleveland Basin at that time.

21.d.2. A euxinic basin

Iron speciation (FeHR/FeT, Fepy/FeHR), TOC/PT and DOPT together with trace-metal enrichment (Mo, U) indicate that euxinic bottom waters became established progressively in the Cleveland Basin during the onset of the T-OAE (Subunit IIIa) and remained present throughout the subsequent phases of the event (Subunits IIIb – d). Isorenieratane/TOC ratios and other biomarkers suggest that PZE, which occurred initially for short periods during the deposition of the black shales of the Sulphur Bands, was commonplace during the T-OAE. However, horizons of low-oxygen specialist bivalves, principally Pseudomytiloides dubius, throughout the black shale succession point to common brief episodes of improved oxygenation that fail to be resolved by the sediment geochemical proxies.

Oxidation events may have been caused by large storms mixing the stratified water column and mobilizing surface sediments but having little impact on the deeper sediment geochemical record. A dramatic increase of tropical cyclone intensity during the T-OAE global warming that affected large areas of NW Tethys, including the European epicontinental seas, has been proposed based on sedimentological evidence obtained in a number of west European and Moroccan sections (Krencker et al., Reference Krencker, Bodin, Suan, Heimhofer, Kabiri and Immenhauser2015) and modelling that indicates two potential storm genesis centres located around NW and SE Tethys (Marsaglia & Klein, Reference Marsaglia and Klein1983; Yan et al., Reference Yan, Li, Kemp, Guo, Zhang and Hu2023).

21.d.3. The Whale Stones Event

Black shale deposition accompanied the T-OAE in the Cleveland Basin. The stratigraphic pattern of TOC enrichment shows a large right-skewed peak centred in the mid-C. exaratum Subzone at Whale Stones ‘bed’ 35. Here, maximum TOC contents of >12% (maximum 19%) immediately precede the reversal to rising δ13C values in the mid-C. exaratum Subzone (‘bed’ 36) and coincide with a maximum in the relative abundance of amorphous organic matter (∼90%) and algal cysts and peak nannofossil species richness.

Whale Stones ‘bed’ 35 shows peak values for detrital grain-size proxies and a CIA minimum with maximum 187Os/188Os values and an increase in the slope of the Sr-isotope profile. An increase in physical weathering and/or relative sea-level fall promoted an increased input of illitic clay and coarser detritus into the basin. However, the large carbonate concretions that characterize this ‘bed’ imply a reduction in bulk sedimentation rate and/or hiatus (Raiswell, Reference Raiswell and Marshall1987, Reference Raiswell1988; Marshall & Pirrie, Reference Marshall and Pirrie2013). Sharp falls in TOC and detrital proxy elements occur at the base of ‘bed’ 36 (Figs 4, 7) together with a marked increase in the slope of the 87Sr/86Srbel profile (Fig. 19). These point to a possible hiatus at this level, followed by a period of slower deposition. Concretion formation likely occurred at this time.

Peak DOPT and FeHR/FeT values, high MoEF, VEF, UEF and a Re/Mo minimum indicate peak euxinia accompanying maximum TOC in the Basin. Despite a near absence of benthic fauna and a lack of visible bioturbation, Whale Stones ‘bed’ 35 yields diverse ammonite and nannofossil assemblages and belemnites demonstrating a continuing connection, at least intermittently, with oxygenated open-marine surface waters and a stratified water column. A maximum Cd/Mo ratio (McArthur, Reference McArthur2019) at this level supports minimum basin restriction but will have been influenced by the global drawdown of redox-sensitive trace metals at this time. A Mo/TOC minimum and low δ98Mo point to a coincident maximum in the global seafloor area indicating that basin water chemistry generally follows a global trend, albeit modified by episodes of basin restriction.

Primary productivity did not collapse in response to the PZE in the Cleveland Basin. Rather, an ecosystem dominated by anoxygenic photosynthesizers provided an increased organic matter flux to the sediment, which was better preserved under the prevailing euxinic bottom-water conditions, generating maximum TOC enrichment. High TOC/PT molar ratios of >200 point to extensive P release from euxinic sediments maintain water column productivity, while δ15Ntot values of ∼+2 indicate that increased algal productivity drove partial denitrification.

A major change in both palaeoenvironmental conditions in the basin and globally occurred following the deposition of Whale Stones ‘bed’ 35, this Whale Stones Event marked the onset of global cooling, associated with decreasing volcanic emissions, falling pCO2, decreased continental weathering and better oxygenation of the global ocean with rising δ13C marking the final stages of the T-OAE. A sharp increase in Cerebropollenites pollen above Whale Stones ‘bed’ 35 that persists to the top of the T-OAE interval is a prominent feature on the Yorkshire coast (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019) and is likely a response of the terrestrial vegetation to the hyperthermal maximum. The coincident changes in multiple marine and terrestrial proxies at this level represent a major event that offers potential as a global marker.

21.d.4. TOC and carbon burial rates

TOC enrichment accompanying the T-OAE is widespread, with a majority of sites globally (albeit heavily biased towards European sections) showing an increase in TOC relative to pre-event values, but each basin displays a unique stratigraphic pattern (Remirez & Algeo, Reference Remirez and Algeo2020; Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b). The highest mean values are recorded in NW and SW German euxinic basins. TOC increases are generally highest where deoxygenation is most severe but multiple local factors including productivity and sedimentation rate fundamentally affect the final stratigraphic record and the distribution of black shales through a succession. Kemp et al. (Reference Kemp, Suan, Fantasia, Jin and Chen2022b) observed that in anoxic – euxinic environments of the T-OAE, TOC content displays a non-linear inverse correlation with sedimentation rate indicating that clastic dilution plays a major role in limiting organic richness.

Surprisingly, the T-OAE was characterized by relatively low organic carbon burial rates, even in places where TOC concentrations are high (e.g. Suan et al., Reference Suan, Schlogl and Mattioli2016; Fantasia et al., Reference Fantasia, Föllmi, Adatte, Spangenberg and Montero-Serrano2018b). An organic carbon burial rate of 1.63 g C m−2 a−1 has been calculated for the T-OAE interval in Yorkshire, assuming a T-OAE duration of 900 ka (Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b). Globally, C burial rates during the CIE interval are <3 g C m−2 a−1 with a mean rate of 0.57 g C m−2 a−1 based on the same time scale. These contrast to a modern global median continental-shelf organic carbon burial rate of 19.6 g C m−2 a−1 (Wilkinson et al., Reference Wilkinson, Besterman, Buelo, Gephart and Pace2018) and rates are up to ∼100 g C m−2 a−1 in high-productivity upwelling zones such as the Peru margin (Föllmi et al., Reference Föllmi, Badertscher, de Kaenel, Stille, John, Adatte and Steinmann2005).

Post-depositional oxidation and diagenetic loss of organic carbon complicate comparison with modern environments but, nonetheless, the relatively low rates of burial in anoxic – euxinic settings like the Cleveland Basin, areas of potentially high burial efficiency, indicate that rates of surface ocean productivity and export production were low during the T-OAE relative to modern shelves. Nonetheless, it has been estimated that organic carbon burial rates may have increased ∼500% on average during the T-OAE compared to the earlier Toarcian in the anoxic–euxinic marine basins of Europe (Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b).

21.d.5. Basin restriction versus global anoxia

Geochemical proxies in the black shales of the Jet Rock and Bituminous Shales predominantly record global-scale changes in ocean chemistry that occurred during and immediately following the T-OAE. Euxinia in the Cleveland Basin was a local manifestation of the global expansion of anoxic and euxinic seafloor area, which led to a fall in seawater sulfate, evidenced by rising δ34SCAS in sediments and depletion of redox-sensitive trace metals (Mo, Re, U, V) in the oceans, reflected in anomalous low trace-metal/TOC and Re/Mo ratios and a depleted isotopic composition (δ98Mo).

Estimates of local basin restriction using geochemical data derived from modern basins are misleading if they fail to incorporate the large changes in global ocean chemistry that occurred during the early Toarcian. However, extreme depletion of Mo (and to a lesser extent U and V) in the Cleveland Basin within the T-OAE Unit III interval during peak euxinia is consistent with a restriction at that time, prior to the development of a more open connection accompanying sea-level rise during the later early Toarcian H. falciferum Subzone (Unit IV). Cyclic decreases in δ98Mo to ∼1‰ coincident with rising Mo concentrations in the T-OAE interval are indicative of basin restriction cycles.

21.d.6. Salinity change

A widespread reduction in salinity across the NW European Shelf during the T-OAE has been widely postulated, including a shift to nearly freshwater conditions in the Cleveland Basin based on geochemical proxies (principally B/Ga ratio; Remirez & Algeo, Reference Remirez and Algeo2020; Remírez & Algeo, Reference Remírez and Algeo2020). Although minor regional freshening of surface waters likely occurred, as indicated by regional ocean circulation models (Bjerrum et al., Reference Bjerrum, Surlyk, Callomon and Slingerland2001; Dera & Donnadieu, Reference Dera and Donnadieu2012; Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018) and interpretation of δ18O data, with salinity falling to a minimum of 30 – 32 psu, there is no geologically consistent evidence of extreme brackish or freshwater conditions in the basin during the T-OAE. The presence of abundant ammonites, belemnites and pseudoplanktonic crinoids demonstrates open-marine (likely stenohaline) conditions, in association with peak levels of amorphous organic matter and marine algal cysts and greatest nannofossil species richness.

21.d.7. A hyperthermal event with enhanced weathering

The T-OAE is recognized as one of the six most significant hyperthermals in the last 300 Ma (Foster et al., Reference Foster, Hull, Lunt and Zachos2018). An increase in seawater temperatures of 6 – 7° C in the Cleveland Basin, estimated from δ18O values in belemnite calcite, led to hyperthermal conditions at the peak of the T-OAE (Subunit IIIb). Globally, the magnitude of seawater warming is estimated to have ranged between +3˚C and +7˚C, depending on latitude (Bailey et al., Reference Bailey, Rosenthal, McArthur, van de Schootbrugge and Thirlwall2003; Suan et al., Reference Suan, Mattioli, Pittet, Mailliot and Lecuyer2008a; Gómez & Goy, Reference Gómez and Goy2011; Dera & Donnadieu, Reference Dera and Donnadieu2012; Danise et al., Reference Danise, Clemence, Price, Murphy, Gomez and Twitchett2019).

A coincident large 187Os/188Osi peak of >1.0, a pulse of illite and rising 87Sr/86Sr in belemnites, point to an episode of accelerated weathering accompanying warming. An associated small increase in the proportion (maximum 3.5%) of freshwater algae, which are present sporadically (<1%) throughout the Yorkshire succession, is consistent with an input of additional terrestrial material that might incorporate freshwater-derived palynomorphs. Comparable temperature rises and weathering pulses have been widely documented coincident with the T-OAE in Wales, Germany, Canada and Japan (e.g. Percival et al., Reference Percival, Cohen, Davies, Dickson, Hesselbo, Jenkyns, Leng, Mather, Storm and Xu2016; Them et al., Reference Them, Gill, Selby, Gröcke, Friedman and Owens2017b; Kemp et al., Reference Kemp, Selby and Izumi2020).

An increase in the proportion of Classopollis spp. pollen with decreasing taxonomic terrestrial palynomorph richness and diversity occur at the onset of the T-OAE in the Cleveland Basin (Subunit IIIa) (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019). This is followed by a large pulse of Cerebropollenites accompanying the Whale Stones Event in the upper C. exaratum Zone. These changes in terrestrial vegetation point to the temporary establishment of a hot climate with extreme wet/dry seasons, followed by a return to a warm climate and strong seasonality immediately following the T-OAE. The disappearance of glendonites in polar sections (North Siberia) coincident with an influx of abundant Classopollis spp. pollen at the onset of the T-OAE has similarly been attributed to high-latitude warming that persisted into the middle Toarcian (Suan et al., Reference Suan, Nikitenko, Rogov, Baudin, Spangenberg, Knyazev, Glinskikh, Goryacheva, Adatte, Riding, Follmi, Pittet, Mattioli and Lecuyer2011).

A phase of overall aridity linked to dominance of dry season conditions during the later stages of the T-OAE (Units IIIc, d) is evidenced by coincident peaks in the relative proportion of Cerebropollenites in Yorkshire and maximum charcoal abundance (incidence of wildfires) in equivalent levels at Mochras and Peniche (Baker et al., Reference Baker, Hesselbo, Lenton, Duarte and Belcher2017). There is increasing evidence of arid phases during the T-OAE throughout the warm temperate belt of the Northern Hemisphere (Baranyi et al., Reference Baranyi, Jin, Dal Corso, Li and Kemp2024). Increased wildfire activity during the later stages of the T-OAE has been linked to increasing pO2 driven by organic carbon and pyrite burial (Baker et al., Reference Baker, Hesselbo, Lenton, Duarte and Belcher2017).

21.d.8. The Karoo–Ferrar LIPs, pCO2 and the T-OAE

The peak of the main magmatic phase of the Karoo LIP between 183.5 and 182.5 Ma (Jiang et al., Reference Jiang, Jourdan, Olierook and Merle2023) and the reinterpreted Ferrar LIP peak between 184 and 182.5 Ma (Ware et al., Reference Ware, Jourdan and Timms2023) coincide with the acceleration of early Toarcian global warming and immediately pre-date the beginning of the T-OAE negative δ13C excursion at 182.77+0.11/−0.15 Ma (Al-Suwaidi et al., Reference Al-Suwaidi, Ruhl, Jenkyns, Damborenea, Manceñido, Condon, Angelozzi, Kamo, Storm, Riccardi and Hesselbo2022). In contrast to the Karoo–Ferrar LIPs, Chon Aike, which has also been implicated in the T-OAE (e.g. Krencker et al., Reference Krencker, Fantasia, Danisch, Martindale, Kabiri, El Ouali and Bodin2020), is a silicic magmatic province that may not be plume-related (Bastias et al., Reference Bastias, Spikings, Riley, Ulianov, Grunow, Chiaradia and Hervé2021). It was emplaced over a long period ∼ 153 – 188 Ma with one of three pulses between 178 – 188 Ma (Pankhurst et al., Reference Pankhurst, Riley, Fanning and Kelley2000) and likely did not result in rapid hydrothermal venting of greenhouse gases.

Variations in the concentration of mercury (Hg) in sedimentary archives, generally expressed as Hg/TOC ratios, are increasingly being used as a proxy for past global volcanic activity (e.g. Percival et al., Reference Percival, Witt, Mather, Hermoso, Jenkyns, Hesselbo, Al-Suwaidi, Storm, Xu and Ruhl2015, Reference Percival, Bergquist, Mather, Sanei, Ernst, Dickson and Bekker2021) but the interpretation of sedimentary Hg anomalies is not straightforward (Grasby et al., Reference Grasby, Them, Chen, Yin and Ardakani2019). Unlike the PlToBE and despite significant mercury increases, no Hg/TOC anomaly has been documented in the T-OAE interval of the Cleveland Basin, in contrast to Mochras and many sections in Portugal, Spain, Canada, Chile and Argentina (e.g. Them et al., Reference Them, Jagoe, Caruthers, Gill, Grasby, Gröcke, Yin and Owens2019; Ruhl et al., Reference Ruhl, Hesselbo, Jenkyns, Xu, Silva, Matthews, Mather, Mac Niocaill and Riding2022).

Kovács et al. (Reference Kovács, Ruhl, Silva, McElwain, Reolid, Korte, Ruebsam and Hesselbo2024) reassessed Hg distributions in the Pliensbachian – Toarcian of the Cleveland Basin and at Mochras and La Cerradura (Subbetic Basin, southern Spain). The study showed that the use of Hg/TOC ratio alone as a volcanic emission proxy may be misleading since it does not always fully correct for host-phase biases, nor does it quantify the additional environmental Hg present over background levels during periods of sedimentary Hg enhancement (Fendley et al., Reference Fendley, Frieling, Mather, Ruhl, Hesselbo and Jenkyns2024). For example, sulfides rather than organic matter provide the main Hg host in some Lower Jurassic successions (Zhu et al. Reference Zhu, La Croix, Kemp, Shen, Huang, Hua, Li and Wei2024).

A strong affinity of Hg with organic carbon – sulfide phases was observed by Kovács et al. (Reference Kovács, Ruhl, Silva, McElwain, Reolid, Korte, Ruebsam and Hesselbo2024) solely in the anoxic – euxinic intervals of the Cleveland Basin. Under oxic – dysoxic conditions, both in Yorkshire and elsewhere, Hg principally co-varied with redox-sensitive trace metals, suggesting an association with Fe–Mn oxyhydroxides and/or detrital (clay) minerals. Nonetheless, substantial Hg enrichment characterized the T-OAE interval in all redox environments, consistent with an increase in Hg loading contemporaneous with LIP volcanism.

Hg/TOC enrichments associated with the T-OAE have been recorded mainly in shallow-marine settings and show evidence of cycling and redistribution of Hg in intermediate terrestrial reservoirs prior to being incorporated in the marine record. Mercury enrichment in fossil leaves from a T-OAE section in south China has been reported (Näslund, Reference Näslund2021), while Hg concentration and Hg-isotope data from a lacustrine T-OAE section in the Ordos Basin of north China point to Hg enrichment solely from a terrestrial source and exclude direct increased atmospheric Hg deposition (Jin et al., Reference Jin, Zhang, Baranyi, Kemp, Feng, Grasby, Sun, Shi, Chen and Dal Corso2022). Thus, although volcanogenic outgassing remains the potential primary source of Hg associated with the T-OAE, an increase in terrestrially derived Hg in the global ocean may have been enhanced by continental weathering (evidenced by Os and Sr isotope data) in response to rapid climate warming.

The primary environmental impact of LIP emplacement was the release of volcanic volatiles, principally CO2, into the atmosphere that drove global warming (cf. Jenkyns, Reference Jenkyns1999). Fendley et al. (Reference Fendley, Frieling, Mather, Ruhl, Hesselbo and Jenkyns2024) used excess Hg loading data from Mochras to calculate LIP-associated carbon emissions during the late Sinemurian – late Toarcian. They estimated that a total of ∼12,000 Gt C was released during the entire T-OAE interval, with an estimated duration of ∼1 Ma, in contrast to faster and larger carbon inputs of up to 81,000 Gt C over 150 ka that have been derived from carbon-cycle modelling (Ullman et al., Reference Ullmann, Boyle, Duarte, Hesselbo, Kasemann, Klein, Lenton, Piazza and Aberhan2020; Heimdal et al., Reference Heimdal, Goddéris, Jones and Svensen2021).

A pCO2 increase from 400 – 1200 ppmv in the early Toarcian to 1100 – 1800 ppmv during the T-OAE is indicated by leaf stomata data from Denmark (McElwain et al., Reference McElwain, Wade-Murphy and Hesselbo2005). A doubling of pCO2 from ∼500 ppmv to ∼1000 ppmv during the T-OAE is supported by the changing offset between terrestrial δ13Cn-alkane and marine δ13Ccarb data (Ruebsam et al., Reference Ruebsam, Reolid and Schwark2020d).

21.d.9. pCO2, ocean circulation and anoxia

Modelling by Dera & Donnadieu (Reference Dera and Donnadieu2012) suggests that a 2 – 6 × increase in atmospheric pCO2 during the early Toarcian would have led to an average global warming of +4.5° C associated with stronger high-latitude precipitation rates, enhanced continental runoff and the demise of polar sea ice, causing a regional freshening of Arctic surface seawater. The model shows a progressive slowdown of global oceanic circulation that would promote widespread ocean stratification and bottom anoxia in deep oceanic settings and epicontinental basins.

The mass balance model for the coupled marine P and C cycles of Slomp & Van Cappellen (Reference Slomp and Van Cappellen2007) predicts that deep-ocean anoxia would occur if the present ocean’s circulation rate was reduced by 50% while the supply of reactive phosphorus from the continents was simultaneously boosted by 20%. They suggested that such an increase could be caused by coastal erosion linked to sea-level rise. The transition between oxic to anoxic deep-ocean state is likely very rapid (Donohue et al., Reference Donohue, Florio and Fowler2023).

Slomp & Van Cappellen’s (Reference Slomp and Van Cappellen2007) model indicates that a slowdown of global ocean circulation decreases primary production in the open ocean, but increases that in marginal seas. Slower ocean circulation increases global organic carbon burial, because of enhanced preservation of organic matter under anoxia – euxinia in deep-sea environments and higher primary productivity along continental margins. Carbon burial is enhanced further when reduced oceanic circulation also causes the spreading of bottom-water anoxia in the coastal ocean.

Unlike Cretaceous oceanic anoxic events (Montoya-Pino et al., Reference Montoya-Pino, Weyer, Anbar, Pross, Oschmann, van de Schootbrugge and Arz2010; Owens et al., Reference Owens, Gill, Jenkyns, Bates, Severmann, Kuypers, Woodfine and Lyons2013; Ostrander et al., Reference Ostrander, Owens and Nielsen2017; Clarkson et al., Reference Clarkson, Stirling, Jenkyns, Dickson, Porcelli, Moy, Pogge von Strandman, Cooke and Lenton2018), little is known about the intensity and global extent of oceanic anoxia in the Toarcian. It seems likely, however, that the area of anoxic seafloor during the T-OAE would have been similar to that during the latest Cenomanian OAE2 (Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017), when that area expanded by ∼40% to occupy perhaps ∼20% of the global seafloor (Ostrander et al., Reference Ostrander, Owens and Nielsen2017). Constraints on the area of seafloor overlain by euxinic seawater in that case indicate that ∼5% of the seafloor was euxinic (Owens et al., Reference Owens, Gill, Jenkyns, Bates, Severmann, Kuypers, Woodfine and Lyons2013; Dickson et al., Reference Dickson, Gill, Ruhl, Jenkyns, Porcelli, Idiz, Lyons and van den Boorn2017). These numbers may seem low (cf. Monteiro et al., Reference Monteiro, Pancost, Ridgwell and Donnadieu2012), but even in warm oceans, it is difficult to establish and maintain an anoxic water column away from productive margins (Lyons & Reinhard, Reference Lyons and Reinhard2012).

It is important to note that the results of Dera & Donnadieu (Reference Dera and Donnadieu2012) contrast with the results of other modelling studies of ocean circulation during OAEs which show no collapse of ocean circulation but rather similar to slightly higher intensities of overturning (Poulsen et al., Reference Poulsen, Barron, Arthur and Peterson2001; Otto-Bliesner et al., Reference Otto-Bliesner, Brady and Shields2002; Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Tuenter, Henstra, van der Zwan, van de Wal, Dijkstra and de Boer2010). The meridional gradient of density at the ocean surface increases with increasing CO2 concentration for warmer climates because the coefficient of expansion of seawater is greatly increased at higher temperatures (Manabe & Bryan Jr, Reference Manabe and Bryan1985). For this reason, ocean circulation in warmer climates does not slow down.

A recent modelling study suggests that the Tethys Ocean and the SW European epicontinental sea were mostly well-oxygenated (Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018). More northern parts of Europe were too shallow for currents of oxygenated Tethyan water to penetrate deep into the epeiric sea with its abundant sills and islands (Fig. 1), which was highly prone to stratification (Ruvalcaba Baroni et al., Reference Ruvalcaba Baroni, Pohl, van Helmond, Papadomanolaki, Coe, Cohen, van de Schootbrugge, Donnadieu and Slomp2018). Salinity stratification, either due to runoff from the surrounding land, to rainfall, or both, would have promoted anoxia and burial of organic carbon in the basins, although suggestions of freshwater episodes in the Cleveland Basin (McArthur et al., Reference McArthur, Algeo, van de Schootbrugge, Li and Howarth2008; Remírez & Algeo, Reference Remírez and Algeo2020) are not supported by the geological evidence.

Relatively low organic carbon burial rates during the T-OAE in the Cleveland Basin and globally (Kemp et al., Reference Kemp, Suan, Fantasia, Jin and Chen2022b) point to high burial efficiency in anoxic – euxinic environments, rather than increased rates of surface ocean productivity and export production being the primary cause of TOC enrichment in epeiric seas. Nonetheless, high ratios of TOC/P indicate strong recycling of phosphorus relative to organic carbon which likely helped sustain productivity and anoxia. However, an expansion of anoxic – euxinic seafloor area and increased organic carbon preservation in the Tethyan and Panthalassic oceans is required to explain the elemental and isotopic changes recorded in Toarcian black shales.

21.d.10. Magnitude of the T-OAE global negative δ13C excursion

The magnitude of the negative δ13Corg excursion of ∼6‰ that characterizes the T-OAE in the Cleveland Basin and many other northern European sections is such that an input of mantle-derived carbon alone would require unrealistic volumes of CO2 emission. However, large variation exists in the amplitudes of negative δ13Ccarb and δ13Corg excursions observed globally (Fig. 3; Remirez & Algeo, Reference Remirez and Algeo2020, figs 4, 5), reflecting a combination of global and local controls.

Organic matter type and bottom-water redox, in particular, significantly impact bulk δ13Corg values (Remirez & Algeo, Reference Remirez and Algeo2020): terrestrial-derived organic matter preserved in relatively well-oxygenated conditions exhibits a mean value that is 3.8‰ higher than marine-derived organic matter preserved during the T-OAE in relatively more reducing bottom waters (–28.8‰ versus –32.6‰), with mixed-source organic matter exhibiting an intermediate value (–30.1‰). These compare to the average Phanerozoic black shale value of −27‰ (Meyers, Reference Meyers2014). Changes in the relative abundance of the isotopic endmembers will significantly influence the carbon isotope curve (Suan et al., Reference Suan, van de Schootbrugge, Adatte, Fiebig and Oschmann2015). This is particularly problematic for the Yorkshire succession where the terrestrial component values from >90% in some upper Pliensbachian samples to <5% during the latter part of the T-OAE (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019).

The HI obtained by Rock-Eval pyrolysis can be used to track organic matter sources because fresh and weakly degraded algal and marine-derived organic matter (kerogen Types I and II) are hydrogen-rich, whereas terrestrial-derived and highly-degraded marine-derived organic matter (kerogen Type III) has a lower hydrogen content (Espitalié et al., Reference Espitalié, Deroo and Marquis1985; Tyson, Reference Tyson and Tyson1995). Suan et al. (Reference Suan, van de Schootbrugge, Adatte, Fiebig and Oschmann2015, fig. 5d) used the Yorkshire coast paired δ13Corg and HI data of Sælen et al. (Reference Sælen, Tyson, Telnaes and Talbot2000) to compensate for the changing proportion of terrestrial vs marine organic matter through the succession. They calculated an amplitude for the corrected negative excursion of 3‰ – 4‰ δ13Corg, consistent with corrected δ13Corg, δ13Cphytane and δ13Ccarb curves from two sections in SW Germany (Denkingen, Dotternhausen) that are dominated by marine organic matter throughout the lower Toarcian.

Remirez & Algeo (Reference Remirez and Algeo2020) concluded that the T-OAE CIE is perhaps best interpreted as a combination of a relatively small negative shift of –2‰ to –3‰ in the isotopic composition of the global carbon cycle, overprinted by larger local or regional shifts of up to ±5‰ (e.g. Fig. 3) caused by, for example, differences in organic matter type, changes in marine productivity, water mass restriction, assimilatory uptake of variable amounts of recycled CO2 from bacterial respiration and burial history. Additionally, the shape of the isotope curve will be affected by local physical sedimentary processes, which means that both the shape and the magnitude of the curves are a combination of global and local factors.

The large amplitude of the T-OAE negative δ13C excursion has been interpreted to require a large release of isotopically light methane to the atmosphere caused by the dissociation of methane clathrates on continental margins (Hesselbo et al., Reference Hesselbo, Gröcke, Jenkyns, Bjerrum, Farrimond, Bell and Green2000; Beerling et al., Reference Beerling, Lomas and Gröcke2002; DB Kemp et al., Reference Kemp, Coe, Cohen and Schwark2005) or associated with permafrost melting (Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019) or thermogenically released by igneous intrusions into coals and carbonaceous mudstones (McElwain et al., Reference McElwain, Wade-Murphy and Hesselbo2005; Svensen et al., Reference Svensen, Planke, Chevallier, Malthe-Sørenssen, Corfu and Jamtveit2007; Deegan et al., Reference Deegan, Bédard, Grasby, Dewing, Geiger, Misiti, Capriola, Callegaro, Svensen, Yakymchuk, Aradi, Feda and Troll2022). Based on a negative CIE of about −3%‰, the total amount of carbon released during the T-OAE was estimated to be ∼10,000 Gt by Ruebsam et al. (Reference Ruebsam, Mayer and Schwark2019) assuming a biogenic methane source of carbon (δ13C = −60‰ to −80‰), but >50,000 Gt would be required for a purely volcanic CO2 source (δ13C = −7‰ to −10%). Furthermore, massive carbon burial during the T-OAE would potentially move the carbon-isotope value of the ocean–atmosphere system to more positive values, against which the input of isotopically negative carbon would be balanced. This has implications for calculating the amount of carbon involved in the generation of the T-OAE negative excursion and/or its isotopic composition. Additional modelling is required to address this further.

The release of 10,000 Gt carbon over a period of about 800 ka (see Section 17 for a discussion of the duration of the T-OAE) has been calculated to increase global temperatures by about 5° C and cause a two- to three-fold increase in pCO2 (Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019). For comparison, an increase in atmospheric pCO2 from ∼500 to ∼1000 ppmv during the T-OAE is derived from reconstructions using the stomatal index and the difference in the magnitude of the CIE in marine and terrestrial environments (McElwain et al., Reference McElwain, Wade-Murphy and Hesselbo2005; Ruebsam et al., Reference Ruebsam, Reolid and Schwark2020d). A seawater temperature rise of 6 – 7° C is indicated for the Cleveland Basin (Section 18).

Kemp et al. (Reference Kemp, Suan, Fantasia, Jin and Chen2022b) estimated that globally an extra ∼9000 Gt of carbon may have been buried in shallow seas during the T-OAE relative to before the event, including ∼800 Gt in European anoxic – euxinic basins.

21.e. Early Toarcian – post-T-OAE anoxia and global cooling

Anoxic – euxinic conditions persisted through the lower H. falciferum Subzone (lower Bituminous Shales) in the Cleveland Basin following the T-OAE with continuing deposition of carbonaceous (>2.5% TOC) mudstones (lower Bituminous Shales) that display the greatest Mo, U, V and other redox-sensitive trace metal enrichment in the succession (Unit IV). Consistently high DOPT and FeHR/FeT values point to continuing anoxic conditions with cyclic redox variation indicated by PEF and trace-metal profiles. δ98Mo data indicate a weakly restricted anoxic basin with sufficient exchange to display open ocean isotopic values. The high levels of redox-sensitive metals relative to TOC reflect their increase in ocean water in response to a decline in the areal extent of global anoxic and euxinic seafloor with the termination of the T-OAE. The step fall in trace metals in the upper H. falciferum Subzone (base Unit V) reflects a shift to dysoxic bottom waters in the basin with sediments no longer capturing a global trace-metal signature except during brief euxinic episodes like those in the Hard Shales.

The T-OAE hyperthermal marks a step change in the long-term δ18Obel and Mg/Cabel records (Fig. 19) indicating a shift from relatively cool-water conditions in the Cleveland Basin before the event and the persistence of warm waters following the termination of the hyperthermal. It has been suggested that this reflects a major change in the Earth’s climate system (Ruebsam et al., Reference Ruebsam, Mayer and Schwark2019, Reference Ruebsam, Reolid, Marok and Schwark2020b; Ruebsam & Schwark, Reference Ruebsam, Schwark, Reolid, Duarte, Mattioli and Ruebsam2021). It is argued that volcanic CO2 emissions from the Karoo–Ferrar LIPs initiated global warming that destabilized the ice caps and cryosphere-stored carbon reservoirs. Massive carbon release from the cryosphere, including highly 13C-depleted CH4, combined with orbitally forced climate cycles accelerated global warming. This caused a runaway effect that drove Earth’s climate system from an icehouse in the late Pliensbachian into a prolonged greenhouse mode in the Toarcian. Deglaciation and ocean warming were responsible for eustatic sea-level rise.

21.f. Middle Toarcian – continuing oxygen depletion

Relative sea-level rise in the Cleveland Basin that began at the Pliensbachian – Toarcian boundary reached a maximum for the Early Jurassic around the H. serpentinumH. bifrons zone boundary (Hesselbo, Reference Hesselbo2008), coincident with maximum flooding linked to the Pliensbachian – Toarcian 2nd order Boreal standard sequence of de Graciansky et al. (Reference de Graciansky, Dardeau, Dommergues, Durlet, Marchand, Dumont, Hesselbo, Jacquin, Goggin, Meister, Mouterde, Rey, Vail, de Graciansky, Hardenbol, Jacquin and Vail1998). This reflects the eustatic sea-level trend. An onset of regression during the early H. falciferum Subzone in the Cleveland Basin is indicated by the beginning of gradually rising Ti/Al ratios and quartz mean grain-size indicative of upward coarsening of grain sizes in the Mulgrave and Alum shales. Maximum flooding associated with third-order sequences in the NW and SW German basins has also been recorded in the late H. serpentinum Zone (e.g. Arp et al., Reference Arp, Gropengießer, Schulbert, Jung and Reimer2021, Reference Arp, Balmuk, Seppelt and Reimer2023). An interval of coarser planar laminated mudstones was observed immediately above the Hard Shales in the Dove’s Nest core (Trabucho-Alexandre et al., Reference Trabucho-Alexandre, Gröcke, Atar, Herringshaw and Jarvis2022) but these are not associated with a significant shift in the geochemical profiles. These may reflect a period of increased current action but, nonetheless, water depths in the study area likely remained generally below storm wave-base (i.e., > 30 m).

Two episodes of renewed anoxia and intermittent euxinia characterize the earliest middle Toarcian D. commune Subzone in the Cleveland Basin (Hard Shales, Subunit Vb) evidenced by levels of increased TOC content and trace metal enrichment. TOC/PT and DOPT proxies demonstrate dysoxia persisting into the middle Toarcian. A gradual return to better oxygenation conditions in the middle of the subzone is indicated by an increase in body fossils and macrofossil diversity, the disappearance of low-oxygen specialist bivalves and the reappearance of trace fossils at the top of the study interval. Nonetheless, the benthos had still not recovered to late Pliensbachian pre-OAE state by the P. fibulatum Subzone (mid-H. bifrons Zone) (Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023) indicating that dysoxic anoxic bottom waters persisted, at least intermittently.

Extended intervals displaying an absence of macroscopic bioturbation, indicative of bottom water anoxia – euxinia during the T-OAE and later early Toarcian, have been observed from sections in NW Tethys, the North African (North Gondwana Palaeomargin) and northern Iberian margins, Panthalassa Ocean sites in Alberta, Canada and the Arctic Basin, northern Siberia (Caswell & Herringshaw, Reference Caswell, Herringshaw, Nawrot, Dominici, Tomašových and Zuschin2023). In the western Canadian and Arctic basins, as in the Cleveland Basin, macroscopic trace fossils were absent for much longer (D. tenuicostatum – lower H. bifrons zones). A cessation of bottom sediment irrigation would have significantly impacted biogeochemical dynamics across the benthic boundary layer, favouring the preservation of organic matter and the release of phosphate and ammonia to bottom waters.

Increased oxygenation in the Cleveland Basin during the early middle Toarcian H. bifrons Zone may be related to an increased influx of cool low-salinity Arctic water through the Viking Corridor, evidenced by the southward spread of high-latitude phytoplankton (Parvocysta – Phallocysta dinocyst suite) following the T-OAE (van de Schootbrugge et al., Reference van de Schootbrugge, Houben, Ercan, Verreussel, Kerstholt, Janssen, Nikitenko and Suan2020). Palynological evidence indicates the transition to a cooler more temperate terrestrial vegetation (Slater et al., Reference Slater, Twitchett, Danise and Vajda2019) although belemnite oxygen isotope and Mg/Ca ratios point to continuing warm waters in the basin.

22. Conclusions

Well-exposed upper Pliensbachian – middle Toarcian sections in the Cleveland Basin along the North Yorkshire coast and core from the Dove’s Nest borehole display a thick succession of carbonaceous mudstones incorporating a large negative δ13C excursion (the T-OAE). They represent a typical expression of the T-OAE in the Boreal epicontinental seas of Europe bordering the NW Tethys Ocean although, in contrast to successions like the Posidonia Shale of the SW German Basin (e.g. at Dotternhausen), periodic intervals showing macroscopic bioturbation are absent. A uniquely comprehensive suite of palaeontological and geochemical data (elemental, isotopic and organic) available from the Yorkshire succession offers a global standard for constraining palaeoenvironmental change preceding, during and following the T-OAE.

Oxic conditions prevailed in the Cleveland Basin during the late Pliensbachian, with fossil evidence and geochemical sediment redox proxies indicating an increasing influence of dysoxic and anoxic bottom waters during the earliest Toarcian, culminating in extended euxinia at the height of the T-OAE during the late D. semicelatum and C. exaratum subzones. Three short intervals of euxinia preceded the T-OAE, the oldest two occurring during the Pliensbachian – Toarcian Boundary Event, a prominent earlier negative δ13C excursion. Biomarkers point to periods of PZE during these events and the T-OAE. Generally, dysoxic – anoxic, temporarily euxinic, bottom-water conditions persisted into the early middle Toarcian D. commune Subzone.

Redox changes in the Cleveland Basin are representative of a global pattern. An ε205Tl increase of −6 to −4 in Pliensbachian – Toarcian sediments of the Western Canada Sedimentary Basin is attributed to a global reduction in Fe–Mn oxyhydroxide precipitation accompanying oxygen depletion. The Tl-isotope data indicate that increasing global bottom-water anoxia began at the Pliensbachian – Toarcian boundary and was sustained into the middle Toarcian H. bifrons Zone, representing an interval of ∼2 Ma.

Redox-sensitive trace-metal and isotopic trends recorded in the Yorkshire mudstones principally reflect changes in global ocean chemistry with a secondary influence of local factors, including episodes of enhanced local basin restriction during the T-OAE. Previous suggestions of severe basin restriction and the establishment of freshwater conditions during the T-OAE are untenable. Elevated Mo, U, V and other trace-metal enrichment factors but low trace-metal/TOC ratios in T-OAE black shales reflect the global drawdown of trace metals from ocean water driven by an expansion of anoxic and euxinic seafloor area in both epicontinental seas and the deep ocean.

Globally, carbon burial rates in anoxic – euxinic basins may have increased ∼500% on average during the T-OAE. Nonetheless, despite the prominence of black shale facies including the Jet Rock, Posidonia Shale and Schistes carton that characterize the T-OAE through much of NW Europe, mass balance calculations using organic matter and pyrite contents and sulfur isotope data indicate that the area provided only a minor contribution (< 10%) to the geochemical changes effecting the global ocean.

Consensus on the duration of the T-OAE negative excursion remains elusive with estimates of ∼400 – 500 ka, ∼800 – 900 ka, ∼1 Ma or ∼1.2 Ma commonly quoted. A value of 1.4 Ma is proposed here. Nonetheless, calculated rates of organic carbon burial during the T-OAE are generally low relative to modern shallow-water settings indicating that preservation rather than productivity provided the dominant control on black shale accumulation during the early Toarcian.

Modelling of the δ34S increase accompanying the T-OAE incorporating drawdown of isotopically light 32S by pyrite and organic sulfur burial in euxinic environments indicates that pyrite deposition in northern Europe accounted for at most 4% of the pyrite burial. Substantial additional pyrite burial is needed to drive the documented S-isotope excursion, requiring a much greater extent of euxinic conditions in the world ocean during the T-OAE, although documented heterogeneity in Toarcian seawater sulfate-δ34S poses challenges for estimating global reduced-sulfur burial fluxes and the extent of oceanic anoxia.

Rhenium and Mo oceanic mass balance models incorporating data from anoxic or euxinic carbonaceous mudstones deposited in unrestricted marine settings in western Canada indicate an expansion of up to ∼7% total global seafloor anoxia – euxinia, dominated by euxinia, during the early stages of the T-OAE. Similarly, estimates of organic carbon burial derived from open ocean records of Mo drawdown into reducing sediments during the T-OAE require that ≥ 3% of the ocean floor was covered by euxinic bottom waters compared to < 0.3% today.

Water temperatures rose sharply by 6 – 7° C in the Cleveland Basin during the onset of the T-OAE, reflecting a global hyperthermal event. Increased temperature was likely accompanied by a reduction in salinity of ∼2 psu due to regional changes in ocean circulation and a wetter climate. Water temperatures peaked in the early C. exaratum Subzone coincident with the δ13Corg minimum (−32‰) and maximum TOC values (10%) that accompanied peak bottom-water euxinia evidenced by TOC/PT, DOPT, iron speciation, Re/Mo and other trace-metal data, an absence of bioturbation and a benthic fauna limited to intermittent influxes of low-oxygen specialist bivalves. Temperatures declined in the H. falciferum Subzone immediately following the T-OAE but remained several degrees higher through the early – mid-Toarcian compared to before the event. A regional change from a cool moist climate during the late Pliensbachian to a hot climate with extreme wet/dry seasons of the hyperthermal back to a warm strongly seasonal climate following the T-OAE is indicated by palynological data.

The biotic changes documented in detail from the Cleveland Basin reflect the global impact of the T-OAE. A stepped second-order mass extinction principally affecting benthic taxa resulted from the global expansion of anoxic – euxinic bottom waters in epicontinental seas, on shelves and in the deep ocean, with additional stress imposed by substantial global warming and climate change.

Increased volcanic emissions, principally CO2, provided the primary climate driver. The peak of the main magmatic phase of the Karoo LIP coincided with the acceleration of early Toarcian global warming and immediately pre-dates the beginning of the T-OAE, with volcanic activity continuing during and immediately following the event. Like the Karoo, Ferrar-LIP activity displayed multiple magmatic pulses, contributing CO2 that likely impacted both the Pliensbachian – Toarcian Boundary Event and the T-OAE. A significant CO2 contribution from the Chon Aike magmatic province is considered to be unlikely.

An accelerated hydrological cycle accompanied the T-OAE hyperthermal. A large positive 187Os/188Osi excursion in the T-OAE interval in the Cleveland Basin is consistent with a 400 – 800% increase in continental weathering rates. A coarsening-upward cycle (increasing Ti/Al ratio) and a peak in K/Al ratio reflecting a pulse of illitic clay coincident with the Os-isotope excursion further evidence the event in the Yorkshire succession. Steeply rising 87Sr/86Sr ratios through the C. exaratum Subzone demonstrate a high flux of radiogenic Sr from the chemical weathering of continental crust, although the stratigraphic trend is influenced by reduced sedimentation rates.

187Os/188Osi and δ13Corg data from Japan, Europe and North America display patterns consistent with the Yorkshire profiles and indicate an increased global weathering rate of up to 600% through the entire T-OAE, although variation in the amplitude of the osmium isotope excursions points to inhomogeneity in the Os-isotopic composition of the Toarcian seawater. Silicate weathering and carbon burial sequestering CO2 provided negative feedback that combined with decreasing volcanism to terminate the T-OAE hyperthermal.

Large variations in the amplitude of the negative δ13C excursions recorded across the T-OAE in different archives and at different sites can be attributed to multiple factors including large stratigraphic and geographic changes in the type, proportions and composition of terrestrial and marine organic matter, regional and depth variation in water column DIC δ13C, mineralogical and sedimentological differences in carbonate composition and diagenesis. A value of 3 – 4‰ δ13C provides the best estimate for the primary decrease in global surface carbon reservoirs accompanying the T-OAE.

In addition to volcanic CO2 the contribution of a highly 13C-depleted source, likely biogenic CH4, is required to explain the rapid fall and large amplitude of the negative δ13C excursion characterizing the T-OAE. An initial modest rise in temperature caused by volcanic CO2 released in the early Toarcian caused the dissociation of terrestrial and seafloor methane clathrates, CH4 release and further warming. A melt-down of Earth’s cryosphere that had expanded during the icehouse episode of the late Pliensbachian, and the rapid release of greenhouse gases, mainly as 13C-depleted CH4, drove the negative δ13C excursion. However, the volume of greenhouse gas emissions may have been underestimated.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S0016756824000244

Data availability

All data generated during this study are included in the Supplementary Material files for this article.

Acknowledgements

Thanks are due to Hugh Jenkyns (Oxford University), Guillaume Suan (Université de Lyon), two anonymous referees and editor Emese Bordy (University of Cape Town) for their thoughtful comments and suggestions that greatly improved the quality of this manuscript. We thank Anglo American Woodsmith Project (formerly Sirius Minerals) for allowing us to study and sample the Dove’s Nest core; Howard Armstrong for help describing and sampling the core; Simon DeMars, Kate Olde and Julian Swinden provided invaluable assistance during analytical work by EA in the Kingston laboratories. Analytical costs were supported by the UK Natural Environment Research Council (I.J. grant number NE/H020756/1), (D.G. grant number NE/H021868/1).

Author contributions

IJ: conceptualization (equal), data curation (equal), formal analysis (equal), funding acquisition (equal), methodology (lead), validation (equal), visualization (lead), writing — original draft (lead), review & editing (lead); EA: conceptualization (equal), data curation (equal), formal analysis (equal), investigation (lead), methodology (supporting), validation (equal) and writing — review & editing (supporting); DRG: formal analysis (supporting), funding acquisition (equal) and methodology (supporting); LGH: conceptualization (supporting), formal analysis (supporting) and writing — review & editing (supporting); JPTA: conceptualization (equal), formal analysis (supporting), methodology (supporting) and writing — review & editing (supporting).

Competing interests

The authors have no potential conflicts of interest.

References

Aberhan, M and Baumiller, TZ (2003) Selective extinction among Early Jurassic bivalves: a consequence of anoxia. Geology 31, 10771080.CrossRefGoogle Scholar
Abubakar, Y, Taylor, KG, Coker, V, Wogelius, RA and van Dongen, BE (2022) Fundamental controls on organic matter preservation in organic- and sulfur-rich hydrocarbon source rocks. Marine and Petroleum Geology 141, 105684.CrossRefGoogle Scholar
Adams, DD, Hurtgen, MT and Sageman, BB (2010) Volcanic triggering of a biogeochemical cascade during Oceanic Anoxic Event 2. Nature Geoscience 3, 201204.CrossRefGoogle Scholar
Ader, M, Sansjofre, P, Halverson, GP, Busigny, V, Trindade, RIF, Kunzmann, M and Nogueira, ACR (2014) Ocean redox structure across the Late Neoproterozoic Oxygenation Event: a nitrogen isotope perspective. Earth and Planetary Science Letters 396, 113.CrossRefGoogle Scholar
Agbi, I, Ozibo, B and Newton, R (2015) Pyrite framboid size distribution of the Grey Shales (Yorkshire UK) as an indication of redox conditions. IOSR Journal of Applied Geology and Geophysics 3, 3642.Google Scholar
Aggett, JR (1990) The sedimentology, mineralogy and geochemistry of the Frodingham ironstone formation : implications for the genesis of ooidal ironstones. PhD thesis, University of Manchester, Manchester, 549 pp. Published thesis https://uomlibrary.access.preservica.com/uncategorized/IO_82f89037-72ca-469d-b569-a65e18cedea1.Google Scholar
Ait-Itto, F-Z, Martinez, M, Price, GD and Addi, AA (2018) Synchronization of the astronomical time scales in the Early Toarcian: a link between anoxia, carbon-cycle perturbation, mass extinction and volcanism. Earth and Planetary Science Letters 493, 111.CrossRefGoogle Scholar
Ait-Itto, F-Z, Price, GD, Ait Addi, A, Chafiki, D and Mannani, I (2017) Bulk-carbonate and belemnite carbon-isotope records across the Pliensbachian–Toarcian boundary on the northern margin of Gondwana (Issouka, Middle Atlas, Morocco). Palaeogeography, Palaeoclimatology, Palaeoecology 466, 128136.CrossRefGoogle Scholar
Aitchison, J (1986) The statistical analysis of compositional data. New York: Chapman & Hall, 416 pp.CrossRefGoogle Scholar
Ajuaba, S, Sachsenhofer, RF, Bechtel, A, Galasso, F, Gross, D, Misch, D and Schneebeli-Hermann, E (2022) Biomarker and compound-specific isotope records across the Toarcian CIE at the Dormettingen section in SW Germany. International Journal of Earth Sciences 111, 16311661.CrossRefGoogle ScholarPubMed
Al-Suwaidi, AH, Ruhl, M, Jenkyns, HC, Damborenea, SE, Manceñido, MO, Condon, DJ, Angelozzi, GN, Kamo, SL, Storm, M, Riccardi, AC and Hesselbo, SP (2022) New age constraints on the Lower Jurassic Pliensbachian–Toarcian Boundary at Chacay Melehue (Neuquén Basin, Argentina). Scientific Reports 12, 4975.CrossRefGoogle ScholarPubMed
Algeo, TJ (2004) Can marine anoxic events draw down the trace element inventory of seawater? Geology 32, 10571060.CrossRefGoogle Scholar
Algeo, TJ and Ingall, E (2007) Sedimentary Corg: P ratios, paleocean ventilation, and Phanerozoic atmospheric pO2 . Palaeogeography, Palaeoclimatology, Palaeoecology 256, 130155.CrossRefGoogle Scholar
Algeo, TJ and Li, C (2020) Redox classification and calibration of redox thresholds in sedimentary systems. Geochimica et Cosmochimica Acta 287, 826.CrossRefGoogle Scholar
Algeo, TJ and Liu, J (2020) A re-assessment of elemental proxies for paleoredox analysis. Chemical Geology 540, 112.CrossRefGoogle Scholar
Algeo, TJ and Lyons, TW (2006) Mo-total organic carbon covariation in modern anoxic marine environments: implications for analysis of paleoredox and paleohydrographic conditions. Paleoceanography 21, 123.CrossRefGoogle Scholar
Algeo, TJ and Maynard, JB (2004) Trace-element behavior and redox facies in core shales of Upper Pennsylvanian Kansas-type cyclothems. Chemical Geology 206, 289318.CrossRefGoogle Scholar
Algeo, TJ, Meyers, PA, Robinson, RS, Rowe, H and Jiang, GQ (2014) Icehouse–greenhouse variations in marine denitrification. Biogeosciences 11, 12731295.CrossRefGoogle Scholar
Algeo, TJ and Rowe, H (2012) Paleoceanographic applications of trace-metal concentration data. Chemical Geology 324–325, 618.CrossRefGoogle Scholar
Algeo, TJ, Rowe, H, Hower, JC, Schwark, L, Merrmann, A and Heckel, P (2008) Changes in ocean denitrification during Late Carboniferous glacial–interglacial cycles. Nature Geoscience 1, 709714.CrossRefGoogle Scholar
Algeo, TJ and Tribovillard, N (2009) Environmental analysis of paleoceanographic systems based on molybdenum–uranium covariation. Chemical Geology 268, 211225.CrossRefGoogle Scholar
Alnazghah, M, Koeshidayatullah, A, Al-Hussaini, A, Amao, A, Song, H and Al-Ramadan, K (2022) Evidence for the early Toarcian Carbon Isotope Excursion (T-CIE) from the shallow marine siliciclastic red beds of Arabia. Scientific Reports 12, 18124.CrossRefGoogle ScholarPubMed
Anbar, AD, Creaser, RA, Papanastassiou, DA and Wasserburg, GJ (1992) Rhenium in seawater: confirmation of generally conservative behavior. Geochimica et Cosmochimica Acta 56, 40994103.CrossRefGoogle Scholar
Arnold, GL, Anbar, AD, Barling, J and Lyons, TW (2004) Molybdenum isotope evidence for widespread anoxia in mid-Proterozoic oceans. Science 304, 8790.CrossRefGoogle ScholarPubMed
Arp, G, Balmuk, Y, Seppelt, S and Reimer, A (2023) Biostratigraphy and sedimentary sequences of the Toarcian Hainberg section (northwestern Harz foreland, northern Germany). Zitteliana 97, 127.CrossRefGoogle Scholar
Arp, G, Gropengießer, S, Schulbert, C, Jung, D and Reimer, A (2021) Biostratigraphy and sequence stratigraphy of the Toarcian Ludwigskanal section (Franconian Alb, Southern Germany). Zitteliana 95, 5794.CrossRefGoogle Scholar
Atar, E (2015) Inorganic geochemistry and palaeoenvironments of the Early Jurassic Cleveland Basin. MSc thesis, Durham University, Durham, 93 pp. Published thesis http://etheses.dur.ac.uk/10948/.Google Scholar
Atar, E, Marz, C, Aplin, AC, Dellwig, O, Herringshaw, LG, Lamoureux-Var, V, Leng, MJ, Schnetger, B and Wagner, T (2019a) Dynamic climate-driven controls on the deposition of the Kimmeridge Clay Formation in the Cleveland Basin, Yorkshire, UK. Climate of the Past 15, 15811601.CrossRefGoogle Scholar
Atar, E, Marz, C, Schnetger, B, Wagner, T and Aplin, A (2019b) Local to global controls on the deposition of organic-rich muds across the Late Jurassic Laurasian Seaway. Journal of the Geological Society 176, 11431153.CrossRefGoogle Scholar
Atkinson, JW, Little, CTS and Dunhill, AM (2023) Long duration of benthic ecological recovery from the early Toarcian (Lower Jurassic) mass extinction event in the Cleveland Basin, UK. Journal of the Geological Society 180, jgs20222126.CrossRefGoogle Scholar
Baghli, H, Mattioli, E, Spangenberg, JE, Ruebsam, W, Schwark, L, Bensalah, M, Sebane, A, Pittet, B, Pellenard, P and Suan, G (2022) Stratification and productivity in the Western Tethys (NW Algeria) during early Toarcian. Palaeogeography, Palaeoclimatology, Palaeoecology 591, 110864.CrossRefGoogle Scholar
Bailey, TR, Rosenthal, Y, McArthur, JM, van de Schootbrugge, B and Thirlwall, MF (2003) Paleoceanographic changes of the Late Pliensbachian–Early Toarcian interval: a possible link to the genesis of an Oceanic Anoxic Event. Earth and Planetary Science Letters 212, 307320.CrossRefGoogle Scholar
Baker, SJ, Hesselbo, SP, Lenton, TM, Duarte, LV and Belcher, CM (2017) Charcoal evidence that rising atmospheric oxygen terminated Early Jurassic ocean anoxia. Nature Communications 8, 15018.CrossRefGoogle ScholarPubMed
Bambach, RK (2006) Phanerozoic biodiversity mass extinctions. Annual Review of the Earth and Planetary Sciences 34, 127155.CrossRefGoogle Scholar
Banta, AB, Wei, JH and Welander, PV (2015) A distinct pathway for tetrahymanol synthesis in bacteria. PNAS 112, 1347813483.CrossRefGoogle ScholarPubMed
Baranyi, V, Jin, X, Dal Corso, J, Li, B and Kemp, DB (2024) Vegetation response to climate change during an Early Jurassic hyperthermal event (Jenkyns Event) from Northern China (Ordos Basin). Palaeogeography, Palaeoclimatology, Palaeoecology 643, 112180.CrossRefGoogle Scholar
Baranyi, V, Jin, X, Dal Corso, J, Shi, Z, Grasby, SE and Kemp, DB (2023) Collapse of terrestrial ecosystems linked to heavy metal poisoning during the Toarcian oceanic anoxic event. Geology 51, 652656.CrossRefGoogle Scholar
Barnard, PC and Cooper, BS (1983) A review of geochemical data related to the northwest European gas province. In Petroleum Geochemistry and Exploration of Europe (ed Brooks, J), pp. 1933. Geological Society London, Special Publications 12.Google Scholar
Barnes, CE and Cochran, JK (1990) Uranium removal in oceanic sediments and the oceanic U balance. Earth and Planetary Science Letters 97, 94101.CrossRefGoogle Scholar
Bastias, J, Spikings, R, Riley, T, Ulianov, A, Grunow, A, Chiaradia, M and Hervé, F (2021) A revised interpretation of the Chon Aike magmatic province: active margin origin and implications for the opening of the Weddell Sea. Lithos 386–387, 106013.CrossRefGoogle Scholar
Beerling, DJ, Lomas, MR and Gröcke, DR (2002) On the nature of methane gas-hydrate dissociation during the Toarcian and Aptian Oceanic Anoxic Events. American Journal of Science 302, 2849.CrossRefGoogle Scholar
Benamara, A, Charbonnier, G, Adatte, T, Spangenberg, JE and Föllmi, KB (2020) Precession-driven monsoonal activity controlled the development of the early Albian Paquier oceanic anoxic event (OAE1b): evidence from the Vocontian Basin, SE France. Palaeogeography, Palaeoclimatology, Palaeoecology 537, 109406.CrossRefGoogle Scholar
Bennett, WW and Canfield, DE (2020) Redox-sensitive trace metals as paleoredox proxies: a review and analysis of data from modern sediments. Earth-Science Reviews 204, 103175.CrossRefGoogle Scholar
Bergman, SC, Eldrett, JS and Minisini, D (2021) Phanerozoic Large Igneous Province, petroleum system, and source rock links. In Large Igneous Provinces: A Driver of Global Environmental and Biotic Changes (eds Ernst, RE, Dickson, AJ and Bekker, A). Geophysical Monograph Series, pp. 191228. Hoboken NJ: American Geophysical Union and John Wiley and Sons, Inc.CrossRefGoogle Scholar
Berner, RA (1970) Sedimentary pyrite formation. American Journal of Science 268, 123.CrossRefGoogle Scholar
Bjerrum, CJ, Surlyk, F, Callomon, JH and Slingerland, RL (2001) Numerical paleoceanographic study of the Early Jurassic Transcontinental Laurasian Seaway. Paleoceanography 16, 390404.CrossRefGoogle Scholar
Blakey, R (2012) Paleogeography of Europe series, Cretaceous ca. 75 Ma. Colorado Plateau Geosystems (now DeepTimeMapsTM), Flagstaff AZ. https://deeptimemaps.com.Google Scholar
Blakey, R (2016) Mesozoic: 180 Ma Moll_Jur, Global Paleogeography and Tectonics in Deep Time. Colorado Plateau Geosystems (now DeepTimeMapsTM), Flagstaff AZ. https://deeptimemaps.com.Google Scholar
Bodin, S, Fantasia, A, Krencker, F-N, Nebsbjerg, B, Christiansen, L and Andrieu, S (2023) More gaps than record! A new look at the Pliensbachian/Toarcian boundary event guided by coupled chemo-sequence stratigraphy. Palaeogeography, Palaeoclimatology, Palaeoecology 610, 111344.CrossRefGoogle Scholar
Bodin, S, Godet, A, Matera, V, Steinmann, P, Vermeulen, J, Gardin, S, Adatte, T, Coccioni, R and Föllmi, KB (2007) Enrichment of redox-sensitive trace metals (U, V, Mo, As) associated with the late Hauterivian Faraoni oceanic anoxic event. International Journal of Earth Sciences (Geologisches Rundschau) 96, 327341.CrossRefGoogle Scholar
Bomou, B, Suan, G, Schlögl, J, Grosjean, A-S, Suchéras-Marx, B, Adatte, T, Spangenberg, J, Fouché, S, Zacaï, A, Gibert, C, Brazier, J-M, Perrier, V, Vincent, P, Janneau, K and Martin, JE (2021) The palaeoenvironmental context of Toarcian vertebrate yielding shales of southern France (Hérault). In Carbon Cycle and Ecosystem Response to the Jenkyns Event in the Early Toarcian (Jurassic) (eds Reolid, M, Duarte, LV, Mattioli, E and Ruebsam, W), pp. 121152. Geological Society London, Special Publications 514.Google Scholar
Bond, DPG, Wignall, PB, Wang, W, Izon, G, Jiang, HS, Lai, XL, Sun, YD, Newton, RJ, Shao, LY, Vedrine, S and Cope, H (2010) The mid-Capitanian (Middle Permian) mass extinction and carbon isotope record of South China. Palaeogeography, Palaeoclimatology, Palaeoecology 292, 282294.CrossRefGoogle Scholar
Bougeault, C, Pellenard, P, Deconinck, JF, Hesselbo, SP, Dommergues, JL, Bruneau, L, Cocquerez, T, Laffont, R, Huret, E and Thibault, N (2017) Climatic and palaeoceanographic changes during the Pliensbachian (Early Jurassic) inferred from clay mineralogy and stable isotope (C–O) geochemistry (NW Europe). Global and Planetary Change 149, 139152.CrossRefGoogle Scholar
Boulila, S, Galbrun, B, Huret, E, Hinnov, LA, Rouget, I, Gardin, S and Bartolini, A (2014) Astronomical calibration of the Toarcian Stage: implications for sequence stratigraphy and duration of the early Toarcian OAE. Earth and Planetary Science Letters 386, 98111.CrossRefGoogle Scholar
Boulila, S, Galbrun, B, Sadki, D, Gardin, S and Bartolini, A (2019) Constraints on the duration of the early Toarcian T-OAE and evidence for carbon-reservoir change from the High Atlas (Morocco). Global and Planetary Change 175, 113128.CrossRefGoogle Scholar
Boulila, S and Hinnov, LA (2017) A review of tempo and scale of the early Jurassic Toarcian OAE: implications for carbon cycle and sea level variations. Newsletters on Stratigraphy 50, 363389.CrossRefGoogle Scholar
Bour, I, Mattioli, E and Pittet, B (2007) Nannofacies analysis as a tool to reconstruct paleoenvironmental changes during the Early Toarcian anoxic event. Palaeogeography, Palaeoclimatology, Palaeoecology 249, 5879.CrossRefGoogle Scholar
Bowden, SA, Farrimond, P, Snape, CE and Love, GD (2006) Compositional differences in biomarker constituents of the hydrocarbon, resin, asphaltene and kerogen fractions: an example from the Jet Rock (Yorkshire, UK). Organic Geochemistry 37, 369383.CrossRefGoogle Scholar
Bown, PR, Lees, JA and Young, JR (2004) Calcareous nannoplankton evolution and diversity through time. In Coccolithophores: from Molecular Processes to Global Impact (eds Thierstein, HR and Young, JR), pp. 481508. Berlin Heidelberg: Springer-Verlag.CrossRefGoogle Scholar
Bradshaw, MJ, Cope, JCW, Cripps, DW, Donovan, DT, Howarth, MK, Rawson, PF, West, IM and Wimbledon, WA (1992) Jurassic. In Atlas of Palaeogeography and Lithofacies (eds JCW Cope, JK Ingham and PF Rawson). Geological Society London, Memoir 13, 107129.CrossRefGoogle Scholar
Brański, P (2012) The mineralogical record of the Early Toarcian stepwise climate changes and other environmental variations (Ciechocinek Formation, Polish Basin). Volumina Jurassica 10, 124.Google Scholar
Brantley, SL, Holleran, ME, Jin, L and Bazilevskaya, E (2013) Probing deep weathering in the Shale Hills Critical Zone Observatory, Pennsylvania (USA): the hypothesis of nested chemical reaction fronts in the subsurface. Earth Surface Processes and Landforms 38, 12801298.CrossRefGoogle Scholar
Brazier, JM, Suan, G, Tacail, T, Simon, L, Martin, JE, Mattioli, E and Balter, V (2015) Calcium isotope evidence for dramatic increase of continental weathering during the Toarcian oceanic anoxic event (Early Jurassic). Earth and Planetary Science Letters 411, 164176.CrossRefGoogle Scholar
Breit, GN and Wanty, RB (1991) Vanadium accumulation in carbonaceous rocks: a review of geochemical controls during deposition and diagenesis. Chemical Geology 91, 8397.CrossRefGoogle Scholar
Bruland, KW and Lohan, MC (2003) Controls of trace metals in seawater. In The Oceans and Marine Geochemistry (ed Elderfield, H), Treatise on Geochemistry 6, pp. 2347. Amsterdam: Elsevier.Google Scholar
Brumsack, H-J (1989) Seawater chemistry - update on trace-metal data. Naturwissenschaften 76, 99106.CrossRefGoogle Scholar
Brumsack, H-J (2006) The trace metal content of recent organic carbon-rich sediments: implications for Cretaceous black shale formation. Palaeogeography, Palaeoclimatology, Palaeoecology 232, 344361.CrossRefGoogle Scholar
Bucefalo Palliani, R, Mattioli, E and Riding, JB (2002) The response of marine phytoplankton and sedimentary organic matter to the early Toarcian (Lower Jurassic) oceanic anoxic event in northern England. Marine Micropaleontology 46, 223245.CrossRefGoogle Scholar
Buckman, SS (1915) A palaeontological classification of the Jurassic rocks of the Whitby district, with a zonal table of Liassic ammonites. In The Geology of the Country between Whitby and Scarborough (eds Fox-Strangways, C and Barrow, G) Memoirs of the Geological Survey, England and Wales, pp. 59102. London: HMSO.Google Scholar
Burnaz, L, Littke, R, Grohmann, S, Erbacher, J, Strauss, H and Amann, F (2024). Lower Jurassic (Pliensbachian–Toarcian) marine paleoenvironment in Western Europe: sedimentology, geochemistry and organic petrology of the wells Mainzholzen and Wickensen, Hils Syncline, Lower Saxony Basin. International Journal of Earth Sciences (Geologische Rundschau) https://doi.org/10.1007/s00531-023-02381-8.CrossRefGoogle Scholar
Caldwell Steele, S (2020) Is Whitby Jet Monkey Puzzle wood? Ebor Jetworks Blog. Available at https://whatiswhitbyjet.com/2020/08/14/is-whitby-jet-monkey-puzzle/ 14/08/2020 (accessed 14/11/2023).Google Scholar
Calvert, SE and Pedersen, TF (1993) Geochemistry of Recent oxic and anoxic marine sediments: implications for the geological record. Marine Geology 13, 6788.CrossRefGoogle Scholar
Calvert, SE and Pedersen, TF (1996) Sedimentary geochemistry of manganese: implications for the environment of formation of manganiferous black shales. Economic Geology 91, 3647.CrossRefGoogle Scholar
Calvert, SE and Pedersen, TF (2007) Elemental proxies for palaeoclimatic and palaeoceanographic variability in marine sediments: Interpretation and application. In Proxies in Late Cenozoic Paleoceanography (eds Hillaire-Marcel, C and de Vernal, A), pp. 568644. Developments in Marine Geology 1, Amsterdam: Elsevier.Google Scholar
Campbell, CV (1967) Lamina, laminaset, bed and bedset. Sedimentology 8, 726.CrossRefGoogle Scholar
Canfield, DE (1994) Factors influencing organic carbon preservation in marine sediments. Chemical Geology 114, 315329.CrossRefGoogle ScholarPubMed
Caruthers, AH, Smith, PL and Gröcke, DR (2013) The Pliensbachian–Toarcian (Early Jurassic) extinction, a global multi-phased event. Palaeogeography, Palaeoclimatology, Palaeoecology 386, 104118.CrossRefGoogle Scholar
Caruthers, AH, Smith, PL and Gröcke, DR (2014) The Pliensbachian–Toarcian (Early Jurassic) extinction: a North American perspective. Geological Society of America, Special Paper 505, 225243.Google Scholar
Caswell, BA and Coe, AL (2013) Primary productivity controls on opportunistic bivalves during Early Jurassic oceanic deoxygenation. Geology 41, 11631166.CrossRefGoogle Scholar
Caswell, BA and Coe, AL (2014) The impact of anoxia on pelagic macrofauna during the Toarcian Oceanic Anoxic Event (Early Jurassic). Proceedings of the Geologists’ Association 125, 383391.CrossRefGoogle Scholar
Caswell, BA, Coe, AL and Cohen, AS (2009) New range data for marine invertebrate species across the early Toarcian (Early Jurassic) mass extinction. Journal of the Geological Society 166, 859872.CrossRefGoogle Scholar
Caswell, BA and Dawn, SJ (2019) Recovery of benthic communities following the Toarcian oceanic anoxic event in the Cleveland Basin, UK. Palaeogeography, Palaeoclimatology, Palaeoecology 521, 114126.CrossRefGoogle Scholar
Caswell, BA and Frid, CLJ (2017) Marine ecosystem resilience during extreme deoxygenation: the Early Jurassic oceanic anoxic event. Oecologia 183, 275290.CrossRefGoogle ScholarPubMed
Caswell, BA and Herringshaw, L (2023) Marine bioturbation collapse during Early Jurassic deoxygenation: implications for post-extinction marine ecosystem functioning. In Conservation Palaeobiology of Marine Ecosystems (eds Nawrot, R, Dominici, S, Tomašových, A and Zuschin, M), pp. 311344. Geological Society of London, Special Publications 529.Google Scholar
Catt, JA, Gad, MA, Le Riche, HH and Lord, AR (1971) Geochemistry, micropalaeontology and origin of the Middle Lias ironstones in northeast Yorkshire (Great Britain). Chemical Geology 8, 6176.CrossRefGoogle Scholar
Chamley, H (1989) Clay Sedimentology. Berlin, Heidelberg: Springer-Verlag, 623 pp.CrossRefGoogle Scholar
Chen, W, Kemp, DB, He, T, Newton, RJ, Xiong, Y, Jenkyns, HC, Izumi, K, Cho, T, Huang, C and Poulton, SW (2023) Shallow-and deep-ocean Fe cycling and redox evolution across the Pliensbachian–Toarcian boundary and Toarcian Oceanic Anoxic Event in Panthalassa. Earth and Planetary Science Letters 602, 112.CrossRefGoogle Scholar
Chen, WH, Kemp, DB, He, TC, Huang, CJ, Jin, SM, Xiong, YJ and Newton, RJ (2021) First record of the early Toarcian Oceanic Anoxic Event in the Hebrides Basin (UK) and implications for redox and weathering changes. Global and Planetary Change 207, 103685.CrossRefGoogle Scholar
Cheng, K, Elrick, M and Romaniello, SJ (2020) Early Mississippian ocean anoxia triggered organic carbon burial and late Paleozoic cooling: evidence from uranium isotopes recorded in marine limestone. Geology 48, 363367.CrossRefGoogle Scholar
Chowns, TM (1966) Depositional environment of the Cleveland Ironstone Series. Nature 211, 12861287.CrossRefGoogle Scholar
Chowns, TM (1968) Environmental and diagenetic studies of the Cleveland Ironstone Formation of north east Yorkshire. PhD thesis, Newcastle University, Newcastle upon Tyne, 432 pp. Published thesis http://theses.ncl.ac.uk/jspui/handle/10443/268.Google Scholar
Clarkson, MO, Stirling, CH, Jenkyns, HC, Dickson, AJ, Porcelli, D, Moy, CM, Pogge von Strandman, PAE, Cooke, IR and Lenton, TM (2018) Uranium isotope evidence for two episodes of deoxygenation during Oceanic Anoxic Event 2. Proceedings of the National Academy of Sciences 115, 29182923.CrossRefGoogle ScholarPubMed
Claypool, GE, Holzer, WT, Kaplan, IR, Sakai, H and Zak, I (1980) The age curve of sulphur and oxygen isotopes in marine sulphates and their mutual interpretation. Chemical Geology 28, 199260.CrossRefGoogle Scholar
Clémence, ME, Gardin, S and Bartolini, A (2015) New insights in the pattern and timing of the Early Jurassic calcareous nannofossil crisis. Palaeogeography, Palaeoclimatology, Palaeoecology 427, 100108.CrossRefGoogle Scholar
Cochlan, WP and Harrison, PJ (1991) Kinetics of nitrogen (nitrate, ammonium and urea) uptake by the picoflagellate Micromonas pusilla (Prasinophyceae). Journal of Experimental Marine Biology and Ecology 153, 129141.CrossRefGoogle Scholar
Cohen, AS (2004) The rhenium–osmium isotope system: applications to geochronological and palaeoenvironmental problems. Journal of the Geological Society 161, 729734.CrossRefGoogle Scholar
Cohen, AS and Coe, AL (2007) The impact of the Central Atlantic Magmatic Province on climate and on the Sr- and Os-isotope evolution of seawater. Palaeogeography, Palaeoclimatology, Palaeoecology 244, 374390.CrossRefGoogle Scholar
Cohen, AS, Coe, AL, Bartlett, JM and Hawkesworth, CJ (1999) Precise Re–Os ages of organic-rich mudrocks and the Os isotope composition of Jurassic seawater. Earth and Planetary Science Letters 167, 159173.CrossRefGoogle Scholar
Cohen, AS, Coe, AL, Harding, SM and Schwark, L (2004) Osmium isotope evidence for the regulation of atmospheric CO2 by continental weathering. Geology 32, 157160.CrossRefGoogle Scholar
Cohen, AS, Coe, AL and Kemp, DB (2007) The Late Palaeocene–Early Eocene and Toarcian (Early Jurassic) carbon isotope excursions: a comparison of their time scales, associated environmental changes, causes and consequences. Journal of the Geological Society 164, 10931108.CrossRefGoogle Scholar
Cole, DB, Planavsky, NJ, Longley, M, Böning, P, Wilkes, D, Wang, X, Swanner, ED, Wittkop, C, Loydell, DK, Busigny, V, Knudsen, AC and Sperling, EA (2020) Uranium isotope fractionation in non-sulfidic anoxic settings and the global uranium isotope mass balance. Global Biogeochemical Cycles 34, e2020GB006649.CrossRefGoogle Scholar
Coleman, ML and Raiswell, R (1981) Carbon, oxygen and sulphur isotope variations in concretions from the Upper Lias of N.E. England. Geochemica et Cosmochemica Acta 45, 329340.CrossRefGoogle Scholar
Colodner, DC, Boyle, EA and Edmond, JM (1993a) Determination of rhenium and platinum in natural waters and sediments, and iridium in sediments by flow injection isotope dilution inductively coupled plasma mass spectrometry. Analytical Chemistry 65, 14191425.CrossRefGoogle Scholar
Colodner, DC, Sachs, J, Ravizza, G, Turekian, K, Edmond, J and Boyle, E (1993b) The geochemical cycle of rhenium: a reconnaissance. Earth and Planetary Science Letters 117, 205221.CrossRefGoogle Scholar
Comas-Rengifo, MJ, Arias, C, Gómez, JJ, Goy, A, Herrero, C, Osete, ML and Palencia, A (2010) A complementary section for the proposed Toarcian (Lower Jurassic) global stratotype: the Almonacid de la Cuba section (Spain). Stratigraphy and Geological Correlation 18, 133152.CrossRefGoogle Scholar
Condie, KC (1993) Chemical composition and evolution of the upper continental crust: contrasting results from surface samples and shales. Chemical Geology 104, 137.CrossRefGoogle Scholar
Cope, JCW, Getty, TA, Howarth, MK, Morton, N and Torrens, HS (1980) A Correlation of Jurassic Rock in the British Isles Part One: Introduction and Lower Jurassic. Geological Society London, Special Report 14, Oxford: Blackwell, 73 pp.Google Scholar
Cox, BM, Sumbler, MG and Ivimey-Cook, HC (1999 ) A formational framework for the Lower Jurassic of England and Wales (onshore area). British Geological Survey Research Report RR/99/01, Keyworth, 28 pp.Google Scholar
Cramer, BS and Jarvis, I (2020) Carbon isotope stratigraphy. In The Geologic Time Scale 2020 (eds Gradstein, F, Ogg, JG and Ogg, G), 1, pp. 309343. Amsterdam: Elsevier.CrossRefGoogle Scholar
Crusius, J, Calvert, SE, Pedersen, TF and Sage, D (1996) Rhenium and molybdenum enrichments in sediments as indicators of oxic, suboxic and sulfidic conditions of deposition. Earth and Planetary Science Letters 145, 6578.CrossRefGoogle Scholar
Cumberland, SA, Douglas, G, Grice, K and Moreau, JW (2016) Uranium mobility in organic matter-rich sediments: a review of geological and geochemical processes. Earth-Science Reviews 159, 160185.CrossRefGoogle Scholar
da Rocha, RB, Mattioli, E, Duarte, LV, Pittet, B, Elmi, S, Mouterde, R, Cabral, MC, Comas-Rengifo, MJ, Gomez, JJ, Goy, A, Hesselbo, SP, Jenkyns, HC, Littler, K, Mailliot, S, de Oliveira, LCV, Osete, ML, Perilli, N, Pinto, S, Ruget, C and Suan, G (2016) Base of the Toarcian Stage of the Lower Jurassic defined by the Global Boundary Stratotype Section and Point (GSSP) at the Peniche section (Portugal). Episodes 39, 460481.CrossRefGoogle Scholar
Danise, S, Clemence, ME, Price, GD, Murphy, DP, Gomez, JJ and Twitchett, RJ (2019) Stratigraphic and environmental control on marine benthic community change through the early Toarcian extinction event (Iberian Range, Spain). Palaeogeography, Palaeoclimatology, Palaeoecology 524, 183200.CrossRefGoogle Scholar
Danise, S, Slater, SM, Vajda, V and Twitchett, RJ (2022) Land-sea ecological connectivity during a Jurassic warming event. Earth and Planetary Science Letters 578, 117290.CrossRefGoogle Scholar
Danise, S, Twitchett, RJ and Little, CTS (2015) Environmental controls on Jurassic marine ecosystems during global warming. Geology 43, 263266.CrossRefGoogle Scholar
Danise, S, Twitchett, RJ, Little, CTS and Clémence, M-E (2013) The impact of global warming and anoxia on marine benthic community dynamics: an example from the Toarcian (Early Jurassic). Plos One 8, 114.CrossRefGoogle ScholarPubMed
De Baets, K, Nätscher, PS, Rita, P, Fara, E, Neige, P, Bardin, J, Dera, G, Duarte, LV, Hughes, Z, Laschinger, P, Garcia-Ramos, JC, Piñuela, L, Übelacker, C and Weis, R (2021) The impact of the Pliensbachian–Toarcian crisis on belemnite assemblages and size distribution. Swiss Journal of Palaeontology 140, 114.CrossRefGoogle Scholar
de Graciansky, P-C, Dardeau, G, Dommergues, JL, Durlet, C, Marchand, D, Dumont, T, Hesselbo, SP, Jacquin, T, Goggin, V, Meister, C, Mouterde, R, Rey, J and Vail, PR (1998) Ammonite biostratigraphic correlation and Early Jurassic sequence stratigraphy in France: Comparisons with some U.K. sections. In Mesozoic and Cenozoic Sequence Stratigraphy of European Basins (eds de Graciansky, P-C, Hardenbol, J, Jacquin, T and Vail, PR). SEPM, Special Publication 60, 583622.CrossRefGoogle Scholar
de Lange, GJ, Jarvis, I and Kuijpers, A (1987) Geochemical characteristics and provenance of late Quaternary sediments from the Madeira Abyssal Plain, N. Atlantic. In Geology and Geochemistry of Abyssal Plains (eds PPE Weaver and J Thomson). Geological Society London, Special Publications 31, 147165.CrossRefGoogle Scholar
De Lena, LF, Taylor, D, Guex, J, Bartolini, A, Adatte, T, van Acken, D, Spangenberg, JE, Samankassou, E, Vennemann, T and Schaltegger, U (2019) The driving mechanisms of the carbon cycle perturbations in the late Pliensbachian (Early Jurassic). Scientific Reports 9, 18430.CrossRefGoogle ScholarPubMed
de Vos, R (2017) Textural and mineralogical characterization of the late Pliensbachian–early Toarcian sediments of the Cleveland Basin, Yorkshire, U.K. MSc thesis, Utrecht University, Utrecht, 86 pp. Published thesis https://studenttheses.uu.nl/handle/20.500.12932/28170.Google Scholar
Dean, WE, Arthur, MA and Stow, DAV (1984) Origin and geochemistry of Cretaceous deep-sea black shales and multicolored claystones, with emphasis on Deep–Sea Drilling Project Site 530, southern Angola Basin. Initial Reports of the Deep Sea Drilling Project 75, 819844.Google Scholar
Deconinck, JF, Hesselbo, SP and Pellenard, P (2019) Climatic and sea-level control of Jurassic (Pliensbachian) clay mineral sedimentation in the Cardigan Bay Basin, Llanbedr (Mochras Farm) borehole, Wales. Sedimentology 66, 27692783.CrossRefGoogle Scholar
Deegan, FM, Bédard, JH, Grasby, SE, Dewing, K, Geiger, H, Misiti, V, Capriola, M, Callegaro, S, Svensen, HH, Yakymchuk, C, Aradi, LE, Feda, C and Troll, VR (2022) Magma–shale interaction in large igneous provinces: implications for climate warming and sulfide genesis. Journal of Petrology 63, 110.CrossRefGoogle Scholar
Dera, G, Brigaud, B, Monna, F, Laffont, R, Puceat, E, Deconinck, JF, Pellenard, P, Joachimski, MM and Durlet, C (2011) Climatic ups and downs in a disturbed Jurassic world. Geology 39, 215218.CrossRefGoogle Scholar
Dera, G and Donnadieu, Y (2012) Modeling evidences for global warming, Arctic seawater freshening, and sluggish oceanic circulation during the Early Toarcian anoxic event. Paleoceanography 27, PA2211.CrossRefGoogle Scholar
Dera, G, Neige, P, Dommergues, J-L, Fara, E, Laffont, R and Pellenard, P (2010) High-resolution dynamics of Early Jurassic marine extinctions: the case of Pliensbachian–Toarcian ammonites (Cephalopoda). Journal of the Geological Society 167, 2133.CrossRefGoogle Scholar
Dera, G, Pellenard, P, Neige, P, Deconinck, J-F, Pucéat, E and Dommergues, J-L (2009) Distribution of clay minerals in Early Jurassic Peritethyan seas: palaeoclimatic significance inferred from multiproxy comparisons. Palaeogeography, Palaeoclimatology, Palaeoecology 271, 3951.CrossRefGoogle Scholar
Dickens, GR, Oneil, JR, Rea, DK and Owen, RM (1995) Dissociation of oceanic methane hydrate as a cause of the carbon-isotope excursion at the end of the Paleocene. Paleoceanography 10, 965971.CrossRefGoogle Scholar
Dickson, AJ (2017) A molybdenum-isotope perspective on Phanerozoic deoxygenation events. Nature Geoscience 10, 721726.CrossRefGoogle Scholar
Dickson, AJ, Davies, M, Bagard, M-L and Cohen, AS (2022a) Quantifying seawater exchange rates in the Eocene Arctic Basin using osmium isotopes. Geochemical Perspective Letters 24, 711.CrossRefGoogle Scholar
Dickson, AJ, Gill, BC, Ruhl, M, Jenkyns, HC, Porcelli, D, Idiz, E, Lyons, TW and van den Boorn, SHJM (2017) Molybdenum-isotope chemostratigraphy and paleoceanography of the Toarcian Oceanic Anoxic Event (Early Jurassic). Paleoceanography 32, 813829.CrossRefGoogle Scholar
Dickson, AJ, Hsieh, Y-T and Bryan, A (2020) The rhenium isotope composition of Atlantic Ocean seawater. Geochimica et Cosmochimica Acta 287, 221228.CrossRefGoogle Scholar
Dickson, AJ, Idiz, E, Porcelli, D, Murphy, MJ, Celestino, R, Jenkyns, HC, Poulton, SW, Hesselbo, SP, Hooker, JN, Ruhl, M and van der Boorn, SHJM (2022b) No effect of thermal maturity on the Mo, U, Cd, and Zn isotope compositions of Lower Jurassic organic-rich sediments. Geology 50, 598602.CrossRefGoogle Scholar
Dinis, P, Garzanti, E, Vermeesch, P and Huvi, J (2017) Climatic zonation and weathering control on sediment composition (Angola). Chemical Geology 467, 110121.CrossRefGoogle Scholar
Donohue, JG, Florio, BJ and Fowler, AC (2023) The development of deep-ocean anoxia in a comprehensive ocean phosphorus model. GEM - International Journal on Geomathematics 14, 12.CrossRefGoogle Scholar
Doyle, P (1990) The British Toarcian (Lower Jurassic) belemnites. Monographs of the Paleontological Society 144, 149.CrossRefGoogle Scholar
Dunham, KC (1961) Black shale, oil and sulphide ore. Advancement of Science 18, 284299.Google Scholar
Emerson, SR and Huested, SS (1991) Ocean anoxia and the concentration of molybdenum and vanadium in seawater. Marine Chemistry 34, 177196.CrossRefGoogle Scholar
Endrizzi, F and Rao, L (2014) Chemical speciation of uranium(VI) in marine environments: complexation of calcium and magnesium ions with [(UO2)(CO3)3]4 and the effect on the extraction of uranium from seawater. Chemistry A European Journal 20, 1449914506.CrossRefGoogle ScholarPubMed
Erba, E (2004) Calcareous nannofossils and Mesozoic oceanic anoxic events. Marine Micropaleontology 52, 85106.CrossRefGoogle Scholar
Erba, E, Bottini, C, Faucher, G, Gambacorta, G and Visentin, S (2019) The response of calcareous nannoplankton to Oceanic Anoxic Events: the Italian pelagic record. Bollettino della Società Paleontologica Italiana 58, 5171.Google Scholar
Erba, E, Cavalheiro, L, Dickson, AJ, Faucher, G, Gambacorta, G, Jenkyns, HC and Wagner, T (2022) Carbon- and oxygen-isotope signature of the Toarcian Oceanic Anoxic Event: insights from two Tethyan pelagic sequences (Gajum and Sogno Cores – Lombardy Basin, northern Italy). Newsletters on Stratigraphy 55, 451477.CrossRefGoogle Scholar
Ernst, TW (1970) Geochemical Facies Analysis. Amsterdam: Elsevier, 152 pp.Google Scholar
Espitalié, J, Deroo, G and Marquis, F (1985) La pyrolyse Rock-Eval et ses applications. Deuxième partie. Revue de l’Institut français du Pétrole 40, 755784.CrossRefGoogle Scholar
Ettinger, NP, Larson, TE, Kerans, C, Thibodeau, AM, Hattori, KE, Kacur, SM and Martindale, RC (2021) Ocean acidification and photic-zone anoxia at the Toarcian Oceanic Anoxic Event: insights from the Adriatic Carbonate Platform. Sedimentology 68, 63107.CrossRefGoogle Scholar
Fagel, N (2007) Clay minerals, deep circulation and climate. In Proxies in Late Cenozoic Paleoceanography (eds Hillaire-Marcel, C and De Vernal, A), pp. 139184. Developments in Marine Geology 1, Amsterdam: Elsevier.CrossRefGoogle Scholar
Fan, H, Nielsen, SG, Owens, JD, Auro, M, Shu, SA, Hardisty, DS, Horner, TJ, Bowman, CN, Young, SA and Wen, H (2020) Constraining oceanic oxygenation during the Shuram excursion in South China using thallium isotopes. Geobiology 18, 348365.CrossRefGoogle ScholarPubMed
Fantasia, A, Adatte, T, Spangenberg, JE, Font, E, Duarte, LV and Föllmi, KB (2019) Global versus local processes during the Pliensbachian–Toarcian transition at the Peniche GSSP, Portugal: a multi-proxy record. Earth-Science Reviews 198, 116.CrossRefGoogle Scholar
Fantasia, A, Föllmi, KB, Adatte, T, Bernárdez, E, Spangenberg, JE and Mattioli, E (2018a) The Toarcian Oceanic Anoxic Event in southwestern Gondwana: an example from the Andean Basin, northern Chile. Journal of the Geological Society 175, 883902.CrossRefGoogle Scholar
Fantasia, A, Föllmi, KB, Adatte, T, Spangenberg, JE and Montero-Serrano, J-C (2018b) The Early Toarcian oceanic anoxic event: paleoenvironmental and paleoclimatic change across the Alpine Tethys (Switzerland). Global and Planetary Change 162, 5368.CrossRefGoogle Scholar
Faucher, G, Visentin, S, Gambacorta, G and Erba, E (2022) Schizosphaerella size and abundance variations across the Toarcian Oceanic Anoxic Event in the Sogno Core (Lombardy Basin, Southern Alps). Palaeogeography, Palaeoclimatology, Palaeoecology 595, 115.CrossRefGoogle Scholar
Fendley, IM, Frieling, J, Mather, TA, Ruhl, M, Hesselbo, SP and Jenkyns, HC (2024) Early Jurassic large igneous province carbon emissions constrained by sedimentary mercury. Nature Geoscience, https://doi.org/10.1038/s41561-024-01378-5.CrossRefGoogle Scholar
Fernández-Martínez, J, Martínez Ruiz, F, Rodríquez-Tovar, FJ, Piñuela, L, García-Ramos, JC and Algeo, TJ (2023) Euxinia and hydrographic restriction in the Tethys Ocean: reassessing global oceanic anoxia during the early Toarcian. Global and Planetary Change 221, 113.CrossRefGoogle Scholar
Ferrari, M, Little, CTS and Atkinson, JW (2021) Upper Toarcian (Lower Jurassic) marine gastropods from the Cleveland Basin, England: systematics, palaeobiogeography and contribution to biotic recovery from the early Toarcian extinction event. Papers in Palaeontology 7, 885912.CrossRefGoogle Scholar
Filatova, NI, Konstantinovskaya, E and Vishnevskaya, V (2022) Jurassic-Lower Cretaceous siliceous rocks and black shales from allochthonous complexes of the Koryak–Western Kamchatka orogenic belt, East Asia. International Geology Review 64, 311330.CrossRefGoogle Scholar
Fleischmann, S, Picotti, V, Caves Rugenstein, JK, Cobianchi, M and Bernasconi, SM (2022) Effects of the Pliensbachian–Toarcian Boundary Event on carbonate productivity of a Tethyan platform and slope. Paleoceanography and Paleoclimatology 37, e2021PA004392.CrossRefGoogle Scholar
Föllmi, KB, Badertscher, C, de Kaenel, E, Stille, P, John, CM, Adatte, T and Steinmann, P (2005) Phosphogenesis and organic-carbon preservation in the Miocene Monterey Formation at Naples beach, California – The Monterey Hypothesis revisited. Geological Society of America Bulletin 117, 589619.CrossRefGoogle Scholar
Foster, GL, Hull, P, Lunt, D and Zachos, J (2018) Placing our current ‘hyperthermal’ in the context of rapid climate change in our geological past. Philosophical Transactions of the Royal Society A 375, 20170086.CrossRefGoogle Scholar
Fox-Strangways, C (1892) The Jurassic Rocks of Britain. Vol. 1. Yorkshire. Memoirs of the Geological Survey of the United Kingdom. London: HMSO, 551 pp.Google Scholar
Fraguas, A, Gómez, JJ, Goy, A and Comas-Rengifo, MJ (2021) The response of calcareous nannoplankton to the latest Pliensbachian–early Toarcian environmental changes in the Camino Section (Basque Cantabrian Basin, northern Spain). In Carbon Cycle and Ecosystem Response to the Jenkyns Event in the Early Toarcian (Jurassic) (eds Reolid, M, Duarte, LV, Mattioli, E and Ruebsam, W), pp. 3158. Geological Society London, Special Publications 514.Google Scholar
French, KL, Sepulveda, J, Trabucho-Alexandre, J, Gröcke, DR and Summons, RE (2014) Organic geochemistry of the early Toarcian oceanic anoxic event in Hawsker Bottoms, Yorkshire, England. Earth and Planetary Science Letters 390, 116127.CrossRefGoogle Scholar
Fu, X, Wang, J, Wen, H, Song, C, Wang, Z, Zeng, S, Feng, X and Wei, H (2021) A Toarcian Ocean Anoxic Event record from an open-ocean setting in the eastern Tethys: implications for global climatic change and regional environmental perturbation. Science China Earth Sciences 64, 18601872.CrossRefGoogle Scholar
Fu, X, Wang, J, Zeng, S, Feng, X, Wang, D and Song, C (2017) Continental weathering and palaeoclimatic changes through the onset of the Early Toarcian oceanic anoxic event in the Qiangtang Basin, eastern Tethys. Palaeogeography, Palaeoclimatology, Palaeoecology 487, 241250.CrossRefGoogle Scholar
Gad, MA (1966) A geochemical study of the Liassic rocks of the Yorkshire coast. PhD thesis, University of London, London, 269 pp. Published thesis https://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.761665.Google Scholar
Gad, MA, Catt, JA and Le Riche, HH (1969) Geochemistry of the Whitbian (Upper Lias) sediments of the Yorkshire coast. Proceedings of the Yorkshire Geological Society 37, 105136.CrossRefGoogle Scholar
Galasso, F, Feist-Burkhardt, S and Schneebeli-Hermann, E (2022) The palynology of the Toarcian Oceanic Anoxic Event at Dormettingen, southwest Germany, with emphasis on changes in vegetational dynamics. Review of Palaeobotany and Palynology 304, 133.CrossRefGoogle Scholar
Gambacorta, G, Brumsack, H-J, Jenkyns, HC and Erba, E (2024) The early Toarcian Oceanic Anoxic Event (Jenkyns Event) in the Alpine-Mediterranean Tethys, north African margin, and north European epicontinental seaway. Earth-Science Reviews 248, 104636.CrossRefGoogle Scholar
Gambacorta, G, Cavalheiro, L, Brumsack, H-J, Dickson, AJ, Jenkyns, HC, Schnetger, B, Wagner, T and Erba, E (2023) Suboxic conditions prevailed during the Toarcian Oceanic Anoxic Event in the Alpine-Mediterranean Tethys: the Sogno Core pelagic record (Lombardy Basin, northern Italy). Global and Planetary Change 223, 104089.CrossRefGoogle Scholar
Gangl, SK, Stirling, CH, Jenkyns, HC, Preston, WJ, Clarkson, MO, Moy, CM, Dickson, AJ and Porcelli, D (2023) Regional conditions cause contrasting behaviour in U-isotope fractionation in black shales: constraints for global ocean palaeo-redox reconstructions. Chemical Geology 623, 121411.CrossRefGoogle Scholar
García Joral, F, Gómez, JJ and Goy, A (2011) Mass extinction and recovery of the Early Toarcian (Early Jurassic) brachiopods linked to climate change in Northern and Central Spain. Palaeogeography, Palaeoclimatology, Palaeoecology 302, 367380.CrossRefGoogle Scholar
Garzanti, E, Padoan, M, Setti, M, López-Galindo, A and Villa, IM (2014) Provenance versus weathering control on the composition of tropical river mud (southern Africa). Chemical Geology 366, 6174.CrossRefGoogle Scholar
Gaynor, SP, Svensen, HH, Polteau, S and Schaltegger, U (2022) Local melt contamination and global climate impact: dating the emplacement of Karoo LIP sills into organic–rich shale. Earth and Planetary Science Letters 579, 110.CrossRefGoogle Scholar
Ghadeer, S (2011) An investigation of the sediment dispersal operating to control lithofacies variability and organic carbon preservation in an ancient mud-dominated succession: a case study of the Lower Jurassic mudstone dominated succession exposed in the Cleveland Basin (North Yorkshire). PhD thesis, University of Manchester, Manchester, 196 pp. Published thesis https://pure.manchester.ac.uk/ws/portalfiles/portal/54510396/full_text.pdf.Google Scholar
Ghadeer, SG and Macquaker, JHS (2011) Sediment transport processes in an ancient mud-dominated succession: a comparison of processes operating in marine offshore settings and anoxic basinal environments. Journal of the Geological Society 168, 11211132.CrossRefGoogle Scholar
Gibbs, RJ (1977) Clay mineral segregation in the marine environment. Journal of Sedimentary Petrology 47, 237242.Google Scholar
Gill, BC, Lyons, TW and Jenkyns, HC (2011) A global perturbation to the sulfur cycle during the Toarcian Oceanic Anoxic Event. Earth and Planetary Science Letters 312, 484496.CrossRefGoogle Scholar
Godet, A, Bodin, S, Adatte, T and Föllmi, KB (2008) Platform-induced clay-mineral fractionation along a northern Tethyan basin-platform transect: implications for the interpretation of Early Cretaceous climate change (Late Hauterivian–Early Aptian). Cretaceous Research 29, 830847.CrossRefGoogle Scholar
Gómez, JJ, Comas-Rengifo, MJ and Goy, A (2016) Palaeoclimatic oscillations in the Pliensbachian (Early Jurassic) of the Asturian Basin (Northern Spain). Climate of the Past 12, 11991214.CrossRefGoogle Scholar
Gómez, JJ and Goy, A (2011) Warming-driven mass extinction in the Early Toarcian (Early Jurassic) of northern and central Spain. Correlation with other time-equivalent European sections. Palaeogeography, Palaeoclimatology, Palaeoecology 306, 176195.CrossRefGoogle Scholar
Gómez, JJ, Goy, A and Canales, ML (2008) Seawater temperature and carbon isotope variations in belemnites linked to mass extinction during the Toarcian (Early Jurassic) in central and northern Spain. Comparison with other European sections. Palaeogeography, Palaeoclimatology, Palaeoecology 258, 2858.CrossRefGoogle Scholar
González López, JM, Bauluz, B, Yuste, A, Mayayo, MJ and Fernández-Nieto, C (2005) Mineralogical and trace element composition of clay-sized fractions from Albian siliciclastic rocks (Oliete Basin, NE Spain). Clay Minerals 40, 565580.CrossRefGoogle Scholar
Gradstein, FM, Ogg, JG, Schmitz, MD and Ogg, GM, eds. (2020) Geologic Time Scale 2020, 1357 pp. Amsterdam: Elsevier.Google Scholar
Grasby, SE, Them, TR II, Chen, ZH, Yin, RS and Ardakani, OH (2019) Mercury as a proxy for volcanic emissions in the geologic record. Earth-Science Reviews 196, 102880.CrossRefGoogle Scholar
Greber, ND, Davies, JHFL, Gaynor, SP, Jourdan, F, Bertrand, H and Schaltegger, U (2020) New high precision U–Pb ages and Hf isotope data from the Karoo large igneous province; implications for pulsed magmatism and early Toarcian environmental perturbations. Results in Geochemistry 1, 100005.CrossRefGoogle Scholar
Greensmith, JT, Rawson, PF and Shalaby, SE (1980) An association of minor fining-upward cycles and aligned gutter marks in the Middle Lias (Lower Jurassic) of the Yorkshire coast. Proceedings of the Yorkshire Geological Society 42, 525538.CrossRefGoogle Scholar
Grice, K and Eiserbeck, C (2014) The analysis and application of biomarkers. In Treatise on Geochemistry (2nd Edn) (eds Holland, HD and Turekian, KK), 12, pp. 4778. Amsterdam: Elsevier.CrossRefGoogle Scholar
Gröcke, DR, Hori, RS, Trabucho-Alexandre, J, Kemp, DB and Schwark, L (2011) An open ocean record of the Toarcian oceanic anoxic event. Solid Earth 2, 245257.CrossRefGoogle Scholar
Gröcke, DR, Rimmer, SM, Yoksoulian, LE, Cairncross, B, Tsikos, H and van Hunen, J (2009) No evidence for thermogenic methane release in coal from the Karoo–Ferrar large igneous province. Earth and Planetary Science Letters 277, 204212.CrossRefGoogle Scholar
Guex, J, Pilet, S, Müntener, O, Bartolini, A, Spangenberg, J, Schoene, B, Sell, B and Schaltegger, U (2016) Thermal erosion of cratonic lithosphere as a potential trigger for mass-extinction. Scientific Reports 6, 23168.CrossRefGoogle ScholarPubMed
Hall, A and Bishop, P (2002) Scotland’s denudational history: An integrated view of erosion and sedimentation at an uplifted passive margin. In Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration (eds Doré, AG, Cartwright, JA, Stoker, MS, Turner, JP and White, N), pp. 271290. Geological Society London, Special Publications 196.Google Scholar
Hallam, A (1962) A band of extraordinary calcareous concretions in the Upper Lias of Yorkshire, England. Journal of Sedimentary Petrology 32, 840847.CrossRefGoogle Scholar
Hallam, A (1986) The Pliensbachian and Tithonian extinction events. Nature 319, 765768.CrossRefGoogle Scholar
Hallam, A (1997) Estimates of the amount and rate of sea-level change across the Rhaetian-Hettangian and Pliensbachian–Toarcian boundaries (latest Triassic to early Jurassic). Journal of the Geological Society 154, 773779.CrossRefGoogle Scholar
Hallam, A (2001) A review of the broad pattern of Jurassic sea-level changes and their possible causes in the light of current knowledge. Palaeogeography, Palaeoclimatology, Palaeoecology 167, 2337.CrossRefGoogle Scholar
Hallimond, AF, Ennos, FR and Sutcliffe, R (1925) Iron Ores: Bedded Ores of England and Wales. Petrography and Chemistry. Memoirs of the Geological Survey Special Reports on the Mineral Resources of Great Britain 29, London: HMSO, 139 pp.Google Scholar
Hammer, Ø, Harper, DAT and Ryan, PD (2001) PAST: Paleontological Statistics Software Package for Education and Data Analysis. Palaeontologia Electronica 4, 19.Google Scholar
Han, Z, Hu, X, He, T, Newton, RJ, Jenkyns, HC, Jamieson, RA and Franceschi, M (2022) Early Jurassic long-term oceanic sulfur-cycle perturbations in the Tibetan Himalaya. Earth and Planetary Science Letter 578, 113.CrossRefGoogle Scholar
Han, Z, Hu, X, Newton, RJ, He, T, Mills, BJW, Jenkyns, HC, Ruhl, M and Jamieson, RA (2023) Spatially heterogenous seawater δ34S and global cessation of Ca-sulfate burial during the Toarcian oceanic anoxic event. Earth and Planetary Science Letters 622, 118404.CrossRefGoogle Scholar
Haq, BU (2018) Jurassic sea-level variations: a reappraisal. GSA Today 28, 410.CrossRefGoogle Scholar
Hardenbol, J, Thierry, J, Farley, MB, Jacquin, T, Graciansky, P-Cd and Vail, PR (1998) Mesozoic and Cenozoic sequence chronostratigraphic framework in European basins. In Mesozoic and Cenozoic Sequence Stratigraphy of European Basins (eds Graciansky, P-Cd, Hardenbol, J, Jacquin, T and Vail, PR), pp. 313. SEPM, Special Publication 60.Google Scholar
Hardisty, DS, Lyons, TW, Riedinger, N, Isson, TT, Owens, JD, Aller, RC, Rye, DM, Planavsky, NJ, Reinhard, CT, Gill, BC, Masterson, AL, Asael, D and Johnston, DT (2018) An evaluation of sedimentary molybdenum and iron as proxies for pore fluid paleoredox conditions. American Journal of Science 318, 527556.CrossRefGoogle Scholar
Harries, PJ and Little, CTS (1999) The early Toarcian (Early Jurassic) and the Cenomanian–Turonian (Late Cretaceous) mass extinctions: similarities and contrasts. Palaeogeography, Palaeoclimatology, Palaeoecology 154, 3966.CrossRefGoogle Scholar
Hartnett, HE, Keil, RG, Hedges, JI and Devol, AH (1998) Influence of oxygen exposure time on organic carbon preservation in continental margin sediments. Nature 391, 572574.CrossRefGoogle Scholar
Hawco, NJ, Lam, PJ, Lee, J-M, Ohnemus, DC, Noble, AE, Wyatt, NJ, Lohan, MC and Saito, MA (2018) Cobalt scavenging in the mesopelagic ocean and its influence on global mass balance: synthesizing water column and sedimentary fluxes. Marine Chemistry 201, 151166.CrossRefGoogle Scholar
Hazen, RM, Ewing, RC and Sverjensky, DA (2009) Evolution of uranium and thorium minerals. American Mineralogist 94, 12931311.CrossRefGoogle Scholar
He, ZW, Clarkson, MO, Andersen, M, Archer, C, Sweere, TC, Kraal, P, Guthauser, A, Huang, F and Vance, D (2021) Temporally and spatially dynamic redox conditions on an upwelling margin: the impact on coupled sedimentary Mo and U isotope systematics, and implications for the Mo–U paleoredox proxy. Geochimica et Cosmochimica Acta 309, 251271.CrossRefGoogle Scholar
Head, MJH, Aubry, M-P, Piller, WE and Walker, M (2023) Standard Auxiliary Boundary Stratotype (SABS) approved to support the Global boundary Stratotype Section and Point (GSSP). Episodes 46, 99100.CrossRefGoogle Scholar
Hecky, RE, Campbell, M and Hendzel, LL (1993) The stoichiometry of carbon, nitrogen, and phosphorus in particulate matter of lakes and oceans. Limnology and Oceanography 38, 709724.CrossRefGoogle Scholar
Heimdal, TH, Goddéris, Y, Jones, MT and Svensen, HH (2021) Assessing the importance of thermogenic degassing from the Karoo Large Igneous Province (LIP) in driving Toarcian carbon cycle perturbations. Nature Communications 12, 17.CrossRefGoogle ScholarPubMed
Helly, JJ and Levin, LA (2004) Global distribution of naturally occurring marine hypoxia on continental margins. Deep-Sea Research I 51, 11591168.CrossRefGoogle Scholar
Helz, GR and Adelson, JM (2013) Trace element profiles in sediments as proxies of dead zone history; rhenium compared to molybdenum. Environmental Science and Technology 47, 12571264.CrossRefGoogle ScholarPubMed
Hemingway, JE (1933 ) Whitby jet and its relation to Upper Lias sedimentation in the Yorkshire basin. PhD thesis, University of Leeds, Leeds, 119 pp. Published thesis https://etheses.whiterose.ac.uk/2312/1/uk_bl_ethos_539893.pdf.Google Scholar
Hemingway, JE (1974) Jurassic. In The Geology and Mineral Resources of Yorkshire (eds Rayner, DH and Hemingway, JE). pp. 161223. Leeds: Maney and Son. Google Scholar
Hermoso, M, Le Callonnec, L, Minoletti, F, Renard, M and Hesselbo, SP (2009a) Expression of the Early Toarcian negative carbon-isotope excursion in separated carbonate microfractions (Jurassic, Paris Basin). Earth and Planetary Science Letters 277, 194203.CrossRefGoogle Scholar
Hermoso, M, Minoletti, F, Le Callonnec, L, Jenkyns, HC, Hesselbo, SP, Rickaby, REM, Renard, M, de Rafelis, M and Emmanuel, L (2009b) Global and local forcing of Early Toarcian seawater chemistry: a comparative study of different paleoceanographic settings (Paris and Lusitanian basins). Paleoceanography 24, PA4208.CrossRefGoogle Scholar
Hermoso, M, Minoletti, F and Pellenard, P (2013) Black shale deposition during Toarcian super-greenhouse driven by sea level. Climate of the Past 9, 27032712.CrossRefGoogle Scholar
Hermoso, M, Minoletti, F, Rickaby, REM, Hesselbo, SP, Baudin, F and Jenkyns, HC (2012) Dynamics of a stepped carbon-isotope excursion: ultra high-resolution study of Early Toarcian environmental change. Earth and Planetary Science Letters 319, 4554.CrossRefGoogle Scholar
Hermoso, M and Pellenard, P (2014) Continental weathering and climatic changes inferred from clay mineralogy and paired carbon isotopes across the early to middle Toarcian in the Paris Basin. Palaeogeography, Palaeoclimatology, Palaeoecology 399, 385393.CrossRefGoogle Scholar
Hesselbo, SP (2008) Sequence stratigraphy and inferred relative sea-level change from the onshore British Jurassic. Proceedings of the Geologists’ Association 119, 1934.CrossRefGoogle Scholar
Hesselbo, SP, Gröcke, DR, Jenkyns, HC, Bjerrum, CJ, Farrimond, P, Bell, HSM and Green, OR (2000) Massive dissociation of gas hydrate during a Jurassic oceanic anoxic event. Nature 406, 392395.CrossRefGoogle ScholarPubMed
Hesselbo, SP, Jenkyns, HC, Duarte, LV and Oliveira, LCV (2007) Carbon-isotope record of the Early Jurassic (Toarcian) Oceanic Anoxic Event from fossil wood and marine carbonate (Lusitanian Basin, Portugal). Earth and Planetary Science Letters 253, 455470.CrossRefGoogle Scholar
Hesselbo, SP and Jenkyns, HC (1995) A comparison of the Hettangian to Bajocian successions of Dorset and Yorkshire. In Field Geology of the British Jurassic (ed Taylor, PD). pp. 105150. London: The Geological Society.Google Scholar
Hesselbo, SP and King, C (2019) Stratigraphic framework for the Yorkshire Lias. In Fossils from the Lias of the Yorkshire Coast (ed Lord, AR). Field Guide to Fossils. pp. 3040. Palaeontological Association.Google Scholar
Hesselbo, SP, Little, CTS, Ruhl, M, Thibault, N and Ullmann, CV (2020a) Comments on “Paleosalinity determination in ancient epicontinental seas: a case study of the T-OAE in the Cleveland Basin (UK)” by Remirez, M.N. and Algeo, T.J. Earth-Science Reviews 208, 103290.CrossRefGoogle Scholar
Hesselbo, SP, Ogg, JG, Ruhl, M, Hinnov, LA and Huang, CJ (2020b) The Jurassic Period. In Geologic Time Scale 2020 (eds Gradstein, FM, Ogg, JG, Schmitz, MD and Ogg, GM). 2, pp. 9551021. Amsterdam: Elsevier.CrossRefGoogle Scholar
Hesselbo, SP and Pieńkowski, G (2011) Stepwise atmospheric carbon-isotope excursion during the Toarcian Oceanic Anoxic Event (Early Jurassic, Polish Basin). Earth and Planetary Science Letters 301, 365372.CrossRefGoogle Scholar
Higgins, MB, Robinson, RS, Husson, JM, Carter, SJ and Pearson, A (2012) Dominant eukaryotic export production during ocean anoxic events reflects the importance of recycled NH4+ . Proceeedings of the Natiional Academy of Sciences 109, 22692274.CrossRefGoogle ScholarPubMed
Hoffmann, R and Stevens, K (2020) The palaeobiology of belemnites – foundation for the interpretation of rostrum geochemistry. Biological Reviews 95, 94123.CrossRefGoogle ScholarPubMed
Hollaar, TP, Hesselbo, SP, Deconinck, J-F, Damaschke, M, Ullmann, CV, Jiang, M and Belcher, CM (2023) Environmental changes during the onset of the Late Pliensbachian Event (Early Jurassic) in the Cardigan Bay Basin, Wales. Climate of the Past 19, 979997.CrossRefGoogle Scholar
Hönisch, B, Ridgwell, A, Schmidt, DN, Thomas, E, Gibbs, SJ, Sluijs, A, Zeebe, R, Kump, L, Martindale, RC, Greene, SE, Kiessling, W, Ries, J, Zachos, JC, Royer, DL, Barker, S, Marchitto, TM, Moyer, R, Pelejero, C, Ziveri, P, Foster, GL and Williams, B (2012) The geological record of ocean acidification. Science 335, 10581063.CrossRefGoogle ScholarPubMed
Houben, AJP, Goldberg, T and Slomp, CP (2021) Biogeochemical evolution and organic carbon deposition on the Northwestern European Shelf during the Toarcian Ocean Anoxic Event. Palaeogeography, Palaeoclimatology, Palaeoecology 565, 113.CrossRefGoogle Scholar
Houben, ME, Barnhoorn, A, Lie-A-Fat, J, Ravestein, T, Peach, CJ and Drury, MR (2016a) Microstructural characteristics of the Whitby Mudstone Formation (UK). Marine and Petroleum Geology 70, 185200.CrossRefGoogle Scholar
Houben, ME, Barnhoorn, A, Wasch, L, Trabucho-Alexandre, J, Peach, CJ and Drury, MR (2016b) Microstructures of Early Jurassic (Toarcian) shales of Northern Europe. International Journal of Coal Geology 165, 7689.CrossRefGoogle Scholar
Howard, AS (1985) Lithostratigraphy of the Staithes Sandstone and Cleveland Ironstone formations (Lower Jurassic) of north-east Yorkshire. Proceedings of the Yorkshire Geological Society 45, 261275.CrossRefGoogle Scholar
Howarth, MK (1955) Domerian of the Yorkshire coast. Proceedings of the Yorkshire Geological Society 30, 147175.CrossRefGoogle Scholar
Howarth, MK (1962) The Jet Rock Series and the Alum Shale Series of the Yorkshire coast. Proceedings of the Yorkshire Geological Society 33, 381422.CrossRefGoogle Scholar
Howarth, MK (1973) The stratigraphy and ammonite fauna of the Upper Liassic Grey Shales of the Yorkshire coast. Bulletin of the British Museum (Natural History) Geology 24, 235277.CrossRefGoogle Scholar
Howarth, MK (1992) The ammonite family Hildoceratidae in the Lower Jurassic of Britain. Monograph of the Palaeontographical Society 145–146, 1200.Google Scholar
Huang, CJ and Hesselbo, SP (2014) Pacing of the Toarcian Oceanic Anoxic Event (Early Jurassic) from astronomical correlation of marine sections. Gondwana Research 25, 13481356.CrossRefGoogle Scholar
Huang, Y, Jin, X, Pancost, RD, Kemp, DB and Naafs, BDA (2024) An intensified lacustrine methane cycle during the Toarcian OAE (Jenkyns Event) in the Ordos Basin, northern China. Earth and Planetary Science Letters 639, 118766.CrossRefGoogle Scholar
Hurst, A (1985) The implications of clay mineralogy to palaeoclimate and provenance during the Jurassic in NE Scotland. Scottish Journal of Geology 21, 143160.CrossRefGoogle Scholar
Ikeda, M and Hori, RS (2014) Effects of Karoo–Ferrar volcanism and astronomical cycles on the Toarcian Oceanic Anoxic Events (Early Jurassic). Palaeogeography, Palaeoclimatology, Palaeoecology 410, 134142.CrossRefGoogle Scholar
Ikeda, M, Hori, RS, Ikehara, M, Miyashita, R, Chino, M and Yamada, K (2018) Carbon cycle dynamics linked with Karoo–Ferrar volcanism and astronomical cycles during Pliensbachian–Toarcian (Early Jurassic). Global and Planetary Change 170, 163171.CrossRefGoogle Scholar
Imber, J, Armstrong, H, Clancy, S, Daniels, S, Herringshaw, L, McCaffrey, K, Rodriguez, J, Trabucho-Alexandre, J and Warren, C (2014) Natural fractures in a United Kingdom shale reservoir analog, Cleveland Basin, northeast England. AAPG Bulletin 98, 24112437.CrossRefGoogle Scholar
Immenhauser, A (2009) Estimating palaeo-water depth from the physical rock record. Earth-Science Reviews 96, 107139.CrossRefGoogle Scholar
Ingall, ED, Bustin, RM and Van Cappellen, P (1993) Influence of water column anoxia on the burial and preservation of carbon and phosphorus in marine shales. Geochimica et Cosmochimica Acta 57, 303316.CrossRefGoogle Scholar
Izumi, K, Kemp, DB, Itamiya, S and Inui, M (2018) Sedimentary evidence for enhanced hydrological cycling in response to rapid carbon release during the early Toarcian oceanic anoxic event. Earth and Planetary Science Letters 481, 162170.CrossRefGoogle Scholar
Jarvis, I, Burnett, WC, Nathan, Y, Almbaydin, FSM, Attia, AKM, Castro, LN, Flicoteaux, R, Hilmy, ME, Husain, V, Qutawnah, AA, Serjani, A and Zanin, YN (1994) Phosphorite geochemistry: state-of-the-art and environmental concerns. Eclogae Geologicae Helvetiae 87, 643700.Google Scholar
Jarvis, I (2003) Sample preparation in ICP-MS. In Handbook of Inductively Coupled Plasma Mass Spectrometry (eds Jarvis, KE, Gray, AL and Houk, RS). pp. 172224. Woking: Viridian.Google Scholar
Jeans, CV (2006) Clay mineralogy of the Jurassic strata of the British Isles. Clay Minerals 41, 187307.CrossRefGoogle Scholar
Jenkyns, HC (1980) Cretaceous anoxic events: from continents to oceans. Journal of the Geological Society 137, 171188.CrossRefGoogle Scholar
Jenkyns, HC (1985) The early Toarcian and Cenomanian–Turonian anoxic events in Europe - comparisons and contrasts. Geologische Rundschau 74, 505518.CrossRefGoogle Scholar
Jenkyns, HC (1988) The early Toarcian (Jurassic) anoxic event - stratigraphic, sedimentary, and geochemical evidence. American Journal of Science 288, 101151.CrossRefGoogle Scholar
Jenkyns, HC (1999) Mesozoic anoxic events and palaeoclimate. Zentralblatt für Geologie und Paläontologie. Teil I, Allgemeine, Angewandte, Regionale und Historische Geologie 7–9, 943949.Google Scholar
Jenkyns, HC (2010) Geochemistry of oceanic anoxic events. Geochemistry Geophysics Geosystems 11, Q03004.CrossRefGoogle Scholar
Jenkyns, HC and Clayton, CJ (1997) Lower Jurassic epicontinental carbonates and mudstones from England and Wales: chemostratigraphic signals and the early Toarcian anoxic event. Sedimentology 44, 687706.CrossRefGoogle Scholar
Jenkyns, HC, Dickson, AJ, Ruhl, M and Van den Boorn, SHJM (2017) Basalt–seawater interaction, the Plenus Cold Event, enhanced weathering and geochemical change: deconstructing Oceanic Anoxic Event 2 (Cenomanian–Turonian, Late Cretaceous). Sedimentology 64, 1643.CrossRefGoogle Scholar
Jenkyns, HC, Gröcke, DR and Hesselbo, SP (2001) Nitrogen isotope evidence for water mass denitrification during the early Toarcian (Jurassic) oceanic anoxic event. Paleoceanography 16, 593603.CrossRefGoogle Scholar
Jenkyns, HC, Jones, CE, Gröcke, DR, Hesselbo, SP and Parkinson, DN (2002) Chemostratigraphy of the Jurassic System: applications, limitations and implications for palaeoceanography. Journal of the Geological Society 159, 351378.CrossRefGoogle Scholar
Jenkyns, HC and Macfarlane, S (2021) The chemostratigraphy and environmental significance of the Marlstone and Junction Bed (Beacon Limestone, Toarcian, Lower Jurassic, Dorset, UK). Geological Magazine 159, 357371.CrossRefGoogle Scholar
Jiang, Q, Jourdan, F, Olierook, HKH and Merle, RE (2023) An appraisal of the ages of Phanerozoic large igneous provinces. Earth-Science Reviews 237, 104314.CrossRefGoogle Scholar
Jiang, SY, Song, HJ, Kemp, DB, Dai, X and Liu, XK (2020) Two pulses of extinction of larger benthic foraminifera during the Pliensbachian–Toarcian and early Toarcian environmental crises. Palaeogeography, Palaeoclimatology, Palaeoecology 560, 109998.CrossRefGoogle Scholar
Jin, X, Shi, Z, Baranyi, V, Kemp, DB, Han, Z, Luo, G, Hu, J, He, F, Chen, L and Preto, N (2020) The Jenkyns Event (early Toarcian OAE) in the Ordos Basin, North China. Global and Planetary Change 193, 115.CrossRefGoogle Scholar
Jin, X, Zhang, F, Baranyi, V, Kemp, DB, Feng, X, Grasby, SE, Sun, G, Shi, Z, Chen, W and Dal Corso, J (2022) Early Jurassic massive release of terrestrial mercury linked to floral crisis. Earth and Planetary Science Letters 598, 112.CrossRefGoogle Scholar
Jones, B and Manning, DAC (1994) Comparison of geochemical indexes used for the interpretation of paleoredox conditions in ancient mudstones. Chemical Geology 111, 111129.CrossRefGoogle Scholar
Jones, CE, Jenkyns, HC and Hesselbo, SP (1994) Strontium isotopes in Early Jurassic seawater. Geochimica et Cosmochimica Acta 58, 12851301.CrossRefGoogle Scholar
Junium, CK, Dickson, AJ and Uveges, BT (2018) Perturbation to the nitrogen cycle during rapid Early Eocene global warming. Nature Communications 9, 3186.CrossRefGoogle Scholar
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, 255286.CrossRefGoogle Scholar
Katchinoff, JAR, Syverson, DD, Planavsky, NJ, Evans, ESJ and Rooney, AD (2021) Seawater chemistry and hydrothermal controls on the Cenozoic osmium cycle. Geophysical Research Letters 48, e2021GL095558.CrossRefGoogle Scholar
Kearsley, AT (1989) Iron-rich ooids, their mineralogy and microfabric: clues to their origin and evolution. In Phanerozoic Ironstones (eds Young, TP and Taylor, WEG), pp. 141164. Geological Society London, Special Publications 46.Google Scholar
Kemp, DB, Chen, W, Cho, T, Algeo, TJ, Shen, J and Ikeda, M (2022a) Deep-ocean anoxia across the Pliensbachian–Toarcian boundary and the Toarcian Oceanic Anoxic Event in the Panthalassic Ocean. Global and Planetary Change 212, 114.CrossRefGoogle Scholar
Kemp, DB, Coe, AL, Cohen, AS and Schwark, L (2005) Astronomical pacing of methane release in the Early Jurassic period. Nature 437, 396399.CrossRefGoogle ScholarPubMed
Kemp, DB, Coe, AL, Cohen, AS and Weedon, GP (2011) Astronomical forcing and chronology of the early Toarcian (Early Jurassic) oceanic anoxic event in Yorkshire, UK. Paleoceanography 26, PA4210.CrossRefGoogle Scholar
Kemp, DB, Fraser, WT and Izumi, K (2018) Stratigraphic completeness and resolution in an ancient mudrock succession. Sedimentology 65, 18751890.CrossRefGoogle Scholar
Kemp, DB and Izumi, K (2014) Multiproxy geochemical analysis of a Panthalassic margin record of the early Toarcian oceanic anoxic event (Toyora area, Japan). Palaeogeography, Palaeoclimatology, Palaeoecology 414, 332341.CrossRefGoogle Scholar
Kemp, DB and Sadler, PM (2022) Incompleteness: Dealing with an imperfect stratigraphical record. In Deciphering Earth’s History: the Practice of Stratigraphy (ed Coe, AL), pp. 213226. GSL Geoscience in Practice.Google Scholar
Kemp, DB, Selby, D and Izumi, K (2020) Direct coupling between carbon release and weathering during the Toarcian oceanic anoxic event. Geology 48, 976980.CrossRefGoogle Scholar
Kemp, DB, Suan, G, Fantasia, A, Jin, SM and Chen, WH (2022b) Global organic carbon burial during the Toarcian oceanic anoxic event: patterns and controls. Earth-Science Reviews 231, 104086.CrossRefGoogle Scholar
Kemp, SJ and McKervey, JA (2001) The mineralogy of mudrocks from the Lias Group of England. British Geological Survey Internal Report IR/01/124, Keyworth: British Geological Survey, 38 pp.Google Scholar
Kemp, SJ, Merriman, RJ and Bouch, JE (2005) Clay mineral reaction progress – the maturity and burial history of the Lias Group of England and Wales. Clay Minerals 40, 4361.CrossRefGoogle Scholar
Kendall, B, Andersen, MB and Owens, JD (2021) Assessing the effect of Large Igneous Provinces on global oceanic redox conditions using non-traditional metal isotopes (molybdenum, uranium, thallium). In Large Igneous Provinces: A Driver of Global Environmental and Biotic Changes (eds Ernst, RE, Dickson, AJ and Bekker, A), pp. 305323. Geophysical Monograph 255, American Geophysical Union and John Wiley and Sons, Inc.CrossRefGoogle Scholar
Kennedy, MJ, Pevear, DR and Hill, RJ (2002) Mineral surface control of organic carbon in black shale. Science 295, 657660.CrossRefGoogle ScholarPubMed
Kent, PE (1980) Subsidence and uplift in east Yorkshire and Lincolnshire: a double inversion. Proceedings of the Yorkshire Geological Society 42, 505524.CrossRefGoogle Scholar
Khaustova, N, Tikhomirova, Y, Korost, S, Poludetkina, E, Voropaev, A, Mironenko, M and Spasennykh, M (2021) The study of uranium accumulation in marine bottom sediments: effect of redox conditions at the time of sedimentation. Geosciences 11, 332.CrossRefGoogle Scholar
Klinkhammer, GP and Palmer, MR (1991) Uranium in the oceans: where it goes and why. Geochimica et Cosmochimica Acta 55, 17911806.CrossRefGoogle Scholar
Koopmans, MP, Koster, J, vanKaamPeters, HME, Kenig, F, Schouten, S, Hartgers, WA, deLeeuw, JW and Sissingh Damsté, JS (1996) Diagenetic and catagenetic products of isorenieratene: molecular indicators for photic zone anoxia. Geochimica et Cosmochimica Acta 60, 44674496.CrossRefGoogle Scholar
Korte, C and Hesselbo, SP (2011) Shallow marine carbon and oxygen isotope and elemental records indicate icehouse–greenhouse cycles during the Early Jurassic. Paleoceanography 26, PA4219.CrossRefGoogle Scholar
Korte, C, Hesselbo, SP, Ullmann, CV, Dietl, G, Ruhl, M, Schweigert, G and Thibault, N (2015) Jurassic climate mode governed by ocean gateway. Nature Communications 6, 10015.CrossRefGoogle ScholarPubMed
Kovács, EB, Ruhl, M, Silva, RL, McElwain, JC, Reolid, M, Korte, C, Ruebsam, W and Hesselbo, SP (2024) Mercury sequestration pathways under varying depositional conditions during Early Jurassic (Pliensbachian and Toarcian) Karoo–Ferrar volcanism. Palaeogeography, Palaeoclimatology, Palaeoecology 637, 111977.CrossRefGoogle Scholar
Krencker, F-N, Bodin, S, Suan, G, Heimhofer, U, Kabiri, L and Immenhauser, A (2015) Toarcian extreme warmth led to tropical cyclone intensification. Earth and Planetary Science Letters 425, 120130.CrossRefGoogle Scholar
Krencker, F-N, Fantasia, A, Danisch, J, Martindale, R, Kabiri, L, El Ouali, M and Bodin, S (2020) Two-phased collapse of the shallow-water carbonate factory during the late Pliensbachian–Toarcian driven by changing climate and enhanced continental weathering in the Northwestern Gondwana Margin. Earth-Science Reviews 208, 103254.CrossRefGoogle Scholar
Krencker, F-N, Fantasia, A, El Ouali, M, Kabiri, L and Bodin, S (2022) The effects of strong sediment-supply variability on the sequence stratigraphic architecture: insights from early Toarcian carbonate factory collapses. Marine and Petroleum Geology 136, 105469.CrossRefGoogle Scholar
Krencker, F-N, Lindström, S and Bodin, S (2019) A major sea-level drop briefly precedes the Toarcian oceanic anoxic event: implication for Early Jurassic climate and carbon cycle. Scientific Reports 9, 12518.CrossRefGoogle Scholar
Ku, T-L, Knauss, KG and Mathieu, GG (1977) Uranium in open ocean: concentration and isotopic composition. Deep-Sea Research 24, 10051017.CrossRefGoogle Scholar
Kuenen, PH (1966) Experimental turbidite lamination in a circular flume. Journal of Geology 74, 523545.CrossRefGoogle Scholar
Kulenguski, JT, Gilleaudeau, GJ, Kaufman, AJ, Kipp, MA, Tissot, FLH, Goepfert, TJ, Pitts, AD, Pierantoni, P, Evans, MN and Elrick, M (2023) Carbonate uranium isotopes across Cretaceous OAE 2 in southern Mexico: new constraints on the global spread of marine anoxia and organic carbon burial. Palaeogeography, Palaeoclimatology, Palaeoecology 628, 111756.CrossRefGoogle Scholar
Kunert, A and Kendall, B (2023) Global ocean redox changes before and during the Toarcian Oceanic Anoxic Event. Nature Communications 14, 110.CrossRefGoogle ScholarPubMed
Küspert, W (1982) Environmental changes during oil shale deposition as deduced from stable isotope ratios. In Cyclic and Event Stratification (eds Einsele, G and Seilacher, A). pp. 482501. Berlin: Springer-Verlag.CrossRefGoogle Scholar
LaGrange, MT, Konhauser, KO, Catuneanu, O, Harris, BS, Playter, TL and Gingras, MK (2020) Sequence stratigraphy in organic-rich marine mudstone successions using chemostratigraphic datasets. Earth-Science Reviews 203, 103137.CrossRefGoogle Scholar
Large, RR, Bull, SW and Maslennikov, VV (2011) A carbonaceous sedimentary source-rock model for Carlin-type and orogenic gold deposits. Economic Geology 106, 331358.CrossRefGoogle Scholar
Lau, KV and Hardisty, DS (2022) Modeling the impacts of diagenesis on carbonate paleoredox proxies. Geochimica et Cosmochimica Acta 337, 123139.CrossRefGoogle Scholar
Lau, KV, Romaniello, SJ and Zhang, F (2019) The Uranium Isotope Paleoredox Proxy. Elements in Geochemical Tracers in Earth System Science, Cambridge: Cambridge University Press, 28 pp.CrossRefGoogle Scholar
Law, CA (1999) Evaluating source rocks. In Exploring for Oil and Gas Traps (eds Beaumont, EA and Foster, NH). AAPG Treatise Handbook pp. 6-1–6-41. Tulsa: American Association of Petroleum Geologists.Google Scholar
Lazar, OR, Bohacs, KM, Macquaker, JHS, Schieber, J and Demko, TM (2015) Capturing key attributes of fine-grained sedimentary rocks in outcrops, cores, and thin sections: nomenclature and description guidelines. Journal of Sedimentary Research 85, 230246.CrossRefGoogle Scholar
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, 7198.CrossRefGoogle Scholar
Levasseur, S, Birck, J-L and Allègre, CJ (1999) The osmium riverine flux and the oceanic mass balance of osmium. Earth and Planetary Science Letters 174, 723.CrossRefGoogle Scholar
Li, Q, McArthur, JM, Thirlwall, MF, Turchyn, AV, Page, K, Bradbury, HJ, Weis, R and Lowry, D (2021) Testing for ocean acidification during the Early Toarcian using δ44/40Ca and δ88/86Sr. Chemical Geology 574, 120228.CrossRefGoogle Scholar
Li, Z, Schieber, J and Pedersen, PK (2021) On the origin and significance of composite particles in mudstones: examples from the Cenomanian Dunvegan Formation. Sedimentology 68, 737754.CrossRefGoogle Scholar
Littke, R, Klussmann, U, Krooss, B and Leythaeuser, D (1991) Quantification of loss of calcite, pyrite, and organic matter due to weathering of Toarcian black shales and effects on kerogen and bitumen characteristics. Geochimica et Cosmochimica Acta 55, 33693378.CrossRefGoogle Scholar
Little, CTS (1995) The Pliensbachian–Toarcian (Lower Jurassic) extinction event. PhD thesis, University of Bristol, Bristol, 144 pp. Published thesis http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.294903.Google Scholar
Little, CTS and Benton, MJ (1995) Early Jurassic mass extinction: a global long-term event. Geology 23, 495498.2.3.CO;2>CrossRefGoogle Scholar
Little, CTS (1996) The Pliensbachian–Toarcian (Lower Jurassic) extinction event. In The Cretaceous–Tertiary Event and other Catastrophes in Earth History (eds Ryder, G, Fastovsky, D and Gartner, S), pp. 505512. Geological Society of America Special Papers 307.Google Scholar
Little, SH, Vance, D, Lyons, TW and McManus, J (2015) Controls on trace metal authigenic enrichment in reducing sediments: insights from modern oxygen-deficient settings. American Journal of Science 315, 77119.CrossRefGoogle Scholar
Littler, K, Hesselbo, SP and Jenkyns, HC (2010) A carbon-isotope perturbation at the Pliensbachian–Toarcian boundary: evidence from the Lias Group, NE England. Geological Magazine 147, 181192.CrossRefGoogle Scholar
Liu, J, Cao, J, He, T, Liang, F, Pu, J and Wang, Y (2022) Lacustrine redox variations in the Toarcian Sichuan Basin across the Jenkyns Event. Global and Planetary Change 215, 103860.CrossRefGoogle Scholar
Liu, M, Ji, CJ, Hu, HW, Xia, GQ, Yi, HS, Them, TR II, Sun, P and Chen, DZ (2021) Variations in microbial ecology during the Toarcian Oceanic Anoxic Event (Early Jurassic) in the Qiangtang Basin, Tibet: evidence from biomarker and carbon isotopes. Palaeogeography, Palaeoclimatology, Palaeoecology 580, 110626.CrossRefGoogle Scholar
Locarnini, RA, Mishonov, AV, Baranova, OK, Boyer, TP, Zweng, MM, Garcia, HE, Reagan, JR, Seidov, D, Weathers, KW, Paver, CR and Smolyar, IV (2019) World Ocean Atlas 2018, Volume 1: Temperature. NOAA Atlas NESDIS 81, Silver Springs MD: National Oceanic and Atmospheric Administration, 52 pp.Google Scholar
Lord, AR (2019) Fossils from the Lias of the Yorkshire Coast. Palaeontological Association Field Guide. London: Palaeontological Association, 403 pp.Google Scholar
Lu, Z, Jenkyns, HC and Rickaby, REM (2010) Iodine to calcium ratios in marine carbonate as a paleo-redox proxy during oceanic anoxic events. Geology 38, 11071110.CrossRefGoogle Scholar
Luther, GWI (2023) Review on the physical chemistry of iodine transformations in the oceans. Frontiers in Marine Science 10, 1085618.CrossRefGoogle Scholar
Luttinen, A, Kurhila, M, Puttonen, R, Whitehouse, M and Andersen, T (2022) Periodicity of Karoo rift zone magmatism inferred from zircon ages of silicic rocks: implications for the origin and environmental impact of the large igneous province. Gondwana Research 107, 107122.CrossRefGoogle Scholar
Lyons, TW and Reinhard, CT (2012) Deoxygenation in warming oceans—looking back to the future. Geology 40, 671672.CrossRefGoogle Scholar
Macquaker, JHS, Keller, MA and Davies, SJ (2010) Algal blooms and “marine snow”: mechanisms that enhance preservation of organic carbon in ancient fine-grained sediments. Journal of Sedimentary Research 80, 934942.CrossRefGoogle Scholar
Macquaker, JHS and Taylor, KG (1996) A sequence-stratigraphic interpretation of a mudstone-dominated succession: the Lower Jurassic Cleveland Ironstone Formation, UK. Journal of the Geological Society 153, 759770.CrossRefGoogle Scholar
Maher, W and Butler, E (1988) Arsenic in the marine environment. Applied Organometallic Chemistry 3, 191214.CrossRefGoogle Scholar
Manabe, S and Bryan, K Jr (1985) CO2-induced change in a coupled ocean–atmosphere model and its paleoclimatic implications. Journal of Geophysical Research: Oceans 90, 1168911707.CrossRefGoogle Scholar
Marley, J (1857 ) Cleveland Ironstone. Outline of the main or thick stratified bed, its discovery, application and results, in connection with the iron works in the north of England. Transactions North of England Institute of Mining Engineers 5, 165219.Google Scholar
Marquez, RTC, Tejada, MLG, Suzuki, K, Peleo-Alampay, AM, Goto, KT, Hyun, S and Senda, R (2017) The seawater osmium isotope record of South China Sea: implications on its history and evolution. Marine Geology 394, 98115.CrossRefGoogle Scholar
Marsaglia, KM and Klein, GD (1983) The paleogeography of Paleozoic and Mesozoic storm depositional systems. Journal of Geology 91, 117142.CrossRefGoogle Scholar
Marshall, JD and Pirrie, D (2013) Carbonate concretions — explained. Geology Today 29, 5362.CrossRefGoogle Scholar
Martin, KD (2004) A re-evaluation of the relationship between trace fossils and dysoxia. In The Application of Ichnology to Palaeoenvironmental and Stratigraphic Analysis (ed McIlroy, D), pp. 141156. Geological Society London, Special Publications 228.Google Scholar
Martinez, M, Krencker, F-N, Mattioli, E and Bodin, S (2017) Orbital chronology of the Pliensbachian – Toarcian transition from the Central High Atlas Basin (Morocco). Newsletters on Stratigraphy 50, 4769.CrossRefGoogle Scholar
März, C, Riedinger, N, Sena, C and Kasten, S (2018) Phosphorus dynamics around the sulphate–methane transition in continental margin sediments: authigenic apatite and Fe(II) phosphates. Marine Geology 404, 8496.CrossRefGoogle Scholar
Maynard, JB (1986) Geochemistry of oolitic iron ores, an electron microprobe study. Economic Geology 81, 14731483.CrossRefGoogle Scholar
McArthur, JM (2019) Early Toarcian black shales: a response to an oceanic anoxic event or anoxia in marginal basins? Chemical Geology 522, 7183.CrossRefGoogle Scholar
McArthur, JM, Algeo, TJ, van de Schootbrugge, B, Li, Q and Howarth, RJ (2008) Basinal restriction, black shales, Re–Os dating, and the Early Toarcian (Jurassic) oceanic anoxic event. Paleoceanography 23, PA4217.CrossRefGoogle Scholar
McArthur, JM, Donovan, DT, Thirlwall, MF, Fouke, BW and Mattey, D (2000) Strontium isotope profile of the early Toarcian (Jurassic) oceanic anoxic event, the duration of ammonite biozones, and belemnite palaeotemperatures. Earth and Planetary Science Letters 179, 269285.CrossRefGoogle Scholar
McArthur, JM, Howarth, RJ, Shields, GA and Zhou, Y (2020) Strontium isotope stratigraphy. In Geologic Time Scale 2020 (eds Gradstein, FM, Ogg, JG, Schmitz, MD and Ogg, GM). pp. 211238. Amsterdam: Elsevier.CrossRefGoogle Scholar
McArthur, JM, Steuber, T, Page, KN and Landman, NH (2016) Sr-isotope stratigraphy: assigning time in the Campanian, Pliensbachian, Toarcian, and Valanginian. Journal of Geology 124, 569586.CrossRefGoogle Scholar
McDonald, BS, Partin, CA, Sageman, B and Holmden, C (2022) Uranium isotope reconstruction of ocean deoxygenation during OAE 2 hampered by uncertainties in fractionation factors and local U-cycling. Geochimica et Cosmochimica Acta 331, 143164.CrossRefGoogle Scholar
McElwain, JC, Wade-Murphy, J and Hesselbo, SP (2005) Changes in carbon dioxide during an oceanic anoxic event linked to intrusion into Gondwana coals. Nature 435, 479482.CrossRefGoogle ScholarPubMed
McKee, ED and Weir, GW (1953) Terminology for stratification and cross-stratification in sedimentary rocks. GSA Bulletin 64, 381390.CrossRefGoogle Scholar
McLennan, SM (1993) Weathering and global denudation. Journal of Geology 101, 295303.CrossRefGoogle Scholar
Meister, C, Aberhan, M, Blau, J, Dommergues, JL, Feist-Burkhardt, S, Hailwood, EA, Hart, M, Hesselbo, SP, Hounslow, MW, Hylton, M, Morton, N, Page, K and Price, GD (2006) The Global Boundary Stratotype Section and Point (GSSP) for the base of the Pliensbachian Stage (Lower Jurassic), Wine Haven, Yorkshire, UK. Episodes 29, 93106.CrossRefGoogle Scholar
Menini, A, Mattioli, E, Hesselbo, SP, Ruhl, M and Suan, G (2021) Primary v. carbonate production in the Toarcian, a case study from the Llanbedr (Mochras Farm) borehole, Wales. In Carbon Cycle and Ecosystem Response to the Jenkyns Event in the Early Toarcian (Jurassic). (eds Reolid, M, Duarte, LV, Mattioli, E and Ruebsam, W), pp. 5981. Geological Society London, Special Publications 514.Google Scholar
Merkel, A and Munnecke, A (2023) Glendonite-bearing concretions from the upper Pliensbachian (Lower Jurassic) of South Germany: indicators for a massive cooling in the European epicontinental sea. Facies 69, 10.CrossRefGoogle Scholar
Meyers, PA (2014) Why are the δ13Corg values in Phanerozoic black shales more negative than in modern marine organic matter? Geochemistry Geophysics Geosystems 15, 30853106.CrossRefGoogle Scholar
Miller, CA, Peucker-Ehrenbrink, B, Walker, BD and Marcantonio, F (2011) Re-assessing the surface cycling of molybdenum and rhenium. Geochimica et Cosmochimica Acta 75, 71467179.CrossRefGoogle Scholar
Min, H, Zhang, T, Li, Y, Zhao, S, Li, J, Lin, D and Wang, J (2019) The albitization of K-feldspar in organic- and silt-rich fine-grained rocks of the Lower Cambrian Qiongzhusi Formation in the southwestern Upper Yangtze Region, China. Minerals 9, 122.CrossRefGoogle Scholar
Monteiro, FM, Pancost, RD, Ridgwell, A and Donnadieu, Y (2012) Nutrients as the dominant control on the spread of anoxia and euxinia across the Cenomanian–Turonian oceanic anoxic event (OAE2): model–data comparison. Paleoceanography 27, PA4209.CrossRefGoogle Scholar
Montero-Serrano, JC, Föllmi, KB, Adatte, T, Spangenberg, JE, Tribovillard, N, Fantasia, A and Suan, G (2015) Continental weathering and redox conditions during the early Toarcian Oceanic Anoxic Event in the northwestern Tethys: insight from the Posidonia Shale section in the Swiss Jura Mountains. Palaeogeography, Palaeoclimatology, Palaeoecology 429, 8399.CrossRefGoogle Scholar
Montoya-Pino, C, Weyer, S, Anbar, AD, Pross, J, Oschmann, W, van de Schootbrugge, B and Arz, HW (2010) Global enhancement of ocean anoxia during Oceanic Anoxic Event 2: a quantitative approach using U isotopes. Geology 38, 315318.CrossRefGoogle Scholar
Moor, C, Lymberopoulou, T and Dietrich, VJ (2001) Determination of heavy metals in soils, sediments and geological materials by ICP-AES and ICP-MS. Mikrochimica Acta 136, 123128.CrossRefGoogle Scholar
Morel, FMM and Price, NM (2003) The biogeochemical cycles of trace metals in the oceans. Science 300, 944947.CrossRefGoogle ScholarPubMed
Morford, JL and Emerson, S (1999) The geochemistry of redox sensitive trace metals in sediments. Geochimica et Cosmochimica Acta 63, 17351750.CrossRefGoogle Scholar
Morford, JL, Martin, WR and Carney, CM (2012) Rhenium geochemical cycling: insights from continental margins. Chemical Geology 324–325, 7386.CrossRefGoogle Scholar
Morris, KA (1979) A classification of Jurassic marine shale sequences: an example from the Toarcian (Lower Jurassic) of Great Britain. Palaeogeography, Palaeoclimatology, Palaeoecology 26, 117126.CrossRefGoogle Scholar
Morris, KA (1980) Comparison of major sequences of organic-rich mud deposition in the British Jurassic. Journal of the Geological Society 137, 157170.CrossRefGoogle Scholar
Moulin, M, Fluteau, F, Courtillot, V, Marsh, J, Delpech, G, Quidelleur, X and Gérard, M (2017) Eruptive history of the Karoo lava flows and their impact on early Jurassic environmental change. Journal of Geophysical Research: Solid Earth 122, 738772.CrossRefGoogle Scholar
Mücke, A and Farshad, F (2005) Whole-rock and mineralogical composition of Phanerozoic ooidal ironstones: comparison and differentiation of types and subtypes. Ore Geology Reviews 26, 227262.CrossRefGoogle Scholar
Müller, T, Jurikova, H, Gutjahr, M, Tomasovych, A, Schlogl, J, Liebetrau, V, Duarte, LV, Milovsky, R, Suan, G, Mattioli, E, Pittet, B and Eisenhauer, A (2020) Ocean acidification during the early Toarcian extinction event: evidence from boron isotopes in brachiopods. Geology 48, 11841188.CrossRefGoogle Scholar
Müller, T, Price, GD, Bajnai, D, Nyerges, A, Kesjár, D, Raucsik, B, Varga, A, Judik, K, Fekete, J, May, Z and Pálfy, J (2017) New multiproxy record of the Jenkyns Event (also known as the Toarcian Oceanic Anoxic Event) from the Mecsek Mountains (Hungary): differences, duration and drivers. Sedimentology 64, 6686.CrossRefGoogle Scholar
Myers, KJ and Wignall, PB (1987) Understanding Jurassic organic rich mudrocks – new concepts using gamma-ray spectrometry and palaeoecology: Examples from the Kimmeridge Clay of Dorset and the Jet Rock of Yorkshire. In Marine Clastic Sedimentology (eds Leggett, JK and Zuffa, GG), pp. 172189. London: Graham and Trotman.CrossRefGoogle Scholar
Myers, KJ (1989) The origin of the Lower Jurassic Cleveland Ironstone Formation of North-East England: evidence from portable gamma-ray spectrometry. In Phanerozoic Ironstones (eds Young, TP and Gordon Taylor, WE), pp. 221228. Geological Society London, Special Publications 46.Google Scholar
Näslund, J (2021) Mercury in fossil leaves as a proxy for tracking Large Igneous Province volcanism. BSc thesis, Uppsala University, Uppsala, 18 pp. Published thesis https://www.diva-portal.org/smash/get/diva2:1574588/FULLTEXT01.pdf.Google Scholar
Nesbitt, HW and Young, GM (1982) Early Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature 299, 715717.CrossRefGoogle Scholar
Newton, RJ, Reeves, EP, Kafousia, N, Wignall, PB, Bottrell, SH and Sha, J-G (2011) Low marine sulfate concentrations and the isolation of the European epicontinental sea during the Early Jurassic. Geology 39, 710.CrossRefGoogle Scholar
Nie, Y, Fu, X, Liang, J, Wei, H, Chen, Z, Lin, F, Zeng, S, Wu, Y, Zou, Y and Mansour, A (2023) The Toarcian Oceanic Anoxic Event in a shelf environment (Eastern Tethys): implications for weathering and redox conditions. Sedimentary Geology 455, 106476.CrossRefGoogle Scholar
Nielsen, SG, Goff, M, Hesselbo, SP, Jenkyns, HC, LaRowe, DE and Lee, CTA (2011) Thallium isotopes in early diagenetic pyrite – A paleoredox proxy? Geochimica et Cosmochimica Acta 75, 66906704.CrossRefGoogle Scholar
Nielsen, SG, Rehkämper, M and Prytulak, J (2017) Investigation and application of thallium isotope fractionation. Reviews in Mineralogy and Geochemistry 82, 759798.CrossRefGoogle Scholar
Nikitenko, BL and Mickey, MB (2004) Foraminifera and ostracodes across the Pliensbachian–Toarcian boundary in the Arctic Realm (stratigraphy, palaeobiogeography and biofacies). In The Palynology and Micropalaeontology of Boundaries (eds AB Beaudoin and MJH Head). Geological Society London Special Publications 230, 137174.CrossRefGoogle Scholar
Nordt, L, Breecker, D and White, J (2022) Jurassic greenhouse ice-sheet fluctuations sensitive to atmospheric CO2 dynamics. Nature Geoscience 15, 5459.CrossRefGoogle Scholar
O’Brien, NR (1990) Significance of lamination in Toarcian (Lower Jurassic) shales from Yorkshire, Great Britain. Sedimentary Geology 67, 2534.CrossRefGoogle Scholar
Oliveira, LCV, Rodriguez, R, Duarte, LV and Lemos, VB (2006) Avaliação do potencial gerador de petróleo e interpretação paleoambiental com base em biomarcadores e isótopos estáveis de carbono da seção Pliensbaquiano – Toarciano inferior (Jurássico Inferior) da região de Peniche (Bacia Lusitânica, Portugal). Boletim de Geociências da Petrobras, Rio de Janeiro 14, 207234.Google Scholar
Ostrander, CM, Owens, JD and Nielsen, SG (2017) Constraining the rate of oceanic deoxygenation leading up to a Cretaceous Oceanic Anoxic Event (OAE-2: ∼94 Ma). Science Advances 3, e1701020.CrossRefGoogle ScholarPubMed
Otto-Bliesner, BL, Brady, EC and Shields, C (2002) Late Cretaceous ocean: coupled simulations with the National Center for Atmospheric Research climate system model. Journal of Geophysical Research - Atmospheres 107, 4019.CrossRefGoogle Scholar
Owens, JD (2019) Application of Thallium Isotopes Tracking Marine Oxygenation Through Manganese Oxide Burial. Elements in Geochemical Tracers in Earth System Science Cambridge: Cambridge University Press, 21 pp.CrossRefGoogle Scholar
Owens, JD, Gill, BC, Jenkyns, HC, Bates, SM, Severmann, S, Kuypers, MMM, Woodfine, RG and Lyons, TW (2013) Sulfur isotopes track the global extent and dynamics of euxinia during Cretaceous Oceanic Anoxic Event 2. Proceedings of the National Academy of Sciences 110, 1840718412.CrossRefGoogle ScholarPubMed
Owens, JD, Lyons, TW and Lowery, CM (2018) Quantifying the missing sink for global organic carbon burial during a Cretaceous oceanic anoxic event. Earth and Planetary Science Letters 499, 8394.CrossRefGoogle Scholar
Owens, JD, Nielsen, SG, Horner, TJ, Ostrander, CM and Peterson, LC (2017) Thallium-isotopic compositions of euxinic sediments as a proxy for global manganese-oxide burial. Geochimica et Cosmochimica Acta 213, 291307.CrossRefGoogle Scholar
Owens, JD, Reinhard, CT, Rohrssen, M, Love, GD and Lyons, TW (2016) Empirical links between trace metal cycling and marine microbial ecology during a large perturbation to Earth’s carbon cycle. Earth and Planetary Science Letters 449, 407417.CrossRefGoogle Scholar
Oxburgh, R (2001) Residence time of osmium in the oceans. Geochemistry, Geophysics, Geosystems 2, 2000GC000104.CrossRefGoogle Scholar
Page, KN (2017) From Oppel to Callomon (and beyond): building a high resolution ammonite-based biochronology for the Jurassic System. Lethaia 50, 336355.CrossRefGoogle Scholar
Pálfy, J and Smith, PL (2000) Synchrony between Early Jurassic extinction, oceanic anoxic event, and the Karoo–Ferrar flood basalt volcanism. Geology 28, 747750.2.0.CO;2>CrossRefGoogle Scholar
Pancost, RD, Crawford, N, Magness, S, Turner, A, Jenkyns, HC and Maxwell, JR (2004) Further evidence for the development of photic-zone euxinic conditions during Mesozoic oceanic anoxic events. Journal of the Geological Society 161, 353364.CrossRefGoogle Scholar
Pankhurst, RJ, Riley, TR, Fanning, CM and Kelley, SP (2000) Episodic silicic volcanism in Patagonia and the Antarctic Peninsula: chronology of magmatism associated with the break-up of Gondwana. Journal of Petrology 41, 605625.CrossRefGoogle Scholar
Papadomanolaki, NM, Lenstra, WK, Wolthers, M and Slomp, CP (2022) Enhanced phosphorus recycling during past oceanic anoxia amplified by low rates of apatite authigenesis. Science Advances 8, 112.CrossRefGoogle ScholarPubMed
Parkinson, DN (1996) Gamma-ray spectrometry as a tool for stratigraphical interpretation: examples from the western European Lower Jurassic. In Sequence Stratigraphy in British Geology (eds SP Hesselbo and DN Parkinson). Geological Society Special Publication 103, 231255.CrossRefGoogle Scholar
Paul, KM, van Helmond, NAGM, Slomp, CP, Jokinen, SA, Virtasalo, JJ, Filipsson, HL and Jilbert, T (2023) Sedimentary molybdenum and uranium: improving proxies for deoxygenation in coastal depositional environments. Chemical Geology 615, 118.CrossRefGoogle Scholar
Paytan, A, Kastner, M, Campbell, D and Thiemens, MH (2004) Seawater sulfur isotope fluctuations in the Cretaceous. Science 304, 16631665.CrossRefGoogle ScholarPubMed
Paytan, A, Yao, W and Gray, ET (2020) Sulfur isotope stratigraphy. In The Geologic Time Scale 2020 (eds Gradstein, F, Ogg, JG and Ogg, G). 1, pp. 259278. Amsterdam: Elsevier.CrossRefGoogle Scholar
Pearce, CR, Cohen, AS, Coe, AL and Burton, KW (2008) Molybdenum isotope evidence for global ocean anoxia coupled with perturbations to the carbon cycle during the Early Jurassic. Geology 36, 231234.CrossRefGoogle Scholar
Percival, LME, Bergquist, BA, Mather, TA and Sanei, H (2021) Sedimentary mercury enrichments as a tracer of Large Igneous Province volcanism. In Large Igneous Provinces: A Driver of Global Environmental and Biotic Changes (eds Ernst, RE, Dickson, AJ and Bekker, A), pp. 247262. Geophysical Monographs 255.CrossRefGoogle Scholar
Percival, LME, Cohen, AS, Davies, MK, Dickson, AJ, Hesselbo, SP, Jenkyns, HC, Leng, MJ, Mather, TA, Storm, MS and Xu, W (2016) Osmium isotope evidence for two pulses of increased continental weathering linked to Early Jurassic volcanism and climate change. Geology 44, 759762.CrossRefGoogle Scholar
Percival, LME, Witt, MLI, Mather, TA, Hermoso, M, Jenkyns, HC, Hesselbo, SP, Al-Suwaidi, AH, Storm, MS, Xu, W and Ruhl, M (2015) Globally enhanced mercury deposition during the end-Pliensbachian extinction and Toarcian OAE: a link to the Karoo–Ferrar Large Igneous Province. Earth and Planetary Science Letters 428, 267280.CrossRefGoogle Scholar
Peti, L and Thibault, N (2017) Abundance and size changes in the calcareous nannofossil Schizosphaerella – Relation to sea-level, the carbonate factory and palaeoenvironmental change from the Sinemurian to earliest Toarcian of the Paris Basin. Palaeogeography, Palaeoclimatology, Palaeoecology 485, 271282.CrossRefGoogle Scholar
Peti, L, Thibault, N, Clemence, ME, Korte, C, Dommergues, JL, Bougeault, C, Pellenard, P, Jelby, ME and Ullmann, CV (2017) Sinemurian–Pliensbachian calcareous nannofossil biostratigraphy and organic carbon isotope stratigraphy in the Paris Basin: calibration to the ammonite biozonation of NW Europe. Palaeogeography, Palaeoclimatology, Palaeoecology 468, 142161.CrossRefGoogle Scholar
Peti, L, Thibault, N, Korte, C, Ullmann, CV, Cachão, M, Fibæk, M (2021) Environmental drivers of size changes in lower Jurassic Schizosphaerella spp. Marine Micropaleontology 168, 102053.CrossRefGoogle Scholar
Peucker-Ehrenbrink, B and Ravizza, G (2002) The marine osmium isotope record. Terra Nova 12, 205219.CrossRefGoogle Scholar
Peucker-Ehrenbrink, B and Ravizza, GE (2020) Osmium isotope stratigraphy. In The Geologic Time Scale 2020 (eds Gradstein, F, Ogg, JG and Ogg, G). pp. 239257. Amsterdam: Elsevier.CrossRefGoogle Scholar
Phillips, J (1829) Illustrations of the geology of Yorkshire or a description of the strata and organic remains of the Yorkshire coast: accompanied by a geological map, sections and plates of the fossil plants and animals. York: J. Phillips, 192 pp.Google Scholar
Piazza, V, Ullmann, CV and Aberhan, M (2020) Ocean warming affected faunal dynamics of benthic invertebrate assemblages across the Toarcian Oceanic Anoxic Event in the Iberian Basin (Spain). PLoS One 15, 127.CrossRefGoogle ScholarPubMed
Piper, DZ and Calvert, SE (2009) A marine biogeochemical perspective on black shale deposition. Earth-Science Reviews 95, 6396.CrossRefGoogle Scholar
Poulsen, CJ, Barron, EJ, Arthur, MA and Peterson, WH (2001) Response of the mid-Cretaceous global oceanic circulation to tectonic and CO2 forcings. Paleoceanography 16, 576592.CrossRefGoogle Scholar
Powell, JH (1984) Lithostratigraphical nomenclature of the Lias Group in the Yorkshire Basin. Proceedings of the Yorkshire Geological Society 45, 5157.CrossRefGoogle Scholar
Powell, JH (2010) Jurassic sedimentation in the Cleveland Basin: a review. Proceedings of the Yorkshire Geological Society 58, 2172.CrossRefGoogle Scholar
Pye, K (1985) Electron microscope analysis of zoned dolomite rhombs in the Jet Rock Formation (Lower Toarcian) of the Whitby area, U.K. Geological Magazine 122, 279286.CrossRefGoogle Scholar
Pye, K and Krinsley, DH (1986) Microfabric, mineralogy and early diagenetic history of the Whitby Mudstone Formation (Toarcian), Cleveland Basin, U.K. Geological Magazine 123, 191203.CrossRefGoogle Scholar
Quinby-Hunt, MS and Wilde, P (1994) Thermodynamic zonation in the black shale facies based on iron–manganese–vanadium content. Chemical Geology 113, 297317.CrossRefGoogle Scholar
Rahman, MW, Rimmer, SM and Rowe, HD (2018) The impact of rapid heating by intrusion on the geochemistry and petrography of coals and organic-rich shales in the Illinois Basin. International Journal of Coal Geology 187, 4553.CrossRefGoogle Scholar
Raiswell, R (1988) Chemical model for the origin of minor limestone-shale cycles by anaerobic methane oxidation. Geology 16, 641644.2.3.CO;2>CrossRefGoogle Scholar
Raiswell, R and Berner, RA (1985) Pyrite formation in euxinic and semi-euxinic sediments. American Journal of Science 285, 710724.CrossRefGoogle Scholar
Raiswell, R, Bottrell, SH, Al-Biatty, HJ and Tan, MM (1993) The influence of bottom water oxygenation and reactive iron content on sulfur incorporation into bitumens from Jurassic marine shales. American Journal of Science 293, 569596.CrossRefGoogle Scholar
Raiswell, R, Buckley, F, Berner, RA and Anderson, TF (1988) Degree of pyritization of iron as a paleoenvironmental indicator of bottom-water oxygenation. Journal of Sedimentary Petrology 58, 812819.Google Scholar
Raiswell, R and Canfield, DE (1998) Sources of iron for pyrite formation in marine sediments. American Journal of Science 298, 219245.CrossRefGoogle Scholar
Raiswell, R, Hardisty, DS, Lyons, TW, Canfield, DE, Owens, JD, Planavsky, NJ, Poulton, SW and Reinhard, CT (2018) The iron paleoredox proxies: a guide to the pitfalls, problems and proper practice. American Journal of Science 318, 491526.CrossRefGoogle Scholar
Raiswell, R (1987) Non-steady state microbiological diagenesis and the origin of concretions and nodular limestones. In Diagenesis of Sedimentary Sequences (ed Marshall, JD), pp. 4154. Geological Society London, Special Publications 36.Google Scholar
Raiswell, R, Newton, R and Wignall, PB (2001) An indicator of water-column anoxia: resolution of biofacies variations in the Kimmeridge Clay (Upper Jurassic, U.K.). Journal of Sedimentary Research 71, 286294.CrossRefGoogle Scholar
Raucsik, B and Varga, A (2008) Climato-environmental controls on clay mineralogy of the Hettangian–Bajocian successions of the Mecsek Mountains, Hungary: an evidence for extreme continental weathering during the early Toarcian oceanic anoxic event. Palaeogeography, Palaeoclimatology, Palaeoecology 265, 113.CrossRefGoogle Scholar
Raup, DM and Sepkoski, JJ (1984) Periodicity of extinctions in the geologic past. Proceedings of the National Acadamy of Science USA 81, 801805.CrossRefGoogle ScholarPubMed
Rawson, PF, Greensmith, JT and Shalaby, SE (1983) Coarsening-upwards cycles in the uppermost Staithes and Cleveland Ironstone Formations (Lower Jurassic) of the Yorkshire coast, England. Proceedings of the Geologists’ Association 94, 9193.CrossRefGoogle Scholar
Rawson, PF and Wright, JK (1992) The Yorkshire Coast. Geologists’ Association Field Guide London: Geologists’ Association, 120 pp.Google Scholar
Rawson, PF and Wright, JK (2000) The Yorkshire Coast. Geologists’ Association Guides 34, London: Geologists’ Association, 130 pp.Google Scholar
Rawson, PF and Wright, JK (2018) Geology of the Yorkshire Coast. Geologists’ Association Guides 34, London: Geologists’ Association, 178 pp.Google Scholar
Rawson, PF and Wright, JK (1995) Jurassic of the Cleveland Basin, North Yorkshire. In Field Geology of the British Jurassic (ed Taylor, PD), pp. 173208. Bath: The Geological Society, London.Google Scholar
Redfield, AC, Ketchum, BH and Richards, FA (1963) The influence of organisms on the composition of seawater. In The Composition of Seawater: Comparative and Descriptive Oceanography. The Sea: Ideas and Observations on Progress in the Study of the Seas, 2 (ed Hill, MN), pp. 2677. New York: Interscience Publishers.Google Scholar
Reershemius, T and Planavsky, NJ (2021) What controls the duration and intensity of ocean anoxic events in the Paleozoic and the Mesozoic? Earth-Science Reviews 221, 103787.CrossRefGoogle Scholar
Rees, CE, Jenkins, WJ and Monster, J (1978) The sulphur isotopic composition of ocean water sulphate. Geochimica et Cosmochimica Acta 42, 377381.CrossRefGoogle Scholar
Reinhard, CT, Planavsky, NJ, Robbins, LJ, Partin, CA, Gill, BC, Lalonde, SV, Bekker, A, Konhauser, KO, Lyons, TW (2013) Proterozoic ocean redox and biogeochemical stasis. Proceedings of the National Academy of Sciences 110, 53575362.CrossRefGoogle ScholarPubMed
Remirez, MN and Algeo, TJ (2020a) Carbon-cycle changes during the Toarcian (Early Jurassic) and implications for regional versus global drivers of the Toarcian oceanic anoxic event. Earth-Science Reviews 209, 103283.CrossRefGoogle Scholar
Remírez, MN and Algeo, TJ (2020b) Paleosalinity determination in ancient epicontinental seas: a case study of the T-OAE in the Cleveland Basin (UK). Earth-Science Reviews 201, 115.CrossRefGoogle Scholar
Reolid, M, Emanuela, M, Nieto, LM and Rodríguez-Tovar, FJ (2014) The early Toarcian Oceanic Anoxic Event in the External Subbetic (Southiberian Palaeomargin, westernmost Tethys): geochemistry, nannofossils and ichnology. Palaeogeography, Palaeoclimatology, Palaeoecology 411, 7994.CrossRefGoogle Scholar
Reolid, M, Mattioli, E, Duarte, LV and Marok, A (2020) The Toarcian Oceanic Anoxic Event and the Jenkyns Event (IGCP-655 final report). Episodes 43, 112.CrossRefGoogle Scholar
Reolid, M, Mattioli, E, Duarte, LV and Ruebsam, W (2021 ) The Toarcian Oceanic Anoxic Event: Where do we stand? In Carbon Cycle and Ecosystem Response to the Jenkyns Event in the Early Toarcian (Jurassic) (eds Reolid, M, Mattioli, E, Duarte, LV and Ruebsam, W), pp. 111. Geological Society London, Special Publications 514.Google Scholar
Richey, JD, Nordt, L, White, JD and Breecker, DO (2023) ISOORG23: an updated compilation of stable carbon isotope data of terrestrial organic materials for the Cenozoic and Mesozoic. Earth-Science Reviews 241, 104439.CrossRefGoogle Scholar
Rickard, D (2019) Sedimentary pyrite framboid size-frequency distributions: a meta-analysis. Palaeogeography, Palaeoclimatology, Palaeoecology 522, 6275.CrossRefGoogle Scholar
Riegraf, W (1985) Microfauna, Biostratigraphie und Fazies im Unteren Toarcium Südwestdeutschlands und Vergleiche mit benachbarten Gebieten. Tübinger Mikropaläontologische Mitteilungen 3, 1232.Google Scholar
Riegraf, W (1982) The bituminous Lower Toarcian at the Truc-de-Balduc near Mende (Departement de la Lozère, S-France). In Cyclic and Event Stratification (eds Einsele, S and Seilacher, A). pp. 506511. New York: Springer.CrossRefGoogle Scholar
Röhl, H-J, Schmid-Röhl, A, Oschmann, W, Frimmel, A and Schwark, L (2001) The Posidonia Shale (Lower Toarcian) of SW-Germany: an oxygen-depleted ecosystem controlled by sea level and palaeoclimate. Palaeogeography, Palaeoclimatology, Palaeoecology 165, 2752.CrossRefGoogle Scholar
Rosales, I, Barnolas, A, Goy, A, Sevillano, A, Armendáriz, M and López-García, JM (2018) Isotope records (C–O–Sr) of late Pliensbachian–early Toarcian environmental perturbations in the westernmost Tethys (Majorca Island, Spain). Palaeogeography, Palaeoclimatology, Palaeoecology 497, 168185.CrossRefGoogle Scholar
Rosales, I, Robles, S and Quesada, S (2004) Elemental and oxygen isotope composition of early Jurassic belemnites: salinity vs. temperature signals. Journal of Sedimentary Research 74, 342354.CrossRefGoogle Scholar
Rudnick, RL and Gao, S (2014) Composition of the continental crust. In Treatise on Geochemistry ( 2nd Edn) (eds Holland, HD and Turekian, KK). 4, pp. 151. Amsterdam: Elsevier.Google Scholar
Ruebsam, W and Al-Husseini, M (2020) Calibrating the Early Toarcian (Early Jurassic) with stratigraphic black holes (SBH). Gondwana Research 82, 317336.CrossRefGoogle Scholar
Ruebsam, W, Mattioli, E and Schwark, L (2022) Molecular fossils and calcareous nannofossils reveal recurrent phytoplanktonic events in the early Toarcian. Global and Planetary Change 212, 103812.CrossRefGoogle Scholar
Ruebsam, W, Mayer, B and Schwark, L (2019) Cryosphere carbon dynamics control early Toarcian global warming and sea level evolution. Global and Planetary Change 172, 440453.CrossRefGoogle Scholar
Ruebsam, W, Muller, T, Kovacs, J, Palfy, J and Schwark, L (2018) Environmental response to the early Toarcian carbon cycle and climate perturbations in the northeastern part of the West Tethys shelf. Gondwana Research 59, 144158.CrossRefGoogle Scholar
Ruebsam, W, Munzberger, P and Schwark, L (2014) Chronology of the Early Toarcian environmental crisis in the Lorraine Sub-Basin (NE Paris Basin). Earth and Planetary Science Letters 404, 273282.CrossRefGoogle Scholar
Ruebsam, W, Pienkowski, G and Schwark, L (2020a) Toarcian climate and carbon cycle perturbations – its impact on sea-level changes, enhanced mobilization and oxidation of fossil organic matter. Earth and Planetary Science Letters 546, 116417.CrossRefGoogle Scholar
Ruebsam, W, Reolid, M, Marok, A and Schwark, L (2020b) Drivers of benthic extinction during the early Toarcian (Early Jurassic) at the northern Gondwana paleomargin: implications for paleoceanographic conditions. Earth-Science Reviews 203, 103117.CrossRefGoogle Scholar
Ruebsam, W, Reolid, M, Sabatino, N, Masetti, D and Schwark, L (2020c) Molecular paleothermometry of the early Toarcian climate perturbation. Global and Planetary Change 195, 103351.CrossRefGoogle Scholar
Ruebsam, W, Reolid, M and Schwark, L (2020d) δ13C of terrestrial vegetation records Toarcian CO2 and climate gradients. Scientific Reports 10, 117.CrossRefGoogle ScholarPubMed
Ruebsam, W, Schmid-Röhl, A and Al-Husseini, M (2023) Astronomical timescale for the early Toarcian (Early Jurassic) Posidonia Shale and global environmental changes. Palaeogeography, Palaeoclimatology, Palaeoecology 623, 111619.CrossRefGoogle Scholar
Ruebsam, W and Schwark, L (2024) Disparity between Toarcian Oceanic Anoxic Event and Toarcian carbon isotope excursion. International Journal of Earth Sciences https://doi.org/10.1007/s00531-024-02408-8.CrossRefGoogle Scholar
Ruebsam, W and Schwark, L (2021) Impact of a northern-hemispherical cryosphere on late Pliensbachian–early Toarcian climate and environment evolution. In Carbon Cycle and Ecosystem Response to the Jenkyns Event in the Early Toarcian (Jurassic) (eds Reolid, M, Duarte, LV, Mattioli, E and Ruebsam, W), pp. 359385. Geological Society London, Special Publications 514.Google Scholar
Ruhl, M, Hesselbo, SP, Jenkyns, HC, Xu, W, Silva, RL, Matthews, KJ, Mather, TA, Mac Niocaill, C and Riding, JB (2022) Reduced plate motion controlled timing of Early Jurassic Karoo–Ferrar large igneous province volcanism. Science Advances 8, eabo0866.CrossRefGoogle ScholarPubMed
Ruvalcaba Baroni, I, Palastanga, V and Stomp, CP (2020) Enhanced organic carbon burial in sediments of oxygen minimum zones upon ocean deoxygenation. Frontiers in Marine Science 6, 115.CrossRefGoogle Scholar
Ruvalcaba Baroni, I, Pohl, A, van Helmond, N, Papadomanolaki, NM, Coe, AL, Cohen, AS, van de Schootbrugge, B, Donnadieu, Y and Slomp, CP (2018) Ocean circulation in the Toarcian (Early Jurassic): a key control on deoxygenation and carbon burial on the European shelf. Paleoceanography and Paleoclimatology 33, 9941012.CrossRefGoogle Scholar
Sælen, G, Doyle, P and Talbot, MR (1996) Stable-isotope analyses of belemnite rostra from the Whitby Mudstone Fm, England: surface water conditions during deposition of a marine black shale. Palaios 11, 97117.Google Scholar
Sælen, G, Tyson, RV, Talbot, MR and Telnaes, N (1998) Evidence of recycling of isotopically light CO2 (aq) in stratified black shale basins: contrasts between the Whitby Mudstone and Kimmeridge Clay formations, United Kingdom. Geology 26, 747750.2.3.CO;2>CrossRefGoogle Scholar
Sælen, G, Tyson, RV, Telnaes, N and Talbot, MR (2000) Contrasting watermass conditions during deposition of the Whitby Mudstone (Lower Jurassic) and Kimmeridge Clay (Upper Jurassic) formations, UK. Palaeogeography, Palaeoclimatology, Palaeoecology 163, 163196.CrossRefGoogle Scholar
Salem, N-E (2013) Geochemical characterisation of the Pliensbachian–Toarcian boundary during the onset of the Toarcian Oceanic Anoxic Event, North Yorkshire, UK. PhD thesis, Newcastle University, Newcastle upon Tyne, 276 pp. Published thesis https://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.618173.Google Scholar
Schieber, J, Miclaus, C, Seserman, A, Liu, B and Teng, J (2019) When a mudstone was actually a “sand”: results of a sedimentological investigation of the bituminous marl formation (Oligocene), Eastern Carpathians of Romania. Sedimentary Geology 384, 1228.CrossRefGoogle Scholar
Schlanger, SO, Arthur, MA, Jenkyns, HC and Scholle, PA (1987) The Cenomanian–Turonian Oceanic Anoxic event, I. Stratigraphy and distribution of organic carbon-rich beds and the marine δ13C excursion. In Marine Petroleum Source Rocks (eds Brooks, J and Fleet, AJ), pp. 371399. Geological Society London, Special Publications 26.Google Scholar
Schlanger, SO and Jenkyns, HC (1976) Cretaceous oceanic anoxic events: causes and consequences. Geologie en Mijnbouw 55, 179184.Google Scholar
Schmitz, MD, Singer, BS and Rooney, AD (2020) Radioisotope geochronology. In The Geologic Time Scale 2020 (eds Gradstein, F, Ogg, JG and Ogg, G). 1, pp. 193209. Amsterdam: Elsevier.CrossRefGoogle Scholar
Schnyder, J, Pons, D, Yans, J, Tramoy, R and Abdulanova, S (2017) Integrated stratigraphy of a continental Pliensbachian–Toarcian Boundary (Lower Jurassic) section at Taskomirsay, Leontiev Graben, southwest Kazakhstan. In Geological Evolution of Central Asian Basins and the Western Tien Shan Range (eds Brunet, M-F, McCann, T and Sobel, ER), pp. 337356. Geological Society London, Special Publications 427.Google Scholar
Scholle, PA and Arthur, MA (1980) Carbon isotope fluctuation in Cretaceous pelagic limestones: potential stratigraphic and petroleum exploration tool. American Association of Petroleum Geologists Bulletin 64, 6787.Google Scholar
Scholz, F (2018) Identifying oxygen minimum zone-type biogeochemical cycling in Earth history using inorganic geochemical proxies. Earth-Science Reviews 184, 2945.CrossRefGoogle Scholar
Scholz, F, Siebert, C, Dale, AW and Frank, M (2017) Intense molybdenum accumulation in sediments underneath a nitrogenous water column and implications for the reconstruction of paleo-redox conditions based on molybdenum isotopes. Geochimica et Cosmochimica Acta 213, 400417.CrossRefGoogle Scholar
Schouten, S, Van Kaam-Peters, HME, Rijpstra, WIC, Schoell, M and Sinninghe Damsté, JSS (2000) Effects of an oceanic anoxic event on the stable carbon isotopic composition of early Toarcian carbon. American Journal of Science 300, 122.CrossRefGoogle Scholar
Schwark, L and Frimmel, A (2004) Chemostratigraphy of the Posidonia Black Shale, SW-Germany II. Assessment of extent and persistence of photic-zone anoxia using aryl isoprenoid distributions. Chemical Geology 206, 231248.CrossRefGoogle Scholar
Scott, C and Lyons, TW (2012) Contrasting molybdenum cycling and isotopic properties in euxinic versus non-euxinic sediments and sedimentary rocks: refining the paleoproxies. Chemical Geology 324–325, 1927.CrossRefGoogle Scholar
Sellwood, BW and Sladen, CP (1981) Mesozoic and Tertiary argillaceous units: distribution and composition. Quarterly Journal of Engineering Geology 14, 263275.CrossRefGoogle Scholar
Shalaby, SE (1980) The middle Liassic sedimentary rocks of the coastal zone of northwest Yorkshire : their petrology and sedimentation. Unpublished PhD thesis, Queen Mary, University of London, 420 pp.Google Scholar
Sheen, AI, Kendall, B, Reinhard, CT, Creaser, RA, Lyons, TW, Bekker, A, Poulton, SW and Anbar, AD (2018) A model for the oceanic mass balance of rhenium and implications for the extent of Proterozoic ocean anoxia. Geochimica et Cosmochimica Acta 227, 7595.CrossRefGoogle Scholar
Silva, RL, Duarte, LV, Comas-Rengifo, MJ, Mendonça Filho, JG and Azerêdo, AC (2011) Update of the carbon and oxygen isotopic records of the Early–Late Pliensbachian (Early Jurassic, ∼187 Ma): insights from the organic-rich hemipelagic series of the Lusitanian Basin (Portugal). Chemical Geology 283, 177184.CrossRefGoogle Scholar
Simms, MJ, Chidlaw, N, Morton, N and Page, KN (2004a) British Lower Jurassic Stratigraphy. The Geological Conservation Review Series 30, Peterborough: Joint Nature Conservation Committee, 458 pp.Google Scholar
Simms, MJ, Chidlaw, N, Page, KN and Morton, N (2004b) Chapter 6 The Cleveland Basin. In British Lower Jurassic Stratigraphy (ed Gallois, R). Geological Conservation Review Series 30, pp. 237304. Peterborough: Joint Nature Conservation Committee.Google Scholar
Singer, A (1984) The paleoclimatic interpretation of clay minerals in sediments – a review. Earth-Science Reviews 21, 251293.CrossRefGoogle Scholar
Sinninghe Damsté, JSS, Kenig, F, Koopmans, MP, Köster, J, Schouten, S, Hayes, JM and de Leeuw, JW (1995) Evidence for gammacerane as an indicator of water column stratification. Geochimica et Cosmochimica Acta 59, 18951900.CrossRefGoogle ScholarPubMed
Slater, SM, Bown, P, Twitchett, RJ, Danise, S and Vajda, V (2022) Global record of “ghost” nannofossils reveals plankton resilience to high CO2 and warming. Science 376, 853856.CrossRefGoogle ScholarPubMed
Slater, SM, Twitchett, RJ, Danise, S and Vajda, V (2019) Substantial vegetation response to Early Jurassic global warming with impacts on oceanic anoxia. Nature Geoscience 12, 462467.CrossRefGoogle Scholar
Slomp, CP and Van Cappellen, P (2007) The global marine phosphorus cycle: sensitivity to oceanic circulation. Biogeosciences 4, 155171.CrossRefGoogle Scholar
Song, J, Littke, R and Weniger, P (2017) Organic geochemistry of the Lower Toarcian Posidonia Shale in NW Europe. Organic Geochemistry 106, 7692.CrossRefGoogle Scholar
Song, J, Littke, R, Weniger, P, Ostertag-Henning, C and Nelskamp, S (2015) Shale oil potential and thermal maturity of the Lower Toarcian Posidonia Shale in NW Europe. International Journal of Coal Geology 150–151, 127153.CrossRefGoogle Scholar
Spirakis, CS (1996) The roles of organic matter in the formation of uranium deposits in sedimentary rocks. Ore Geology Reviews 11, 5369.CrossRefGoogle Scholar
Storm, MS, Hesselbo, SP, Jenkyns, HC, Ruhl, M, Ullmann, CV, Xu, W, Leng, MJ, Riding, JB and Gorbanenko, O (2020) Orbital pacing and secular evolution of the Early Jurassic carbon cycle. Proceedings of the National Academy of Sciences 117, 39743982.CrossRefGoogle ScholarPubMed
Suan, G, Mattioli, E, Pittet, B, Lecuyer, C, Sucheras-Marx, B, Duarte, LV, Philippe, M, Reggiani, L and Martineau, F (2010) Secular environmental precursors to Early Toarcian (Jurassic) extreme climate changes. Earth and Planetary Science Letters 290, 448458.CrossRefGoogle Scholar
Suan, G, Mattioli, E, Pittet, B, Mailliot, S and Lecuyer, C (2008a) Evidence for major environmental perturbation prior to and during the Toarcian (Early Jurassic) oceanic anoxic event from the Lusitanian Basin, Portugal. Paleoceanography 23, PA1202.CrossRefGoogle Scholar
Suan, G, Nikitenko, BL, Rogov, MA, Baudin, F, Spangenberg, JE, Knyazev, VG, Glinskikh, LA, Goryacheva, AA, Adatte, T, Riding, JB, Follmi, KB, Pittet, B, Mattioli, E and Lecuyer, C (2011) Polar record of Early Jurassic massive carbon injection. Earth and Planetary Science Letters 312, 102113.CrossRefGoogle Scholar
Suan, G, Pittet, B, Bour, I, Mattioli, E, Duarte, LV and Mailliot, S (2008b) Duration of the Early Toarcian carbon isotope excursion deduced from spectral analysis: consequence for its possible causes. Earth and Planetary Science Letters 267, 666679.CrossRefGoogle Scholar
Suan, G, Schlogl, J and Mattioli, E (2016) Bio- and chemostratigraphy of the Toarcian organic-rich deposits of some key successions of the Alpine Tethys. Newsletters on Stratigraphy 49, 401419.CrossRefGoogle Scholar
Suan, G, van de Schootbrugge, B, Adatte, T, Fiebig, J and Oschmann, W (2015) Calibrating the magnitude of the Toarcian carbon cycle perturbation. Paleoceanography 30, 495509.CrossRefGoogle Scholar
Svensen, H, Planke, S, Chevallier, L, Malthe-Sørenssen, A, Corfu, F and Jamtveit, B (2007) Hydrothermal venting of greenhouse gases triggering Early Jurassic global warming. Earth and Planetary Science Letters 256, 554566.CrossRefGoogle Scholar
Swanson, VE (1961) Geology and geochemistry of uranium in marine black shales A review. USGS Professional Paper 356-C, 67112.Google Scholar
Sweere, TC, Dickson, AJ, Jenkyns, HC, Porcelli, D, Ruhl, M, Murphy, MJ, Idiz, E, van der Boorn, SHJM, Eldrett, JS and Henderson, GM (2020) Controls on the Cd-isotope composition of Upper Cretaceous (Cenomanian–Turonian) organic-rich mudrocks from south Texas (Eagle Ford Group). Geochimica et Cosmochimica Acta 287, 251262.CrossRefGoogle Scholar
Sweere, TC, van den Boorn, S, Dickson, AJ and Reichart, GJ (2016) Definition of new trace-metal proxies for the controls on organic matter enrichment in marine sediments based on Mn, Co, Mo and Cd concentrations. Chemical Geology 441, 235245.CrossRefGoogle Scholar
Tate, R and Blake, JF (1876) The Yorkshire Lias. London: John Van Voorst, 475 pp.Google Scholar
Taylor, SR and McLennan, SM (2001) Chemical composition and element distribution in the Earth’s crust (3rd Edn). In Encyclopedia of Physical Science and Technology (ed Meyers, RA), pp. 697719. San Diego: Academic Press.Google Scholar
Them, TR II, Gill, BC, Caruthers, AH, Gerhardt, AM, Gröcke, DR, Lyons, TW, Marroquin, SM, Nielsen, SG, Alexandre, JPT and Owens, JD (2018) Thallium isotopes reveal protracted anoxia during the Toarcian (Early Jurassic) associated with volcanism, carbon burial, and mass extinction. Proceedings of the National Academy of Sciences of the United States of America 115, 65966601.CrossRefGoogle ScholarPubMed
Them, TR II, Gill, BC, Caruthers, AH, Gröcke, DR, Tulsky, ET, Martindale, RC, Poulton, TP and Smith, PL (2017a) High-resolution carbon isotope records of the Toarcian Oceanic Anoxic Event (Early Jurassic) from North America and implications for the global drivers of the Toarcian carbon cycle. Earth and Planetary Science Letters 459, 118126.CrossRefGoogle Scholar
Them, TR II, Gill, BC, Selby, D, Gröcke, DR, Friedman, RM and Owens, JD (2017b) Evidence for rapid weathering response to climatic warming during the Toarcian Oceanic Anoxic Event. Scientific Reports 7, 5003.CrossRefGoogle ScholarPubMed
Them, TR II, Jagoe, CH, Caruthers, AH, Gill, BC, Grasby, SE, Gröcke, DR, Yin, R and Owens, JD (2019) Terrestrial sources as the primary delivery mechanism of mercury to the oceans across the Toarcian Oceanic Anoxic Event (Early Jurassic). Earth and Planetary Science Letters 507, 6272.CrossRefGoogle Scholar
Them, TR II, Owens, JD, Marroquín, SM, Caruthers, AH, Trabucho-Alexandre, JP and Gill, BC (2022) Reduced marine molybdenum inventory related to enhanced organic carbon burial and an expansion of reducing environments in the Toarcian (Early Jurassic) oceans. AGU Advances 3, 120.CrossRefGoogle Scholar
Thibault, N, Ruhl, M, Ullmann, CV, Korte, C, Kemp, DB, Gröcke, DR and Hesselbo, SP (2018) The wider context of the Lower Jurassic Toarcian oceanic anoxic event in Yorkshire coastal outcrops, UK. Proceedings of the Geologists’ Association 129, 372391.CrossRefGoogle Scholar
Trabucho-Alexandre, J, Gröcke, DR, Atar, E, Herringshaw, L and Jarvis, I (2022) A new subsurface record of the Pliensbachian–Toarcian of Yorkshire. Proceedings of the Yorkshire Geological Society 64, 112.CrossRefGoogle Scholar
Trabucho-Alexandre, J (2015) More gaps than shale: Erosion of mud and its effect on preserved geochemical and palaeobiological signals. In Strata and Time: Probing the Gaps in Our Understanding (eds Smith, DG, Bailey, RJ, Burgess, PM and Fraser, AJ), pp. 251270. Geological Society London, Special Publications 404.Google Scholar
Trabucho-Alexandre, J, Tuenter, E, Henstra, GA, van der Zwan, KJ, van de Wal, RSW, Dijkstra, HA and de Boer, PL (2010) The mid-Cretaceous North Atlantic nutrient trap: black shales and OAEs. Paleoceanography 25, PA4201.Google Scholar
Trabucho-Alexandre, J, van Gilst, RI, Rodríguez-López, JP and de Boer, PL (2011) The sedimentary expression of oceanic anoxic event 1b in the North Atlantic. Sedimentology 58, 12171246.CrossRefGoogle Scholar
Tribovillard, N (2020) Arsenic in marine sediments: how robust a redox proxy? Palaeogeography, Palaeoclimatology, Palaeoecology 550, 114.CrossRefGoogle Scholar
Tribovillard, N (2021) Conjugated enrichments in arsenic and antimony in marine deposits used as paleoenvironmental proxies: preliminary results. BSGF Earth Sciences Bulletin 192, 39.CrossRefGoogle Scholar
Tribovillard, N, Algeo, TJ, Baudin, F and Riboulleau, A (2012) Analysis of marine environmental conditions based on molybdenum–uranium covariation – applications to Mesozoic paleoceanography. Chemical Geology 324, 4658.CrossRefGoogle Scholar
Tribovillard, N, Algeo, TJ, Lyons, T and Riboulleau, A (2006) Trace metals as paleoredox and paleoproductivity proxies: an update. Chemical Geology 232, 1232.CrossRefGoogle Scholar
Tucker, ME (2011) Sedimentary Rocks in the Field (4th Edn). The Geological Field Guide Series. Chichester: Wiley-Blackwell, 275 pp.Google Scholar
Turgeon, S and Brumsack, HJ (2006) Anoxic vs dysoxic events reflected in sediment geochemistry during the Cenomanian–Turonian Boundary Event (Cretaceous) in the Umbria–Marche Basin of central Italy. Chemical Geology 234, 321339.CrossRefGoogle Scholar
Tyson, RV (2005) The “productivity versus preservation” controversy: Cause, flaws, and resolution. In The Deposition of Organic-Carbon-Rich Sediments: Models, Mechanisms, and Consequences (ed Harris, NB). SEPM Special Publication 82, 1733.CrossRefGoogle Scholar
Tyson, RV (1995) Bulk geochemical characterization and classification of organic matter: stable carbon isotopes (δ13C). In Sedimentary Organic Matter. Organic Facies and Palynofacies (ed Tyson, RV). pp. 395416. Dordrecht: Springer.CrossRefGoogle Scholar
Ullmann, CV, Boyle, R, Duarte, L, Hesselbo, S, Kasemann, SA, Klein, T, Lenton, TM, Piazza, V and Aberhan, M (2020) Warm afterglow from the Toarcian Oceanic Anoxic Event drives the success of deep-adapted brachiopods. Scientific Reports 10, 6549.CrossRefGoogle ScholarPubMed
Ullmann, CV, Frei, R, Korte, C and Hesselbo, SP (2015) Chemical and isotopic architecture of the belemnite rostrum. Geochimica et Cosmochimica Acta 159, 231243.CrossRefGoogle Scholar
Ullmann, CV, Szȕcs, D, Jiang, M, Hudson, AJL and Hesselbo, SP (2022) Geochemistry of macrofossil, bulk rock and secondary calcite in the Early Jurassic strata of the Llanbedr (Mochras Farm) drill core, Cardigan Bay Basin, Wales, UK. Journal of the Geological Society 179, jgs20212018.CrossRefGoogle Scholar
Ullmann, CV, Thibault, N, Ruhl, M, Hesselbo, SP and Korte, C (2014) Effect of a Jurassic oceanic anoxic event on belemnite ecology and evolution. Proceedings of the National Academy of Sciences of the United States of America 111, 1007310076.CrossRefGoogle ScholarPubMed
Vaes, B, van Hinsbergen, DJJ, van de Lagemaat, SHA, van der Wiel, E, Lom, N, Advokaat, EL, Boschman, LM, Gallo, LC, Greve, A, Guilmette, C, Li, S, Lippert, PC, Montheil, L, Qayyum, A and Langereis, CG (2023) A global apparent polar wander path for the last 320 Ma calculated from site-level paleomagnetic data. Earth-Science Reviews 245, 104547.CrossRefGoogle Scholar
van Acken, D, Tütken, T, Daly, JS, Schmid-Röhl, A and Orr, PJ (2019) Rhenium–osmium geochronology of the Toarcian Posidonia Shale, SW Germany. Palaeogeography, Palaeoclimatology, Palaeoecology 534, 112.CrossRefGoogle Scholar
Van Breugel, Y, Baas, M, Schouten, S, Mattioli, E and Sinninghe Damsté, JSS (2006) Isorenieratane record in black shales from the Paris Basin, France: constraints on recycling of respired CO2 as a mechanism for negative carbon isotope shifts during the Toarcian oceanic anoxic event. Paleoceanography 21, 18.CrossRefGoogle Scholar
Van Buchem, FSP and Knox, RWOB (1998) Lower and Middle Liassic depositional sequences of Yorkshire (U.K.). In Mesozoic and Cenozoic Sequence Stratigraphy of European Basins (eds Graciansky, P-Cd, Hardenbol, J, Jacquin, T and Vail, PR). SEPM, Special Publication 60, 545559.Google Scholar
Van Cappellen, P and Ingall, ED (1994) Benthic phosphorus regeneration, net primary production, and ocean anoxia – a model of the coupled marine biogeochemical cycles of carbon and phosphorus. Paleoceanography 9, 677692.CrossRefGoogle Scholar
van de Schootbrugge, B, Houben, AJP, Ercan, FEZ, Verreussel, R, Kerstholt, S, Janssen, NMM, Nikitenko, B and Suan, G (2020) Enhanced Arctic–Tethys connectivity ended the Toarcian Oceanic Anoxic Event in NW Europe. Geological Magazine 157, 15931611.CrossRefGoogle Scholar
van de Schootbrugge, B, McArthur, JM, Bailey, TR, Rosenthal, Y, Wright, JD and Miller, KG (2005) Toarcian oceanic anoxic event: an assessment of global causes using belemnite C isotope records. Paleoceanography 20, PA3008.CrossRefGoogle Scholar
van Hinsbergen, DJJ, de Groot, LV, van Schaik, SJ, Spakman, W, Bijl, PK, Sluijs, A, Langereis, CG, Brinkhuis, H (2015) A paleolatitude calculator for paleoclimate studies. PLoS ONE 10, 121.CrossRefGoogle ScholarPubMed
van Hulten, M, Middag, R, Dutay, J-C, De Baar, H, Roy-Barman, M, Gehlen, M, Tagliabue, A and Sterl, A (2017) Manganese in the west Atlantic Ocean in the context of the first global ocean circulation model of manganese. Biogeosciences 14, 11231152.CrossRefGoogle Scholar
Vasseur, R, Lathuiliére, B, Lazăr, I, Martindale, RC, Bodin, S and Durlet, C (2021) Major coral extinctions during the early Toarcian global warming event. Global and Planetary Change 207, 103647.CrossRefGoogle Scholar
Vine, JD and Tourtelot, EB (1970) Geochemistry of black shale deposits – a summary report. Economic Geology 65, 253272.CrossRefGoogle Scholar
Visentin, S and Erba, E (2021) High-resolution calcareous nannofossil biostratigraphy across the Toarcian Oceanic Anoxic Event in northern Italy: clues from the Sogno and Gajum cores (Lombardy Basin, Southern Alps). Rivista Italiana di Paleontologia e Stratigrafia 127, 539556.Google Scholar
Von Eynatten, H, Barceló-Vidal, C and Pawlowsky-Glahn, V (2003) Modelling compositional change: the example of chemical weathering of granitoid rocks. Mathematical Geology 35, 231251.CrossRefGoogle Scholar
Wadley, MR, Stevens, DP, Jickells, TD, Hughes, C, Chance, R, Hepach, H, Tinerl, L and Carpenter, LJ (2020) A global model for iodine speciation in the upper ocean. Global Biogeochemical Cycles 34, e2019GB006467.CrossRefGoogle Scholar
Walker, RG and Plint, AG (1992) Wave- and storm-dominated shallow marine systems. In Facies Models: Response to Sea Level Change (eds Walker, RG and James, NP). GeoText 1, pp. 219238. St John’s, Newfoundland: Geological Association of Canada.Google Scholar
Wang, Y (2022) Anomalous weathering records in the Cleveland Basin (Yorkshire, UK) during the T-OAE global warming. Terra Nova 35, 153166.CrossRefGoogle Scholar
Wang, Y, Lu, W, Costa, KM and Nielsen, SG (2022) Beyond anoxia: exploring sedimentary thallium isotopes in paleo-redox reconstructions from a new core top collection. Geochimica et Cosmochimica Acta 333, 347361.CrossRefGoogle Scholar
Wang, Y, Ossa, FO, Spangenberg, JE and Schoenberg, R (2021) Restricted oxygen-deficient basins on the northern European epicontinental shelf across the Toarcian carbon isotope excursion interval. Paleoceanography and Paleoclimatology 36, e2020PA004207.CrossRefGoogle Scholar
Wang, Y, Ossa, FO, Wille, M, Schurr, S, Saussele, M-E, Schmid-Rohl, A and Schoenberg, R (2020) Evidence for local carbon-cycle perturbations superimposed on the Toarcian carbon isotope excursion. Geobiology 18, 682709.CrossRefGoogle ScholarPubMed
Ware, B, Jourdan, F and Timms, NE (2023) The Ferrar Continental Flood Basalt: a ∼1.6 Ma long duration evidenced by high-precision 40Ar/39Ar ages suggest a potential role in the Pliensbachian–Toarcian extinction event. Earth and Planetary Science Letters 622, 118369.CrossRefGoogle Scholar
Whitehead, TH, Anderson, W, Wilson, V and Wray, DA (1952) Liassic iron ores of the Cleveland district. In The Liassic Ironstones. Memoirs of the Geological Survey of Great Britain. The Mesozoic Ironstones of England. pp. 3567. London: HMSO.Google Scholar
Wignall, PB and Bond, DPG (2008) The end-Triassic and Early Jurassic mass extinction records in the British Isles. Proceedings of the Geologists’ Association 119, 7384.CrossRefGoogle Scholar
Wignall, PB and Newton, R (1998) Pyrite framboid diameter as a measure of oxygen deficiency in ancient mudrocks. American Journal of Science 298, 537552.CrossRefGoogle Scholar
Wignall, PB, Newton, RJ and Little, CTS (2005) The timing of paleoenvironmental change and cause-and-effect relationships during the Early Jurassic mass extinction in Europe. American Journal of Science 305, 10141032.CrossRefGoogle Scholar
Wilkin, RT, Barnes, HL and Brantley, SL (1996) The size distribution of framboidal pyrite in modern sediments: an indicator of redox conditions. Geochimica et Cosmochimica Acta 60, 38973912.CrossRefGoogle Scholar
Wilkinson, GM, Besterman, A, Buelo, C, Gephart, J and Pace, ML (2018) A synthesis of modern organic carbon accumulation rates in coastal and aquatic inland ecosystems. Scientific Reports 8, 15736.CrossRefGoogle ScholarPubMed
Woodfine, RG, Jenkyns, HC, Sarti, M, Baroncini, F and Violante, C (2008) The response of two Tethyan carbonate platforms to the early Toarcian (Jurassic) oceanic anoxic event: environmental change and differential subsidence. Sedimentology 55, 10111028.CrossRefGoogle Scholar
Worden, RH, Utley, JEP, Butcher, AR, Griffiths, J, Wooldridge, LJ and Lawan, AY (2020) Improved imaging and analysis of chlorite in reservoirs and modern day analogues: New insights for reservoir quality and provenance. In Application of Analytical Techniques to Petroleum Systems (eds Dowey, P, Osborne, M and Volk, H). Geological Society London, Special Publications 484, 189204.Google Scholar
Wright, JK (2022) The Market Weighton High in the 21st century – new understanding of a long-standing problem. Proceedings of the Yorkshire Geological Society 64, pygs20212008.CrossRefGoogle Scholar
Wu, F, Owens, JD, Huang, T, Sarafian, A, Huang, K-F, Sen, IS, Horner, TJ, Blusztajn, J, Mprton, P and Nielsen, SG (2019) Vanadium isotope composition of seawater. Geochimica et Cosmochimica Acta 244, 403415.CrossRefGoogle Scholar
Xia, G and Mansour, A (2022) Paleoenvironmental changes during the early Toarcian Oceanic Anoxic Event: insights into organic carbon distribution and controlling mechanisms in the eastern Tethys. Journal of Asian Earth Sciences 237, 105344.CrossRefGoogle Scholar
Xu, WM, Ruhl, M, Jenkyns, HC, Hesselbo, SP, Riding, JB, Selby, D, Naafs, BDA, Weijers, JH, Pancost, RD, Tegelaar, E and Idiz, EF (2017) Carbon sequestration in an expanded lake system during the Toarcian oceanic anoxic event. Nature Geoscience 10, 129134.CrossRefGoogle Scholar
Xu, WM, Ruhl, M, Jenkyns, HC, Leng, MJ, Huggett, JM, Minisini, D, Ullmann, CV, Riding, JB, Weijers, JWH, Storm, MS, Percival, LME, Tosca, NJ, Idiz, EF, Tegelaar, EW and Hesselbo, SP (2018) Evolution of the Toarcian (Early Jurassic) carbon-cycle and global climatic controls on local sedimentary processes (Cardigan Bay Basin, UK). Earth and Planetary Science Letters 484, 396411.CrossRefGoogle Scholar
Xu, WM, Weijers, JWH, Ruhl, M, Idiz, EF, Jenkyns, HC, Riding, JB, Gorbanenko, O and Hesselbo, SP (2021) Molecular and petrographical evidence for lacustrine environmental and biotic change in the palaeo-Sichuan mega-lake (China) during the Toarcian Oceanic Anoxic Event. In Carbon Cycle and Ecosystem Response to the Jenkyns Event in the Early Toarcian (Jurassic) (eds Reolid, M, Duarte, LV, Mattioli, E and Ruebsam, W). Geological Society London, Special Publications 514, 335357.Google Scholar
Yan, Q, Li, X, Kemp, DB, Guo, J, Zhang, Z and Hu, Y (2023) Elevated atmospheric CO2 drove an increase in tropical cyclone intensity during the early Toarcian hyperthermal. PNAS 120, e2301018120.CrossRefGoogle ScholarPubMed
Yang, T, Shen, Y, Qin, YJ, Jin, JS, Zhang, Y, Tong, G and Liu, J (2021) Distribution of radioactive elements (Th, U) and formation mechanism of the bottom of the Lopingian (Late Permian) coal-bearing series in western Guizhou, SW China. Journal of Petroleum Science and Engineering 205, 108779.CrossRefGoogle Scholar
Yano, M, Yasukawa, K, Nakamura, K, Ikehara, M and Kato, Y (2020) Geochemical features of redox-sensitive trace metals in sediments under oxygen-depleted marine environments. Minerals 10, 120.CrossRefGoogle Scholar
Young, GAM and Bird, J (1822) A Geological Survey of the Yorkshire Coast: Describing the Strata and Fossils occurring between the Humber and the Tees, from the German Ocean to the Plain of York. Whitby: George Clark, 332 pp.Google Scholar
Young, TP, Aggett, JR and Howard, AS (1990) The Cleveland Ironstone Formation. In Jurassic and Ordovician Ooidal Ironstones (ed Young, TP). pp. 131. Nottingham: British Sedimentological Research Group.Google Scholar
Zakharov, VA, Shurygin, BN, Il’ina, VI and Nikitenko, BL (2006) Pliensbachian–Toarcian biotic turnover in north Siberia and the Arctic region. Stratigraphy and Geological Correlation 14, 399417.CrossRefGoogle Scholar
Zhai, R, Zeng, Z, Zhang, R and Yao, W (2023) The response of nitrogen and sulfur cycles to ocean deoxygenation across the Cenomanian–Turonian boundary. Global and Planetary Change 227, 104182.CrossRefGoogle Scholar
Zhang, F, Lenton, TM, del Rey, A, Romaniello, SJ, Chen, X, Planavsky, NJ, Clarkson, MO, Dahl, TW, Lau, KV, Wang, W, Li, Z, Zhao, M, Isson, T, Algeo, TJ and Anbar, AD (2020) Uranium isotopes in marine carbonates as a global ocean paleoredox proxy: a critical review. Geochimica et Cosmochimica Acta 287, 2749.CrossRefGoogle Scholar
Zhang, R, Kemp, DB, Thibault, N, Jelby, ME, Li, M, Huang, C, Sui, Y, Wang, Z, Liu, D and Jia, S (2023) Astrochronology and sedimentary noise modeling of Pliensbachian (Early Jurassic) sea-level changes, Paris Basin, France. Earth and Planetary Science Letters 614, 118199.CrossRefGoogle Scholar
Zhou, XL, Jenkyns, HC, Owens, JD, Junium, CK, Zheng, XY, Sageman, BB, Hardisty, DS, Lyons, TW, Ridgwell, A and Lu, ZL (2015) Upper ocean oxygenation dynamics from I/Ca ratios during the Cenomanian–Turonian OAE 2. Paleoceanography 30, 510526.CrossRefGoogle Scholar
Zhu, Y, La Croix, A, Kemp, DB, Shen, J, Huang, C, Hua, X, Li, Y and Wei, M (2024) Are sulfides the primary host of sedimentary Hg? A case study from the Lower Jurassic of the Surat Basin (Australia). Chemical Geology 652, 122028.CrossRefGoogle Scholar
Figure 0

Figure 1. Early Jurassic palaeogeography, regional setting and location of the Dove’s Nest study core in the Cleveland Basin. (a) Palaeogeographic reconstruction of Europe showing the location of the basin on the European epicontinental shelf; interpreted bottom-water redox conditions associated with the T-OAE are based on geological data and ocean circulation modelling (Ruvalcaba Baroni et al., 2018). (b) Global palaeogeography of the Early Jurassic showing continent configuration, major ocean basins and location of the Karoo–Ferrar Large Igneous Provinces (LIPs) that were emplaced during the early – middle Toarcian (Heimdal et al., 2021; Gaynor et al., 2022). Yellow box shows the location of the Europe map. Palaeogeographic base maps in (a) and (b) modified from Blakey (2012, 2016); palaeolatitude in (a) revised based on the online palaeolatitude calculator of van Hinsbergen et al. (2015) at 183 Ma (https://paleolatitude.org) with the palaeomagnetic reference frame of Vaes et al. (2023). (c) Map of eastern North Yorkshire showing the geographic distribution of Jurassic sediments in the Cleveland Basin, isopachs for the Lias and location of the Dove’s Nest borehole. Redrawn after Kent (1980) and Rawson & Wright (2000).

Figure 1

Figure 2. Stratigraphic log of the studied section of the Dove’s Nest core and correlation to a composite outcrop section along the North Yorkshire coast between Hawsker Bottoms and Port Mulgrave. Organic carbon isotopes (δ13Corg) and whole-rock total organic carbon (TOCWR) profiles are shown with their correlation (modified from Trabucho-Alexandre et al., 2022). Organic-rich facies associated with the large negative δ13Corg excursion of the Toarcian Oceanic Anoxic Event (T-OAE; Jenkyns, 1985) together with the negative δ13Corg excursion defining the Pliensbachian – Toarcian Boundary Event (Littler et al., 2010) provide prominent tie points. The δ13Corg maximum of the A. gibbosus Subzone is equated to the Late Pliensbachian Event positive excursion of Korte & Hesselbo (2011), De Lena et al. (2019) and Hollaar et al. (2023). Grain size scale: fm, fine mudstone; mm, medium mudstone; cm, coarse mudstone; and fs, very fine sandstone. Yorkshire coast ‘bed’ numbers and named marker beds from Hawsker Bottoms (Fig. 1) for the Pliensbachian (Howarth, 1955) and Whitby composite section for the Toarcian (Howarth, 1962, 1973, 1992). Dove’s Nest data from Trabucho-Alexandre et al. (2022) and this study. Yorkshire coast δ13Corg data from Hawsker Bottoms: orange, Littler et al. (2010); turquoise, Cohen et al. (2004); and pink, DB Kemp et al. (2005). Port Mulgrave: dark blue, DB Kemp et al. (2005); turquoise, Cohen et al. (2004). Saltwick Bay: turquoise, Cohen et al. (2004). TOC data for the Yorkshire coast are composite section values from Kemp et al. (2011; thin green high-resolution curve), Ruvalcaba Baroni et al. (2018; green-filled triangles) and McArthur (2019; green-filled circles). Thick green line shows the trend of the two low-resolution datasets. Vertical dotted lines and numbers are the δ13Corg reference value for average Phanerozoic black shale (Meyers, 2014) and the TOCWR content of average shale (Law, 1999) and average black shale (Vine & Tourtelot, 1970). SB2 and SB3 are the middle and upper Sulphur Bands of the basal lower Toarcian (Salem, 2013; McArthur, 2019). Abbreviations of biostratigraphic zonation: H. bifrons = Hildoceras bifrons; H. serpentinum = Harpoceras serpentinum; D. tenui. = Dactylioceras tenuicostatum; P. spin. = Pleuroceras spinatum; A. margaritatus = Amaltheus margaritatus; D. commune = Dactylioceras commune; H. falciferum = Harpoceras falciferum; C. exa. = Cleviceras exaratum; Ds = Dactylioceras semicelatum; * = Dactylioceras tenuicostatum; † = Dactylioceras clevelandicum; Pp = Protogrammoceras paltum; Ph = Pleuroceras hawskerense; Pa = Pleuroceras apyrenum; A. gib. = Amaltheus gibbosus; As = Amaltheus subnodosus; A. stokesi = Amaltheus stokesi. Ages after GTS2020 (Gradstein et al., 2020) with revisions of Al-Suwaidi et al. (2022). Chemostratigraphic units modified from Remírez & Algeo (2020) and defined by multi-element proxies (see text); note that a – d, to the left of the TOCWR profile for the Yorkshire coast, are subunits of Unit III, the T-OAE.

Figure 2

Figure 3. Carbon isotope correlation of selected European Pliensbachian – Toarcian successions. The map (bottom right) shows the palaeogeographic location of the sites (see Fig. 1 for details). Cleveland Basin δ13Corg profiles from this study (Dove’s Nest = black, coast composite = grey; see Fig. 2 for sources). The top of the D. commune Subzone lies ∼16 m above the top of the Hard Shales on the Yorkshire coast (Hesselbo & Jenkyns, 1995). Mochras δ13Corg data from Xu et al. (2018) and Storm et al. (2020); δ13Ccarb after Ullmann et al. (2022). CIEs as Figure 2 with Stokesi Event of Peti et al. (2017) and Storm et al. (2020). Sancerre δ13Corg data from Hermoso et al. (2013) and Peti et al. (2021); δ13Ccarb after Hermoso et al. (2009a, 2009b; 2013) and Peti et al. (2021). Pliensbachian biostratigraphy follows Peti et al. (2017, 2021) and Zhang et al. (2023). Peniche δ13Corg profile from Fantasia et al. (2019). Peniche Pliensbachian δ13Ccarb values after Oliveira et al. (2006) with stratigraphic revisions and additional data from Silva et al. (2011); Pliensbachian – Toarcian boundary and Toarcian δ13Ccarb after Hesselbo et al. (2007). Yorkshire stratigraphy follows Figure 2. Other abbreviations: PlToBE = Pliensbachian – Toarcian Boundary Event; H. falcif. = Harpoceras falciferum; Dc = Dactylioceras commune; Pf = Peronoceras fibulatum; Cc = Catacoeloceras crissum; D. ten. = Dactylioceras tenuicostatum; Ast. = Amaltheus stokesi; Ps = Pleuroceras spinatum; Dp = Dactylioceras polymorphum; H. levisoni = Hildaites levisoni.

Figure 3

Figure 4. Geochemical profiles for lithofacies proxies Al2O3 (aluminosilicates, principally clay minerals), CaCO3e (carbonates; calcite, siderite) and TOCWR (organic fraction) through the upper Pliensbachian – middle Toarcian of the Dove’s Nest core, with selected detrital proxies. ‘Bed’ numbers, names and biostratigraphy are derived from chemostratigraphic correlation to Hawsker Bottoms for the Pliensbachian (Howarth, 1955) and a Whitby composite section for the Toarcian (Howarth, 1962, 1973, 1992): red, sideritic beds; blue, limestones; and black other beds (see Fig. 2). Vertical dotted lines and numbers are reference values for Post-Archean Average Shale (PASS; = average mud of Taylor & McLennan, 2001). Prominent limestone ‘bed’ 35 (Whale Stones) and ‘beds’ 39 – 40 (Top Jet Dogger and Millstones) are clearly expressed by their high CaCO3e contents. Significant shifts in the elemental (Figs S1, S2) and element-ratio profiles (Figs 4, 5) combined with coincident changes in δ13Corg and TOC (Fig. 2) are used to define the chemostratigraphic units (see text), modified from the scheme of Remírez & Algeo (2020). LPlE = Late Pliensbachian Event; other abbreviations as Figure 2. Detrital proxies show multiple stacked CU cycles superimposed on a longer-term fining-upward trend through the top Staithes Sandstone to mid-Cleveland Ironstone, followed by a marked upward increase in grain size comprising 3 stacked CU cycles (cf. Macquaker & Taylor, 1996). Cycle 3 is most prominent and coincides with an interval of high δ13Corg and TOCWR values ascribed to the LPlE. The base of the Whitby Mudstone is a sharp facies break to clay-mineral dominated sediments illustrated by a steep rise in K/Al, Rb/Al and Cs/Al (not plotted) ratios. The T-OAE is expressed by a sharp increase and peak in TOCWR (maximum 9.2%), although elevated organic matter contents continue upward through the Whitby Mudstone. Upward-coarsening, 405 ka cycles in the Jet Rock (Si/Al profile) derived from cycle analysis of the coastal section by Thibault et al. (2018, fig. 2) are also displayed in the Dove’s Nest record.

Figure 4

Figure 5. Geochemical profiles for redox and productivity proxies through the upper Pliensbachian – middle Toarcian of the Dove’s Nest core. Stratigraphic framework as in Figure 4. SB2 – SB3 are the middle and upper Sulphur Bands, consisting of laminated pyritic carbonaceous mudstones.

Figure 5

Figure 6. Field photograph of the lower cliff face immediately west of the old harbour of Port Mulgrave (Fig. 1) annotated with the TOC, CaCO3 and δ13Corg data of DB Kemp et al. (2005) and Kemp et al. (2011) and the stratigraphic framework of Howarth (1962). The five distinctive concretionary horizons in the Jet Rock, extinction level iii of Caswell et al. (2009) and chemostratigraphic units (IIIa – IVa) with key intervals of change are indicated. D. semi. = Dactylioceras semicelatum; C. exaratum = Cleviceras exaratum; H.f. = H. falciferum. Chemostratigraphic correlation to Dove’s Nest is illustrated in Figure 7. Note that the shore platform in the foreground occurs at the level of the base Jet Rock (‘bed’ 33). The sedimentary log and geochemical profiles below this (‘bed’ 32) represent variations in the subsurface at this site.

Figure 6

Figure 7. Chemostratigraphic correlation of the T-OAE interval in the Yorkshire coastal outcrop reference sections with the Dove’s Nest core. Stratigraphy after Howarth (1962, 1973, 1992) and Howarth (in Cope et al., 1980). Dt = Dactylioceras tenuicostatum Subzone. Lithological log is based on the Hawsker Bottoms and Port Mulgrave sections from DB Kemp et al. (2005): lithologies are dark-grey laminated mudrocks (dark-grey shading), medium-grey mudrocks (pale-grey shading) and carbonate bands and nodules (brick pattern). Major carbonate markers – ‘Stone’ bands and ‘Doggers’ – are indicated. Sample heights of Hesselbo et al. (2000, fig. 3) were recalculated based on the positions of major bed contacts. Data sources: Hesselbo et al. (2000); DB Kemp et al. (2005); Kemp et al. (2011); Thibault et al. (2018); Trabucho-Alexandre et al. (2022); this study. Si/Al and Ti/Al ratios for the coastal sections were recalculated by Thibault et al. (2018 supplementary data) after correction for analytical bias. Shaded intervals a – d represent subdivisions of chemostratigraphic Unit III. This unit corresponds to the interval displaying the large negative carbon-isotope excursion that characterizes the T-OAE. Dashed horizontal grey lines show the correlation of major bed bases; dotted horizontal grey lines correlate significant chemostratigraphic tie points. Cyclostratigraphic filtered output for carbon isotopes (orange curve) and the detrital fraction (yellow curve, derived from Zr/Rb data) are plotted after Thibault et al. (2018). Vertical dotted lines indicate: the δ13Corg value of average Phanerozoic black shales (grey; Meyers, 2014); the oxic–anoxic- and anoxic–euxinic-facies boundaries defined by TOC content (green, 2.5% and 10%) proposed by Algeo & Maynard (2004).

Figure 7

Figure 8. Box plot summary of stratigraphic trends in major- and trace-element contents and Al-ratio data of chemostratigraphic units through the upper Pliensbachian – middle Toarcian of the Dove’s Nest core. Boxes represent 25–75% quartiles with median values shown by the vertical line inside the box. Whiskers are drawn from the top of the box up to the largest data point less than 1.5 times the box height from the box (the ‘upper inner fence’) and similarly below the box (Hammer et al., 2001). Outlier values outside the inner fences are shown as circles, values further than 3 times the box height from the box (the ‘outer fences’) are shown as stars.

Figure 8

Figure 9. Principal component analysis biplot of PC1 vs PC2 for geochemical data from the upper Pliensbachian – middle Toarcian of the Dove’s Nest core. Compositional data for samples (n = 96) having a full major- and trace-element dataset were transformed using a Centre Log-Ratio to remove closure effects prior to PCA. The first two principal components account for 48.4% and 22.9% of the variance, respectively (Table S4).

Figure 9

Figure 10. Ti/Al and K/Al ratio profiles from Dove’s Nest compared to quartz grain size, proportion of silt, illite content, CIA, and Os isotopes. Quartz grain size and illite content (peak-area integration with height ratios method) after de Vos (2017). Silt percentage was determined from Port Mulgrave (Ghadeer, 2011). CIA = Chemical Index of Alteration (Nesbitt & Young, 1982); CIA = [AI2O3/(Al2O3 + CaO* + Na2O + K2O)] × 100 (molecular proportions). CaO* moles assumed to be equivalent to Na2O (filled blue circles) with additional values derived from CaO determinations (open blue circles), where a number of moles was less than that of Na2O (McLennan, 1993). Osmium isotope plot from Cohen et al. (2004, fig. 1), with composite data from 3 coastal sections: Hawsker Bottoms; Port Mulgrave; Saltwick Bay. Cleveland Basin relative sea-level curve (Hesselbo, 2008) replotted relative to biostratigraphic zones interpreted for the Dove’s Nest core. 187Os/188Os ratio of early Toarcian ocean water (0.377 ± 0.065) after van Acken et al. (2019). Cycles are based principally on the Ti/Al profile.

Figure 10

Figure 11. Geochemical cross-plots for selected detrital proxies and grain size in the Dove’s Nest core. (a) Si/ Al vs Ti/Al. (b) Na/Al vs K/Al. (c) Ti/Al vs quartz mean grain size (de Vos, 2017). Mineral reference compositions from webmineral.com, average shale (PAAS) after Taylor & McLennan (2001). Plotted regression lines are (a) ordinary least square and (b, c) reduced major axis, with 95% confidence envelope in (c). Grey shading in (b) represents the field of the Subunit IIIb illite and weathering pulse including Whale Stones ‘bed’ 35, characterized by anomalous low CIA values (Fig. 10).

Figure 11

Figure 12. Ternary diagram of Al2O3–(CaO*+Na2O)–K2O in Dove’s Nest rock samples. Values are molecular proportions. (a) CIA = Chemical Index of Alteration (Nesbitt & Young, 1982) with CaO* moles assumed to be equivalent to Na2O (McLennan, 1993; see text). Tonalite, granodiorite and granite compositions after Condie (1993). (b) Enlargement of plotted data in the top sector of the diagram (red outline in a) with selected sample details. Grey shading represents the field of the Subunit IIIb illite and weathering pulse with high K/Al and anomalous low CIA values.

Figure 12

Figure 13. Ternary plot of iron and carbonate-associated elements Fe+Mn–Ca–Mg in Dove’s Nest rock samples compared to constituent mineral compositions. Stars are electron microprobe determinations of mineral fractions from the Cleveland Ironstone of Staithes (Aggett, 1990). Open triangles are calculated pure mineral values (webmineral.com). Average shale (PASS) composition after Taylor & McLennan (2001). Plot is scaled based on the maximum and minimum values of the three components (cf. de Lange et al., 1987).

Figure 13

Figure 14. Correlation of TOC/PT, DOPT, FeEF, MnEF and PEF between the Dove’s Nest core and Yorkshire coastal outcrop sections. Coast geochemical profiles from McArthur et al. (2008), McArthur (2019) and Remírez & Algeo (2020) with additional high-resolution data (Thibault et al., 2018); Yorkshire chemostratigraphic Units I – V modified from Remírez & Algeo (2020). Stratigraphy as in Figs 2, 4. WS = Whale Stones (‘bed’ 35’); TJD = Top Jet Dogger (‘bed’ 39). Enrichment factors (EF) are calculated relative to PASS. Data of Thibault et al. (2018) are recalibrated relative to McArthur et al. (2008) and McArthur (2019): stratigraphic heights are increased by 1.1 m; Al values are increased by 20% to remove analytical bias. Vertical grey dotted lines are EF values of 1. Vertical green dotted lines indicate the Redfield ratio, a TOC/P ratio of ∼106:1, typical of marine plankton biomass (Redfield et al., 1963). Values of >106 indicate P-release from the sediment under reducing conditions. Vertical dotted lines on the TOC/PT and DOPT plots mark the positions of redox boundaries typically associated with values of 50 and 0.25 (red, oxic/suboxic), 106 and 0.5 (green, dysoxic/anoxic) respectively, following Algeo & Ingall (2007) and Algeo & Maynard (2004). More conservative threshold DOP values of <0.45 for oxic or dysoxic depositional environments and >0.75 for a euxinic environment (gold vertical dotted line) have been proposed by Raiswell et al. (2018).

Figure 14

Figure 15. Bottom-water redox proxy interpretation for the upper Pliensbachian – lower Toarcian of the Cleveland Basin derived from TOC, TOC/P, DOP and Fe speciation. δ13Corg profile for Dove’s Nest (Trabucho-Alexandre et al., 2022, black) rescaled to match coast composite data (see Fig. 2 for sources). Rescaled whole-rock TOC profile for Dove’s Nest (Trabucho-Alexandre et al., 2022, dark green) with coast composite data of Kemp et al. (2011; thin yellow-green high-resolution curve) and trend of the low-resolution coast datasets of Ruvalcaba Baroni et al. (2018, open triangles) and McArthur (2019) (thin pale green low-resolution curve; see Fig. 2). Average shale and black shale values as in Figure 2; ‘anoxic threshold’ of TOCWR = 2.5 wt% follows Algeo & Maynard (2004). Low-resolution TOC/PT (dark green) and DOPT (dark orange) curves from McArthur et al. (2008) and Remírez & Algeo (2020) with high-resolution data (thin pale curves) from Thibault et al. (2018); see Figure 14 for further information. Iron speciation data from Salem (2013, cream-filled circles) and Houben et al. (2021, yellow-filled circles) with redox field boundaries after Raiswell et al. (2018). Extinction levels (i)– (iii) after Caswell et al. (2009). Pliensb. = Pliensbachian; Bitumin. Sh. = Bituminous Shales; D. semic. = Dactylioceras semicelatum; Dt = D. tenuicostatum; Dc = D. clevelandicum; P. hawk. = Pleuroceras hawskerense; Pa = P. apyrenum.

Figure 15

Figure 16. Correlation of Mo, TOC, V and U between the Dove’s Nest core and Yorkshire coastal outcrop sections. Dove’s Nest data this study. Yorkshire coast Mo and Mo/TOC and TOC plots from McArthur (2019) with additional high-resolution TOC (thin black line, see Fig. 2), Mo and Mo/TOC curves (thin dark red lines; Thibault et al., 2018). Yorkshire VEF data were calculated from Ruvalcaba Baroni et al. (2018) with MoEF, U and UEF data after Remírez & Algeo (2020). Stratigraphy as in Figures 2, 4. WS = Whale Stones (‘bed’ 35’); TJD = Top Jet Dogger (‘bed’ 39). Enrichment factors (Section 6.d), e.g. MoEF, are calculated relative to PASS. Vertical grey dotted lines are EF values of 1. Vertical green dotted lines indicate the ‘intermittent euxinia’ boundary of 25 ppm Mo (Scott & Lyons, 2012) and the ‘anoxic threshold’ of TOC = 2.5 wt% (Algeo & Maynard, 2004). Consistent enrichment in authigenic uranium (UEF >1) characterizing Units III and IV is also well displayed in spectral gamma-ray logs of the coastal sections (Myers & Wignall, 1987; Parkinson, 1996).

Figure 16

Figure 17. Cross-plots for key redox-sensitive trace metals. (a) Mo vs Al. Mo shows no clear relationship with Al. (b) Mo vs TOC. The steep upper regression line (small grey dots) for Units IV and V (left) derived from the Dove’s Nest data, displaying a positive correlation between Mo and TOC, contrasts to the shallow lines (small grey dots) derived for T-OAE Unit III (right – upper line and statistics is for Dove’s Nest samples, lower line is for Yorkshire coast samples of McArthur et al. 2008). Regression lines from selected modern anoxic silled basins representing increasing deep water renewal times of <10 – 650 ka (Algeo & Rowe, 2012) are Mo/TOC (ppm/%) ∼45 Saanich Inlet (purple); ∼25 Cariaco Basin (green); ∼9 Framvaren Fjord (red); ∼4.5 Black Sea (blue). Bottom water restriction trends after Algeo & Lyons (2006). (c) U vs Al. Lower regression line is for Pliensbachian Subunits Ia–c; upper line is for Toarcian Units IV and V (all samples ≥3 ppm U). (d) U vs TOC. Lower regression line (left) for Subunits Ia–c; upper regression line (right) for Units IV and V. (e) V vs Al. Regression line is for Pliensbachian Subunits Ia–c, excluding the three Fe-rich flyers. (f) V vs TOC. Regression lines are for Pliensbachian Subunits Ia–c (left, Dove’s Nest) and Toarcian Units II–V (right, coast samples; Ruvalcaba Baroni et al., 2018). TOC-based anoxic (2.5%, vertical dashed green line marking boundary between oxic and anoxic non-sulfidic conditions) and euxinic (10%, upper limit of x-axis) thresholds after Algeo & Maynard (2004). Solid symbols are from the Dove’s Nest core (this study), faded symbols are for Yorkshire coastal outcrop samples (McArthur et al., 2008; Ruvalcaba Baroni et al., 2018; McArthur, 2019; Remírez & Algeo, 2020). Dove’s Nest TOC data are whole-rock values. Average shale (PASS) composition after Taylor & McLennan (2001). Note that samples with 10 – 20% TOC reported from Unit III of the coastal outcrops fall outside the plot area of (B), (D) and (F) but lie on the trends of the regression lines shown.

Figure 17

Figure 18. MoEF vs UEF cross-plot for stratigraphic units comprising the upper Pliensbachian – middle Toarcian of Dove’s Nest core. The three diagonal lines represent multiples (0.3, 1, 3) of the Mo:U ratio of present-day seawater (SW) converted to an average weight ratio of 3.1 for the purpose of comparison with sediment Mo:U weight ratios (Tribovillard et al., 2012). General patterns of MoEF vs UEF covariation in modern marine environments modified from Tribovillard et al. (2012) and Yano et al. (2020): unrestricted open ocean field based on the eastern tropical Pacific; particulate shuttle field based on the Cariaco Basin and Saanich Inlet. Trend lines show deoxygenation trends in modern marine environments (Tribovillard et al., 2012) with positions based on data from restricted basins and coastal settings (Paul et al., 2023). The anomalous high EF values of Whale Stones sample 167.89, interpreted as sampling a carbonate concretion, are likely an artifact of the high carbonate content (81%).

Figure 18

Figure 19. Stratigraphic variation in selected isotope geochemistry in the upper Pliensbachian – middle Toarcian of Yorkshire. δ13Corg profile for Dove’s Nest (Trabucho-Alexandre et al., 2022, black) rescaled to coastal succession, with compiled high-resolution coast curve (grey, see Fig. 2; Cohen et al., 2004; DB Kemp et al., 2005; Littler et al., 2010). Stratigraphic framework as in Figure 2. Rescaled whole-rock TOC profile for Dove’s Nest (Trabucho-Alexandre et al., 2022, dark green) with coast composite data of Kemp et al. (2011; thin yellow-green high-resolution curve), McArthur (2019; thin pale green low-resolution curve) and Ruvalcaba Baroni et al. (2018; green-filled triangles). Yorkshire coast Mo profile of McArthur (2019; thick pink line) with elemental results for Mo-isotope samples of Pearce et al. (2008; red filled triangles) and high-resolution data of Thibault et al. (2018; thin dark red line, see Fig. 16). δ98/95Mo coast profile from Pearce et al. (2008). Details of the Mo concentration and isotope curves within Unit III (T-OAE) are presented in Figure 20. Belemnite carbonate-associated sulfur isotope profile (δ34SCAS) from Gill et al. (2011; white filled squares) incorporating the data of Newton et al. (2011; orange squares). Belemnite 87Sr/86Sr coast curve after McArthur et al. (2000). 187Os/188Osi profile of Cohen et al. (1999; 2004). Belemnite carbonate oxygen-isotope (δ18O, blue dots) and Mg/Ca ratios (blue circles) after McArthur et al. (2000).

Figure 19

Figure 20. Stratigraphic profiles of δ13Corg, TOC and selected trace-metal isotopes within Unit III, the T-OAE interval of the Yorkshire coast. Stratigraphy as in Figures 6, 7. Bulk rock δ13Corg and TOC profiles of the composite section from Hesselbo et al. (2000, pale coloured lines) and DB Kemp et al. (2005, dark lines), Kemp et al. (2011, dark lines) – see Figure 7. Shaded grey bands indicate Unit III Subunits a – c (see Section 8.c); shaded blue band is the interval of the carbonate maximum, Subunit IIId. A – D mark coincident sharp falls in δ13Corg (after DB Kemp et al., 2005; Cohen et al., 2007) and δ98Mo, with increased Mo, as noted by Kemp et al. (2011). Biotic extinction levels ii and iii after Caswell et al. (2009); base of trace fossil absent interval follows Caswell & Herringshaw (2023). Thallium isotopes (ε205Tl) from Nielsen et al. (2011). Rhenium, Mo, Re/Mo and δ98Mo profiles from Pearce et al. (2008). Note that Nielsen et al.’s (2011, fig. 5) comparison figure of ε205Tl vs δ98Mo at Port Mulgrave incorrectly plotted the position of the Mo dataset relative to the stratigraphy, as presented by Pearce et al. (2008, fig. 2). The replotted data in our figure do not support an anti-correlation between these two isotope systems, as proposed by Nielsen et al. (2011) and modelled by Owens et al. (2017).

Figure 20

Figure 21. Stratigraphic profiles of δ13Corg, δ13Cn-alkanes, isorenieratane, AOM, TOC and δ15Ntot within Unit III, the T-OAE interval of the Yorkshire coast. Bulk rock δ13Corg and TOC profiles of the composite section from Hesselbo et al. (2000, pale coloured lines) and DB Kemp et al. (2005, dark lines), Kemp et al. (2011, dark lines) – see Figure 7; ‘anoxic threshold’ of TOCWR = 2.5 wt% follows Algeo & Maynard (2004). δ13C data for terrestrial wood (Hesselbo et al., 2000, brown squares) are offset but track the bulk sediment δ13Corg curve. Carbon isotope values from Hawsker Bottoms of representative long-chain n-alkane biomarkers (δ13Cn-alkane) derived from terrestrial plants (n-C27, n-C29) also display the negative excursion of the T-OAE (French et al., 2014). Short-chain n-alkanes (n-C17n-C19) attributed to marine plants follow an identical δ13C trend (French et al., 2014, fig. 7). The isorenieratane profile, a biomarker for anaerobic phototrophic green sulfur bacteria, provides evidence of reducing conditions developing during the initial phase of the T-OAE, peaking at the time of deposition of ‘bed’ 40 (Millstones); chlorobactane and okenane (not shown) display identical patterns (French et al., 2014, fig. 5). Amorphous organic matter (AOM) and algal cysts dominate (50 – >90%) the palynological assemblages (after Slater et al., 2019) of the high-TOC anoxic – euxinic facies. Peak species richness of calcareous nannofossils, preserved as carbonate and external moulds in organic matter throughout the section (Slater et al., 2022), occurs in Whale Stones ‘bed’ 35, together with a pulse of prasinophyte algae and dense granular organic matter (Houben et al., 2021). Nitrogen isotopes (δ15Ntot) display low values attributable to enhanced N2 fixation by cyanobacteria in a strongly redox-stratified marine environment (Wang et al., 2021). Toarcian open ocean seawater field derived from Tethyan sections (Section 16.a.1). The δ15Ntot profile (Jenkyns et al., 2001) broadly follows TOC. A – D mark coincident sharp falls in δ13Corg (after DB Kemp et al., 2005; Cohen et al., 2007) and δ98Mo, as noted by Kemp et al. (2011) – see Figure 20. Biotic extinction levels ii and iii after Caswell et al. (2009); base of trace fossil absent interval follows Caswell & Herringshaw (2023).

Figure 21

Figure 22. TOC, biotic trends, extinction levels and palaeoredox change through the upper Pliensbachian – middle Toarcian of the Cleveland Basin. Yorkshire coast stratigraphy as in Figure 2. Rescaled whole-rock TOC profile for Dove’s Nest (Trabucho-Alexandre et al., 2022, dark green) with coast composite data of Kemp et al. (2011; thin yellow-green high-resolution curve) and trend of the low-resolution coast datasets of Ruvalcaba Baroni et al. (2018) and McArthur (2019) (thin pale green low-resolution curve; see Fig. 2). Ranges of low-oxygen specialist bivalve taxa compiled from Little (1995) and Caswell et al. (2009). Trace-fossil taxonomic richness from Caswell & Frid (2017), Caswell & Dawn (2019) and Caswell & Herringshaw (2023). Macrofaunal diversity after Danise et al. (2013,2015). Profiles rescaled to reference stratigraphy (Fig. 2) using subzone and marker bed datum levels. Extinction levels (i) – (iii) from Caswell et al. (2009). The boundary between pre-extinction and post-extinction survival intervals of Atkinson et al. (2023) lies at the base of the C. exaratum Subzone (level iii). The top of the ‘survival interval’, the base of recovery phase 1, occurs in the lower H. bifrons Zone above the top of our study interval. Geochemical palaeoredox interpretations from this study (see text). Climate interpretation incorporates palynological interpretation of Slater et al. (2019) with oxygen isotope and Mg/Ca trends from belemnites (Fig. 21).

Supplementary material: File

Jarvis et al. supplementary material 1

Jarvis et al. supplementary material
Download Jarvis et al. supplementary material 1(File)
File 4 MB
Supplementary material: File

Jarvis et al. supplementary material 2

Jarvis et al. supplementary material
Download Jarvis et al. supplementary material 2(File)
File 235.8 KB
Supplementary material: File

Jarvis et al. supplementary material 3

Jarvis et al. supplementary material
Download Jarvis et al. supplementary material 3(File)
File 34 KB