114
Manuscript version: Accepted Manuscript This is a PDF of an unedited manuscript that has been accepted for publication. The manuscript will undergo copyediting, typesetting and correction before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain. Although reasonable efforts have been made to obtain all necessary permissions from third parties to include their copyrighted content within this article, their full citation and copyright line may not be present in this Accepted Manuscript version. Before using any content from this article, please refer to the Version of Record once published for full citation and copyright details, as permissions may be required. Accepted Manuscript Proceedings of the Yorkshire Geological Society Cretaceous Oceanic Anoxic Event 2 in eastern England: further palynological and geochemical data from Melton Ross Paul Dodsworth, James S. Eldrett & Malcolm B. Hart DOI: https://doi.org/10.1144/pygs2019-017 Received 30 October 2019 Revised 20 May 2020 Accepted 20 May 2020 © 2020 The Author(s). This is an Open Access article distributed under the terms of the Creative Commons Attribution 4.0 License (http://creativecommons.org/licenses/by/4.0/). Published by The Geological Society of London for the Yorkshire Geological Society. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics Supplementary material at https://doi.org/10.6084/m9.figshare.c.4987205 When citing this article please include the DOI provided above. by guest on August 28, 2020 http://pygs.lyellcollection.org/ Downloaded from

Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Manuscript version: Accepted Manuscript This is a PDF of an unedited manuscript that has been accepted for publication. The manuscript will undergo copyediting,

typesetting and correction before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Although reasonable efforts have been made to obtain all necessary permissions from third parties to include their

copyrighted content within this article, their full citation and copyright line may not be present in this Accepted Manuscript version. Before using any content from this article, please refer to the Version of Record once published for full citation and

copyright details, as permissions may be required.

Accepted Manuscript

Proceedings of the Yorkshire Geological

Society

Cretaceous Oceanic Anoxic Event 2 in eastern England:

further palynological and geochemical data from Melton Ross

Paul Dodsworth, James S. Eldrett & Malcolm B. Hart

DOI: https://doi.org/10.1144/pygs2019-017

Received 30 October 2019

Revised 20 May 2020

Accepted 20 May 2020

© 2020 The Author(s). This is an Open Access article distributed under the terms of the Creative

Commons Attribution 4.0 License (http://creativecommons.org/licenses/by/4.0/). Published by The Geological Society of London for the Yorkshire Geological Society. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

Supplementary material at https://doi.org/10.6084/m9.figshare.c.4987205

When citing this article please include the DOI provided above.

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 2: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Cretaceous Oceanic Anoxic Event 2 in eastern England: further palynological and geochemical

data from Melton Ross

Paul Dodsworth1, James S. Eldrett2 & Malcolm B. Hart3

1StrataSolve Ltd, 15 Francis Road, Stockton Heath, Warrington WA4 6EB, U.K.

2Shell International Exploration & Production B.V., Lange Kleiweg 40, 2288 GK Rijswijk,

Netherlands.

3School of Geography, Earth & Environmental Sciences, Plymouth University, Drake Circus,

Plymouth PL4 8AA, U.K.

Correspondence: Paul Dodsworth ([email protected])

Abstract. The lowermost 1.45 m of the Welton Chalk Formation, including the regional sedimentary

record of Oceanic Anoxic Event 2 (OAE-2), has been sampled at Melton Ross Quarry in eastern

England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first

time, while a detailed palynological study follows previously published preliminary results. It

comprises a condensed interval that spans the Cenomanian–Turonian Stage boundary. A locally

preserved, lower ‗anomalous‘ succession (Beds I–VII) and a ‗Central Limestone‘ (Bed A) are shown

to correlate respectively with the pre-Plenus sequence and Plenus Bed at Misburg and Wunstorf in the

Lower Saxony Basin (LSB), NW Germany. They are overlain by a succession of variegated marls

(Bed B to Bed H), including the Black Band (Beds C–E), that can be correlated across eastern

England. Based on a carbon isotope (δ13C) profile and dinoflagellate cyst and acritarch bio-event

correlation, Beds B–H appear to be a highly attenuated post-Plenus equivalent of the LSB succession,

including part of the ‗Fish Shale‘. The δ13C profile shows possible ‗precursor‘ / ‗build-up‘ events in

the lower succession at Melton Ross, with the main OAE-2 δ13C excursion occurring in the Central

Limestone and overlying Beds B–H. The darker coloured marls from the Black Band and Bed G

contain 1.43–3.47% total organic carbon (TOC), hydrogen index values of 78–203 mg HC/g TOC and

oxygen index values of 15–26 mg CO2/g TOC, indicating type III and type II–III organic matter, of

mixed terrigenous and marine algal sources. The corresponding palynological assemblages are

dominated by marine dinoflagellate cysts, comprising mainly gonyaulacoid taxa, with subordinate

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 3: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

terrigenous miospores, mainly gymnosperm bisaccate pollen, consistent with a distal marine setting.

The interbedded lighter-coloured marls contain less than 0.4% TOC and lower proportions of

miospores and peridinioid dinoflagellate cysts compared with the darker layers. This is suggestive of

moderately raised levels of productivity during deposition of the darker layers, possibly related to

greater nutrient availability from land-derived sources. The occurrence of the peridinioid taxa

Eurydinium saxoniense and Bosedinia spp., together with higher proportions of prasinophyte

phycomata in the darker layers, may also point to stimulation of organic-walled phytoplankton

productivity by reduced nitrogen chemo-species encroaching the photic zone, possibly by expansion

of an oxygen-minimum zone. Exceptionally high concentrations of palynomorphs (in the tens of

thousands to lower hundreds of thousands per gramme range) in the darker layers at Melton Ross and

eight other eastern England localities is consistent with increased quality of sea floor preservation in a

low oxygen environment, coupled with a high degree of stratigraphic condensation. Two new

dinoflagellate cyst species are described from Melton Ross, Canninginopsis? lindseyensis sp. nov. and

Trithyrodinium maculatum sp. nov., along with two taxa described in open nomenclature.

Supplementary material: One pdf file, with detailed sample positions and descriptions, tables of

supporting information (also available in Excel format), quarry photographs and a palynological

distribution chart, is available at

The Late Cretaceous Epoch was characterized by sustained warm climate, resulting in high eustatic

sea levels (Miller et al. 2005; Haq & Huber 2016; Hay et al. 2018). Numerous epicontinental seaways

became established, submerging large areas of Western Europe (e.g. Gale 1995) and the Western

Interior of North America (Kauffman & Caldwell 1993). Major global perturbations in the carbon

cycle occurred, termed Oceanic Anoxic Events, the most prominent spanning the Cenomanian–

Turonian Stage boundary (CTB, 93.9 Ma; International Commission on Stratigraphy 2019), named

OAE-2 (Schlanger & Jenkins 1976; Schlanger et al. 1987) and lasting for up to ca. 900 kyr (Eldrett et

al. 2015a; Gangl et al. 2019). This interval was marked by a globally recognized positive carbon

isotope (δ13C) excursion, reflecting the widespread sequestration of δ12C-enriched organic matter in

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 4: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

marine sediments under global anoxic conditions (Jenkyns 2010 and references therein). However, the

deposition of dark-coloured, organic-rich, fine-grained sediments (‗black shales‘) varied both

temporally and spatially, being modulated and ultimately dependent on local and regional processes

(basin restriction, water stratification, bottom-currents, sediment input) in addition to global

phenomena (Large Igneous Province activity, sea level change, orbital forcing; e.g. Ernst & Youbi

2017; Clarkson et al. 2018; Minisini et al. 2018). A global mass extinction / turnover bio-event

occurred around the CTB and is probably associated with OAE-2 (Raup & Sepkoski 1982, 1984;

Kauffman 1984; Milne et al. 1985; Hart 2005, 2019).

In the North Sea Basin and adjacent outcrop areas in NW Germany and eastern England, deposition of

dark coloured mudstones interbedded with light coloured mudstones and limestones commenced

during the Late Cenomanian, and continued into the Early Turonian in some areas, including

Wunstorf and Misburg in the Lower Saxony Basin, NW Germany (Fig. 1). At these locations, the

CTB succession is 12–30 m thick (Ernst et al. 1983; Hilbrecht 1986). The dark mudstone layers form

a distinctive interval in predominantly whitish limestone chalk successions. In eastern England (Fig.

2), a highly condensed (relative to NW Germany) ‗Black Band‘ is developed in Yorkshire (0.3–0.7 m

thick at Flixton, East Knapton, Bishop Wilton and Market Weighton; Jeans et al. 1991; Dodsworth

1996) and northern Lincolnshire (0.1–0.2 m thick at South Ferriby, Elsham, Melton Ross, Bigby and

Caistor). Jenkyns (1985) envisaged a pelagic shelf depositional setting for the Black Band and

adjacent strata in a relatively shallow (several hundred metres) epicontinental sea. The Black Band

wedges out to the south of Louth (Hart et al. 1991). It appears to represent the ‗feather-edge‘ of OAE-

2, which dies out when traced toward a palaeo-high that appears to have been located in the region of

the Wash (Hart et al. 1991). Schlanger et al. (1987) suggested that it may have been deposited near

the upper limit of an oxygen-minimum zone that lapped onto the continental shelf.

The Cenomanian and Turonian stages probably record the highest sea levels within the Cretaceous

Period, with maximum eustatic sea levels being reached during the Early Turonian (Haq et al. 1988;

Sahagian et al. 1996; Miller et al. 2003, 2005). They were probably the time of minimal continental

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 5: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

relief during the Phanerozoic, and the time of minimal detrital sediment delivery to the ocean (Hay et

al. 2018). The CTB interval was also a time of extreme warmth with mid-latitude sea-surface

temperature possibly exceeding 35°C (Huber et al. 2002; Voigt et al. 2004; Forster et al. 2007;

Robinson et al. 2019). This warmth was likely associated with a more stratified water column,

resulting in poor atmosphere–ocean gas exchange, oxygen-deficient photic zone waters, including

euxinic conditions in some regions, and enhanced organic matter preservation at the sediment–water

interface (Sinninghe Damsté & Köster 1998; Monteiro et al. 2012). Rhythmic bedding in CTB

successions has been linked to orbital-forcing (e.g. Arthur et al. 1986; Eldrett et al. 2015b; Boulila et

al. 2020), which in some regions, including NW Germany, may have caused an intensified

hydrological cycle during warmer/wetter periods (van Helmond et al. 2015; Charbonnier et al. 2018;

Gharaie & Kalanat 2018).

The early stages of OAE-2 were characterised by a relatively extended interval (ca. 150–200 kyr;

Clarkson et al. 2018) of ca. 3–5°C cooling of sea surface temperatures in the proto-Atlantic and the

European shelf (Forster et al. 2007; Pearce et al. 2009; Sinninghe Damsté et al. 2010; Jarvis et al.

2011; van Helmond et al. 2014, 2015). Boreal Realm fauna, including the belemnite Praeactinocamax

plenus (Blainville), migrated southwards across Europe (Jefferies 1962, 1963; Gale & Christensen

1996; Marcinowski et al. 1996; Košták et al. 2004). The interval was termed the ‗Plenus Cold Event‘

(PCE) by Gale & Christensen (1996) and probably marks a shift towards improved oxygenation of

bottom waters during OAE-2 (Forster et al. 2007; Eldrett et al. 2014, 2017; van Helmond et al. 2014),

possibly on a global scale (Clarkson et al. 2018; O‘Connor et al. 2020), and relatively drier climates

(Heimhofer et al. 2018; Gharaie & Kalanat 2018). Across regions of Europe that record organic-rich

sedimentation during OAE-2, including NW Germany and SE France (e.g. Pont d'Issole, Fig. 1),

‗black shale‘ deposition temporarily gave way in some areas to more oxygenated sediments during the

PCE, and a resumption of deposition of relatively thick biogenic pelagic limestones (Wiese et al.

2009; Jarvis et al. 2011; Grosheny et al. 2017; Jenkyns et al. 2017). In NW Germany, a conspicuous

tripartite limestone bed, called the Plenus Bed or Plenus Bank because it yields the eponymous

belemnite, is developed at this level (Ernst et al. 1984; Hilbrecht & Dahmer 1994), separating

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 6: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

underlying pre-Plenus and overlying post-Plenus successions of dark coloured mudstones interbedded

with light coloured mudstones and limestones.

Wood & Mortimore (1995) and Wood et al. (1997) reported the CTB succession from Melton Ross

Quarry in Lincolnshire (Fig. 2). Temporary deep excavations in the late 1990s (Fig. 3) exposed a

section hitherto unreported in eastern England, presumed eroded by unconformity elsewhere, that

appears to correlate with the lowermost 2–3 m of the NW Germany CTB succession, including the

pre-Plenus interval and the Plenus Bed. These deposits are overlain by the regionally correlative Bed

B, Black Band and Beds F–H (Fig. 4).

In a recent review of the Black Band, Hart (2019) discussed advances in foraminiferal knowledge a

quarter of a century after previous publications on their distribution across the CTB in eastern

England (Hart & Bigg 1981; Hart et al. 1993). The present paper follows up on this work, reviewing

advances in palaeoenvironmental knowledge and biostratigraphic dating using organic-walled

phytoplankton since this was last discussed for eastern England in the publications of Dodsworth

(1996, 2000) and Wood et al. (1997). Previous analysis of palynological recovery from the Black

Band at South Ferriby and Flixton (Hart et al. 1993; Dodsworth 1996) has revealed exceptionally high

concentrations of dinoflagellate cysts, in the thousands to lower hundreds of thousands per gramme

range. Hart & Koutsoukos (2015) recommended further investigation of whether the abundance of

dinoflagellate cysts is a function of their increased productivity under eutrophic conditions, increased

quality of sea floor preservation in a low oxygen environment, or a normal level of organic

productivity accentuated by a loss of biogenic carbonate sediment (including planktonic foraminifera

and calcareous nannofossils) during the CTB mass extinction (Jarvis et al. 1988a; Lamolda et al.

1994; Paul & Mitchell 1994; Hart 1996). This paper provides the first detailed palynological data

from Melton Ross and attempts to discriminate the relative contribution of these factors. A taxonomic

review of new and problematic dinoflagellate cysts is included. Other marine palynomorphs are

documented, including prasinophyte phycomata and acanthomorph acritarchs, along with land-

derived (terrigenous) pollen and spores. For regional comparison, summary palynological data are

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 7: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

published for eight correlative eastern England localities. From south to north, these are Caistor,

Bigby, South Ferriby, Market Weighton, Bishop Wilton, East Knapton, Flixton and Speeton (Fig. 2).

Total organic carbon, Rock-Eval pyrolysis and stable isotope data are presented for the first time from

Melton Ross. The integrated bio- and chemostratigraphy of the section is assessed in an inter-regional

context.

In this paper, chronostratigraphic substages and their corresponding ages are treated as formal units,

using initial upper case letters, consistent with the usual practice, although not all have been ratified.

The base of the Cenomanian Stage and Lower Cenomanian Substage have a ratified Global boundary

Stratotype Section and Point (GSSP) at Mont Risou in southern France (Kennedy et al. 2004). The

base of the Turonian Stage and Lower Turonian Substage have a ratified GSSP near Pueblo, Colorado

(Fig. 1; Kennedy et al. 2005). The base of a Middle Cenomanian Substage has a proposed GSSP in

Sussex, southern England (Tröger et al. 1996), and that of a Middle Turonian Substage is proposed

near Pueblo (Bengtson et al. 1996; Kennedy et al. 2000; Dodsworth & Eldrett 2019). Tröger et al.

(1996) recommended southern France as a suitable region for an Upper Cenomanian Substage GSSP,

though they did not propose a site. Replacement of Acanthoceras ammonites by the genus

Calycoceras (Hancock 1991), which can also be correlated using carbon isotopes (Kennedy & Gale

2006), provides a possible datum for the base of an Upper Cenomanian Substage.

Lithostratigraphy

The Late Cretaceous Epoch in England is represented by the bio-micritic, white limestones of the

Chalk Group (Fig. 2), which Wood & Smith (1978) grouped into three major faunal and depositional

provinces: a Northern Province (investigated here) which links eastern England to the north of the

Wash with contemporaneous North Sea and NW German successions; a Southern Province linking

southern England and northern France (Anglo-Paris Basin); and a Transitional Province in the

Chilterns and East Anglia. In eastern England, chalk bedding is developed on a decimetre scale with

more clay-rich marls forming thin (<5 cm) inter-beds (e.g. Hancock 1976; Jeans 1980). From Louth

northward in Lincolnshire and Yorkshire, the lowermost ca. 0.5 to 1.5 m of the Welton Chalk

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 8: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Formation (Wood & Smith 1978) is characterised by an atypically thick succession of variegated

marls, including the organic-rich Black Band (Fig. 4).

Hopson (2005) gave a comprehensive review of the historical naming of the lowermost part of the

Welton Chalk Formation, along with evidence for its stratigraphic position, based on

chemostratigraphy (stable carbon isotopes: Schlanger et al. 1987; Hart et al. 1991; Gale et al. 1993;

Wood & Mortimore 1995; and subsequently, Clarkson et al. 2018) and macrofossils (e.g. Jefferies

1963; Whitham 1991; Gaunt et al. 1992; Wood et al. 1997). Hopson (2005) proposed using the term

Plenus Marl Member for marl and limestone beds between the base of the Welton Chalk Formation

and the base of the Black Band, on the grounds of their probable correlation with the Plenus Marls of

the Anglo-Paris Basin. He also proposed using the term Black Band Member for marls and limestones

between the base of the Black Band and the base of the Buckton Member (Mitchell 2000), on the

grounds of these deposits probably being stratigraphically higher than the Plenus Marls, equivalent to

part of the Melbourn Rock Member / Ballard Cliff Member in southern England. However,

subsequent workers have retained one unit, the Flixton Member of Jeans (1980), for both intervals

(Hart 2019; Mitchell 2019). In Table 1, the subdivision of the Flixton Member into two sub-members

(Brett et al. 2018) is suggested, based on the definitions of Hopson (2005).

In the Anglo-Paris Basin, Jefferies (1963) labelled the eight correlative beds of a ‗standard‘ Plenus

Marls succession as Beds 1–8 in ascending order. Any strata that could not be correlated with these

beds were labelled locally with lower case Roman numerals. Jefferies (1963) could not trace Beds 1–8

farther north than Marham in Norfolk (Fig. 2), though subsequent work (Voigt et al. 2006) tentatively

identified the uppermost units (Bed 7 and Bed 8) at Heacham and Barret Ringstead, near Hunstanton.

Jefferies (1963) investigated one locality from Lincolnshire (South Ferriby) and another from

Yorkshire (Speeton), with the Black Band assigned to ‗bed i‘ at Speeton and ‗bed iii‘ at South Ferriby.

Hart et al. (1991, 1993) applied Jefferies‘ (1963) informal bed names at South Ferriby. Subsequently,

individual beds within the Flixton Member have been assigned different names by different authors

during research published in the 1990s (Table 1). Most of these schemes use ascending numbers

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 9: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

comparable to the Anglo-Paris Basin succession whilst not intending to imply lithostratigraphic

correlation with southern England beds of the same number. To avoid confusion, Dodsworth (1996)

alternatively proposed using letters rather than numbers for the eastern England succession, and

erected a regional lithostratigraphic scheme of comparable resolution to previous schemes.

In hydrocarbon wells from the central North Sea Basin, e.g. well 47/10-1 (Fig. 2), the distinctive

black, grey and green lithologies described from the CTB interval (Rhys 1974; Burnhill & Ramsay

1981) are comparable to those from the lowermost part of the Welton Chalk Formation of onshore

eastern England, and contain comparable foraminiferal assemblages (Burnhill & Ramsay 1981;

Crittenden et al. 1991) and palynological assemblages (Marshall & Batten 1988; Dodsworth 1996).

The term ‗Plenus Marl Formation‘ was initially applied to the corresponding high gamma log-

response unit in U.K. sector North Sea wells by Deegan & Scull (1977), but this has since been

replaced by ‗Black Band Bed‘ (Johnson & Lott 1993; Surlyk et al. 2003; van der Molen & Wong

2007) to reflect lithologies and stratigraphy different from that of the Plenus Marls in the Anglo-Paris

Basin. Farther north, in the Norwegian sector, e.g. well 35/6-2 S (Fig. 1), the term Blodøks Formation

(of Isaksen & Tonstad 1989) is applied to the correlative interval developed within the siliciclastic

Shetland Group. The Blodøks Formation is usually a few metres thick and rarely exceeds 20 m

(Gradstein & Waters 2016).

In this paper, the lithostratigraphy of Dodsworth (1996) is retained (Fig. 4). Higher resolution

subdivision around the Central Limestone (of Wood et al. 1997; Bed A), as described by Wood et al.

(1997) and Hildreth (1999) at Melton Ross Site 2 and some other northern Lincolnshire quarries,

including Bigby, are treated here as subdivisions of Beds A and B (Table 1). In Yorkshire, Beds A

and B may pinch out at some localities, e.g. Bishop Wilton and East Knapton (Dodsworth 1996), and

Speeton (Jefferies 1963; Mitchell et al. 1996; Mitchell 2000), with the Black Band at the base of the

Welton Chalk Formation or a short distance above it, i.e. above an attenuated, 2–12 cm thick Bed B.

The local Beds I–VII of Wood et al. (1997) are applied here to the lower succession at Melton Ross.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 10: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Previous research

Wood & Mortimore (1995) made an initial description of a relatively expanded stratigraphic

succession at Melton Ross Quarry (Site 1; Fig. 3), which exposed a section of marls, hitherto

unreported in eastern England, above the base of the Welton Chalk Formation. The marls are

presumed to have been eroded by unconformity elsewhere. Wiese et al. (2009) suggested this lower

‗anomalous‘ section may have been fortuitously preserved in a down-faulted block, though there is no

regional geophysical evidence of such a structure.

Wood et al. (1997) described three further Melton Ross excavations made during 1995–1996 (sites 2–

4; Fig. 3), which contain a thicker lower succession than Site 1, and categorised the localised deposits

into seven units, Beds I–VII. They expressed uncertainty about how Beds III–VI correlate with Site 1.

At the time of writing (2019), sites 1–4 are buried approximately 10 m below the restored surface of

the quarry.

In addition, Wood et al. (1997) collected their own set of 27 samples from Site 2 (labelled CJW-1 to -

26) and presented Rare Earth Element data, analysis of clay mineralogy and preliminary palynological

comments from those samples. They reported shale-normalised Rare Earth Element profiles to be sub-

horizontal with no Europium anomaly, suggesting to them a detrital rather than volcanogenic

provenance for the CTB succession (Wood et al. 1997, fig. 5). They further reported the clay

mineralogy to comprise a mixture of illite, kaolinite and smectite at the base of the succession, but

becoming predominantly smectite upwards with the loss of kaolinite in the regionally correlative

upper succession (Wood et al. 1997, fig. 4). They interpreted this change in mineralogy as probably

suggesting either deepening or increasing distance from shorelines through the CTB interval (Wood et

al. 1997, p. 339).

Wood et al. (1997, p. 342, fig. 3) recovered Praeactinocamax plenus from Site 1 at Melton Ross, in

silty lithologies immediately above the Central Limestone (Fig. 4). They considered the latter to

correlate with the Plenus Bed in NW Germany. Wood & Mortimore (1995) and Wood et al. (1997)

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 11: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

did not publish logs of the chalk section above the variegated marls (Bed B to Bed H) at Melton Ross,

i.e. from the upper part of the Flixton Member. This section is logged in the present study (Fig. 4) and

reveals a thickness of chalk (0.65 m) between the top of the variegated marls and ‗Adrian‘s pair of

marls‘ (of Mortimore 2014), immediately below the base of the Buckton Member, comparable to that

at nearby South Ferriby and Caistor (Hart et al. 1991, fig. 2; Mortimore 2014, fig. 4.23a; Hart 2019,

fig. 4).

Previous palynological work

There have been several previous palynological investigations of the lowermost part of the Welton

Chalk Formation. R.J. Davey, in Hart & Bigg (1981), reported dinoflagellate cysts to be the dominant

palynomorphs at Elsham (Fig. 2), with subordinate bisaccate pollen and very rare spores. Hart & Bigg

(1981) suggested that marine algae were probably the main source of the abundant organic matter in

the Black Band. They also noted that adjacent marls and chalks at Elsham were palynologically

barren, probably a function of unfavourable lithologies for the preservation of organic-walled

microfossils. A species list and distribution chart were not provided. Marshall (1983, p.172–174) and

Marshall & Batten (1988) undertook detailed analysis of six South Ferriby samples. Cyclonephelium

compactum–membraniphorum was reported to be the dominant dinoflagellate cyst in the Black Band,

with sparse recovery of palynomorphs from marls immediately below and above (Beds B and F), and

palynologically barren samples from adjacent chalks. Microplankton compose 76% of an assemblage

from the lower part of the Black Band (Bed C) and 89% from the upper part (Bed E). The

Cenomanian marker taxon Litosphaeridium siphoniphorum was not recorded in any of the samples.

Duane (1992, p. 277–292; in Hart et al. 1993) provided palynological data for four productive

samples from the east wall at South Ferriby quarry (Beds C to G); adjacent chalks and marls from ca.

75 cm above and 65 cm below were palynologically barren. Cyclonephelium compactum–

membraniphorum, along with Eurydinium saxoniense, are dominant in the dinoflagellate cyst

assemblages of the productive samples, while terrigenous bisaccate pollen and spores compose < 5%

of the total palynological assemblages. Exceptionally high concentrations of dinoflagellate cysts

(thousands per gramme), relative to Middle Cenomanian chalks and marls of southern England

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 12: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

(hundreds per gramme; Paul et al. 1994), were reported. Duane (1992; in Hart et al. 1993) did not

record Litosphaeridium siphoniphorum in any of the South Ferriby samples.

Dodsworth (1996) published palynological data from three samples of the Black Band from the south

wall at South Ferriby quarry (recovery from adjacent marl samples was sparse), six samples from the

Black Band at Flixton (splits of the same samples analysed for geochemistry by Jeans et al. 1991),

two samples from Bed G at Market Weighton, and commented on the palynology of the Black Band

at Caistor, Bigby, Market Weighton, Bishop Wilton, East Knapton and Speeton. As with previous

investigations, Litosphaeridium siphoniphorum was not recorded in any of the samples. However,

taxa whose last occurrences Marshall & Batten (1988) calibrated approximately to an influx of the

latest Cenomanian zonal ammonite Neocardioceras juddii Barrois & Guerne in NW Germany

(Hilbrecht 1986), namely Adnatosphaeridium tutulosum (in Bed C at Flixton) and Carpodinium

obliquicostatum (in Bed C at South Ferriby, and Beds C and E at Flixton), were recorded and

illustrated. Batten, in Wood et al. (1997), studied two samples from the Black Band at Melton Ross

(one each from Bed C and Bed E) and confirmed an absence of L. siphoniphorum. However, analysis

of five samples from local Beds II–VI at Site 2 revealed the first records of this taxon in the Northern

Province, where it is common in occurrence. Dodsworth (2000, fig. 12), provided preliminary marker

taxa distribution data from the samples documented more fully here (Fig. 4), corroborating Batten‘s

record of common L. siphoniphorum in local Beds I–?III at Melton Ross (Site 1), and indicating the

presence of A. tutulosum and C. obliquicostatum higher in the succession (Site 4), from Beds C–F and

Bed C, respectively.

Material and methods

Twenty-seven channelled (composite) rock samples (MR97 series) were collected from the top of the

Ferriby Chalk Formation and lowermost 1.45 m of the Welton Chalk Formation in Melton Ross

Quarry by one of us (PD) on 24th February 1997 and are documented in the present paper. Eleven of

our samples (MR97-23 to -14) were collected from Site 1 of Wood & Mortimore (1995). Sites 2 and 3

and the lower part of Site 4 of Wood et al. (1997) were submerged below the quarry‘s water table in

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 13: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

February 1997 and were unavailable for sampling. The upper part of Site 4 was accessible and 16 of

our samples (MR97-13 to -1) were collected from the relatively freshly-exposed, regionally

correlative beds there. The thickness of channelled samples varies from 3 mm to 230 mm, depending

on the thickness of rock layers (Fig. 4; see the Supplementary Appendix for details of the samples).

The samples were prepared and analysed for organic geochemistry, palynology, and carbonate stable

isotopes. In the following text, the sample series prefix is not written out in full for each sample; thus,

sample MR97-1 is shortened to sample -1.

Organic geochemistry

All samples were analysed for Total Organic Carbon (TOC). Samples with > 0.25% TOC (-1 to -11.5,

and -22) were also analysed for Rock-Eval pyrolysis. Analyses were undertaken in the laboratory of

Applied Petroleum Technology (APT) AS, Oslo, Norway. All procedures followed Weiss et al.

(2000). For TOC, a Leco SC-632 instrument was used. Diluted hydrochloric acid (HCl) was added to

the crushed rock samples to remove carbonate. The samples were then introduced into the Leco

combustion oven, and the amount of carbon in the sample was measured as carbon dioxide (CO2) by

an IR-detector. For Rock-Eval pyrolysis, a HAWK instrument was used. Jet-Rock 1 was run as every

tenth sample and checked against the acceptable range given in Weiss et al. (2000). The temperature

programme was a five minutes purge before pyrolysis: 300°C (three minutes) plus 25°C per minute

until 650°C was reached.

Palynological processing

Five grammes (dark coloured lithologies) or ten grammes (light coloured lithologies) of crushed,

dried material from samples -1 to -22 were dissolved in hydrochloric acid (35% HCl) and

hydrofluoric acid (40% HF) in order to remove carbonate and silicate minerals, respectively.

Preparations were sieved with a 10 μm mesh. Kerogen slides were prepared at this stage.

Palynomorphs (illustrated in Figs. 5–7), brown and black wood fragments (vitrinite and inertinite

phytoclasts) and clumps of granular amorphous organic matter (AOM) tend to occur in comparable

proportions in the > 10 μm kerogen from the < 1% TOC samples at Melton Ross (Table 2; Fig. 8). To

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 14: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

improve remaining residues for the counting of palynomorphs, many of these preparations were

cleaned by treatment with a ‗nitric wash‘, i.e. one minute of oxidation with nitric acid (70% HNO3),

or a rinse with potassium hydroxide solution (2% KOH). In the > 10 μm kerogen fraction from the >

1% TOC samples, dark coloured, clumped AOM is dominant (Table 2; Fig. 8). To liberate

palynomorphs from the AOM, extended oxidation was given with: 1, nitric acid (18–24 hours),

followed by one to two minutes in an ultrasonic bath with a 2% KOH solution that was supersaturated

with potassium permanganate (KMnO4; samples -8, -8.5, -10, -11, -11.5); 2, Schulze‘s solution (nitric

acid supersaturated with potassium chlorate, KClO3), followed by one subsequent rinse with a 2%

KOH solution (sample -2). The former technique was found to give better results than the latter in

terms of palynomorph preservation and in not selectively destroying gonyaulacoid dinoflagellate cysts

(cf. Dodsworth 1995; 2004a, fig. 4; Dodsworth & Eldrett 2019). All oxidised preparations were

stained with Safranin O solution (red stain). Full laboratory processing records are available in

Supplementary Table A.

Where palynological recovery permitted, approximately equal portions of quantified organic residues

from each sample were strewn over four 22 x 22 mm cover slips, dried, and mounted onto microscope

slides using Norland Optical Adhesive 61. The proportion of organic residue strewn on each cover

slip (e.g. 10% of that derived from a 5 g sample) was used to calculate the equivalent mass of original

dried rock material represented (e.g. 5 g x 0.10 = 0.50 g represented on the cover slip).

To obtain an estimate of the number of palynomorphs on coverslips, the number in a 1/44 traverse of

each was counted and multiplied by 44. To give an estimate of ‗absolute abundance‘, i.e. the

concentration of palynomorphs in each sample (counts per gramme, cpg), the mean number of

palynomorphs per coverslip was divided by the approximate mass of dried rock material represented

on each coverslip. Calculations for each sample are given in Supplementary Tables B–E. Relative

abundances were estimated by counting the first three hundred palynomorphs identified (0.3% = 1

specimen; 0.7% = 2 specimens; 1% = 3 specimens, etc.). The remainder of the first coverslip and,

where applicable, the three additional coverslips, were subsequently scanned for additional rare taxa.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 15: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Relative abundances are described as ‗rare‘ (outside the count), ‗frequent‘ (0.3–0.7%), ‗common‘ (1–

9.7%) or ‗abundant‘ (10% +). Separate counts of 100 kerogen particles (AOM, phytoclasts and

palynomorphs) were made from the un-oxidised kerogen slides. Further samples from Caistor, Bigby,

South Ferriby, Market Weighton, Bishop Wilton, East Knapton, Flixton and Speeton were processed

and analysed for palynology only, using the methods described above, with preliminary summary

results presented herein.

Standard palaeoenvironmental parameters have been calculated for the Melton Ross section,

including: (i) the ratio between terrestrial (T) and marine (M) palynomorphs (T/M ratio) as a proxy for

terrestrial input; (ii) the ratio between peridinioid and gonyaulacoid dinoflagellate cysts (P-cysts and

G-cysts, the P/G ratio) as a proxy of nutrient input; (iii) the species richness, i.e. the number of

dinoflagellate cyst taxa recorded as a proxy of their diversity. Our detailed discussion of

palaeoenvironmental parameters is provided in Eldrett et al. (2017); see also McLachlan et al. (2018)

for a recent review of the P/G ratio.

All Melton Ross palynological slides are curated in the MPA and MPK collections of the British

Geological Survey, Keyworth, Nottingham, U.K. (slide numbers MPA 70686 to 70778 and MPA

71640 to 71642; type and figured specimen numbers MPK 14662 to 14716). For the relationship

between MPA numbers and the original sample numbers used in this paper, see Supplementary Table

F. A full range chart of palynological data is available in the Supplementary Material. Full author

names and synonyms of dinoflagellate cysts, prasinophyte phycomata and acritarchs can be found in

Fensome et al. (2019).

Stable isotopes

All samples were analysed for stable carbon and oxygen isotopes on carbonates (Supplementary Table

G). Analyses were undertaken in the laboratory of the Department of Earth Sciences, Oxford

University, U.K. Oxygen and carbon isotope analytical methods were adapted from those described in

Day & Henderson (2011, section 2.7). All oxygen isotope measurements were performed on a Delta

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 16: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

V Advantage isotope mass spectrometer fitted with a Gas Bench II. The Gas Bench II device

converted the carbonates to carbon dioxide (CO2) with 100% phosphoric acid (H3PO4) at 72°C

(McCrea 1950). The relative 18O/16O values (δ18O) of carbonate are expressed in per mil (‰) relative

to Vienna Pee Dee Belemnite (VPDB) on a normalised scale such that the δ18O of NBS-19 is -2.2‰

and the δ18O of NBS-18 is -23.01‰. The relative 13C/12C values (δ13C) of carbonate are expressed in

per mil relative to VPDB on a normalised scale such that the δ13C of NBS-19 is 1.95‰ and the δ13C

of NBS-18 is -5.01‰. External error (0.07 and 0.09 for δ13C and δ18O, respectively) is calculated from

repeat measurements of Oxford University‘s in-house standard NOCZ. It is assumed that the

phosphoric acid–carbonate fractionation is the same for NBS-19 and Oxford University‘s calcite

samples (Coplen 1996). For carbonates and waters, results are expressed on the same normalised scale

such that δ18O of SLAP2 reference water is -55.5‰.

Results

Organic geochemistry

The two chalk samples (top Ferriby Chalk Formation, sample -23, 0.04% TOC; Central Limestone,

sample -14, 0.10% TOC) contain the lowest organic carbon values in the sampled section. The

‗anomalous‘ lower succession marls (local Beds I–VII; samples -22 to -15) range from 0.11% to

0.16% TOC, with the exception of the darker coloured marl layer (local Bed II, sample -21), which

has 0.28% TOC. In the ‗standard‘ upper succession, the lighter coloured lithologies of Bed B (sample

-13, 0.12% TOC; sample -12, 0.20% TOC), Bed D (sample -9, 0.36% TOC) and Bed F (samples -3 to

-5, 0.12% to 0.34% TOC) contain the lowest values while the dark grey marl samples contain the

highest values: Bed C (samples -11.5, -11 and -10, 1.43% to 2.2% TOC), Bed E (samples -8.5 and -8,

2.74% and 2.18% TOC respectively) and Bed G (sample -2, 3.47% TOC). The hydrogen index (HI) in

the dark grey marl samples ranges from 78 (sample -11) to 203 mg HC/g TOC (sample -2) while the

oxygen index (OI) ranges from 15 (sample -11.5) to 26 mg CO2/g TOC (sample -8) in the same

samples (Supplementary Table H). On a modified Van Krevelen diagram (Fig. 9), these samples plot

within the Type III organic matter (mainly terrestrially derived) field, with samples -11.5 and -2

plotting towards Type III-II organic matter (mixed terrigenous and marine type). The lighter coloured

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 17: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

lithologies between the dark grey marls give HI values of 40 to 78 mg HC/g TOC and relatively high

OI values (27 to 165 mg CO2/g TOC), as does sample -22 from local Bed II (HI = 39, OI = 53;

Supplementary Table H). However, with regard to characterising organic matter type, the OI data may

be unreliable in these low TOC, carbonate-rich samples, due to probable elevation by occluded CO2

within carbonate of inorganic origin, in addition to that derived from early diagenesis of organic

matter (P. Barnard pers. comm. 2019).

Palynology

In the lower succession, local Bed I (sample -22) and Bed ?III (sample -20) respectively yielded an

estimated 53 and 95 palynomorph counts per gramme (cpg) while Bed II (sample -21) has a much

higher concentration of palynomorphs, 12,408 cpg (Table 2). Beds ?IV to VII (samples -19 to -17)

yielded between 5 and 88 cpg, dominated by one taxon, Dinoflagellate? type D of Ioannides (1986)

(Table 3; Figs 6.13, 6.14). The topmost part of the lower succession (samples -15 and -16) and

overlying Central Limestone (sample -14) are palynologically barren or contain up to two

dinoflagellate cysts.

Palynological recovery and preservation from the upper succession in the freshly excavated Site 4

samples is better overall than that from many samples in other quarries investigated to date

(Supplementary Table I). Recovery from Bed B is sparse in sample -13 (28 cpg) and slightly richer in

sample -12 (288 cpg). The dark grey marl samples of the Black Band (Beds C to E) and Bed G

contain exceptionally high concentrations of palynomorphs, 85,184 to 219,648 cpg in samples -11.5, -

11, -10, 8.5, -8 and -2, i.e. in the same samples that have the highest TOC values of 1.43–3.47%. The

lighter coloured lithologies between the dark grey marls show a broad correlation between

palynomorph concentration and TOC, 17,204 to 53,284 cpg within the Black Band (samples -10.5, -9,

-7, -6), where concentration is still exceptionally high, and 1,470 to 8,668 cpg in Bed F. The

assemblage in Bed H is comparable to that in Bed G and is interpreted to be mainly derived from

reworked clasts of the latter (see Supplementary Appendix), given that Bed H is palynologically

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 18: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

barren at four other localities where it has been sampled (Louth, Bigby, Caistor and South Ferriby;

Hart et al. 1993; Supplementary Table I).

The dinoflagellate cyst assemblage from local Bed II (sample -21) is diversified with 77 taxa present

(Table 2). The adjacent samples (-22 and -20) from local Beds I and ?III, along with the samples from

Bed B (-12 and -13) contain fewer taxa (32–33) but this may mainly reflect a much smaller number of

specimens inspected from these lower recovery samples. Relatively high diversity dinoflagellate cyst

assemblages are noted from the darker lithologies of the Black Band (68–79 taxa) with slightly fewer

taxa in the lighter coloured inter-beds (Bed C, sample -10.5, and Bed D; 65 and 63 taxa respectively)

and overlying Bed F (53–69 taxa). Bed G yielded 62 taxa, fewer than from comparable dark

lithologies in the Black Band.

The lower succession has the highest P/G ratio in the sampled section (average 0.61), with P-cyst taxa

Palaeohystrichophora infusorioides (Fig. 5.13) and Subtilisphaera pontis-mariae (Figs 5.14, 5.15)

respectively comprising an average 19.4% and 23.0% of total palynological assemblages from

samples -22 to -20 (Table 4). In the upper succession, P-cysts are subordinate to G-cysts in all

samples, though they occur in higher relative and absolute abundances in the more organic-rich (> 1%

TOC), dark grey marl samples (average P/G ratio of 0.33) than the lighter coloured lithologies of

Beds B–D (average P/G ratio, 0.04) and upper Bed E to Bed F (average P/G ratio, 0.26; Table 2).

Trithyrodinium suspectum (Fig. 5.12) and Ginginodinium? sp. A of Prauss (2006, 2012a) (Figs 6.3,

6.7) occur in most of the productive samples from both lower and upper successions, respectively

comprising 0.3–7% and 0.3–1.7% of total palynological assemblages. Palaeohystrichophora

infusorioides and S. pontis-mariae also occur consistently in the upper succession. With the exception

of sample -6, the former is relatively rare (1–5%) in the Black Band and more common (7.7–11.3%)

in Bed F while the latter is more common in the > 1% TOC samples of the Black Band (10.7–18.7%).

Bosedinia cf. sp. 1 of Prauss (2012b) (Figs 7.5, 7.9, 7.10) is also relatively common in the > 1% TOC

samples (1.3–6%). Eurydinium saxoniense (Fig. 5.6) occurs consistently in Beds C and D, with its

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 19: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

first confirmed occurrence in sample -11.5 at the base of the former, and is relatively common (1.7–

8%) in Beds E, F and G (Table 4).

The G-cyst taxon Spiniferites spp. is prominent in most productive samples throughout the section,

5.4–28.3% of total palynological assemblages, while Pterodinium spp. comprises 1–5% of

assemblages, with an acme (10%) in sample -3 from Bed F (Table 3). Litosphaeridium siphoniphorum

(Figs 5.1, 8.12) is restricted to the lower succession, where it is common (1–2.7%). Cyclonephelium

compactum–membraniphorum (Figs 5.5, 6.4) is rare in the lower succession but is consistently

abundant throughout the upper succession, from basal Bed B to Bed H (average 21.5%). An isolated

specimen of Oligosphaeridium totum (Figs 5.9, 5.10, 8.9) was recorded from sample -17 but the taxon

is only consistently recorded from Bed B to Bed F and has an acme in upper Bed B (40.3%; Table 3).

Sepispinula? huguoniotii (Fig. 5.16) is common to abundant from Bed B to lower Bed E (samples -13

to -8.5).

There is a relative increase in prasinophyte phycomata, mainly the genera Leiosphaeridia,

Pterospermella and Tasmanites, in the darker intervals at Melton Ross (0.7–1% in Bed C; 1.3–6% in

Bed E; 6.3% in Bed G). Acanthomorph acritarchs are also present in higher percentages in the darker

> 1% TOC lithologies (average 3.9% of the total palynological assemblage) than in the < 1% TOC

samples (average 0.6%). They are mainly represented by Veryhachium spp., apart from Bed G

(sample -2), in which Micrhystridium is prominent (11%; Table 2).

Terrigenous palynomorphs are present in higher relative and absolute abundances in the more

organic-rich (> 1% TOC) samples (average T/M ratio of 0.103 and concentration of 12,555 cpg) than

in adjacent lithologies with lower (< 1%) TOC values (Table 2; average T/M ratio, 0.023; average

concentration, 524 cpg; Supplementary Table E). Terrigenous palynomorphs are mainly composed of

gymnosperm taxa (4.7–13.7% of total palynological assemblages from > 1% TOC samples),

predominantly bisaccate pollen, including Rugubivesiculites rugosus (Fig. 5.17) and Alisporites

microsaccus (Fig. 5.18), along with subordinate Classopollis spp. (Fig. 5.19) and Inaperturopollenites

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 20: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

hiatus. Pteridophyte spores, including Deltoidospora spp. and Gleicheniidites spp. are persistent and

rare to common but do not exceed 3% of the total palynological assemblages. Normapolles

angiosperm pollen, Atlantopollis microreticulatus (Fig. 5.20) and Complexiopollis spp. (Fig. 5.21)

occur in most productive samples but are rare and compose less than 1% of the palynological

assemblages. With the probable exception of an occurrence of the dinoflagellate cyst Sindridinium

borealis in sample -11 (reported Albian–Early Cenomanian stratigraphic range; Nøhr-Hansen et al.

2018), palynomorphs reworked from older formations were not recorded in the Melton Ross section.

Stable isotopes

The sample (-23) from the top of the Ferriby Chalk Formation yielded a δ13C value of 2.49‰. In the

lower succession, there are three ‗peaks‘ in the δ13C data above 3‰ (samples -22 to -21, -19, and -17

to -15), separated by two ‗troughs‘ below 3‰ (samples -20, -18 to -18.5). Maximum values for the

section are from the Central Limestone (Bed A, sample -14, 4.25‰) and Bed B (sample -13, 4.21‰;

sample -12, 4.17‰). Values show an overall decline through the Black Band, from 3.91‰ (sample -

11.5) to 3.52‰ (sample -6). Above a trough in Bed F (3.34‰ in sample -4), there is a subsequent

peak in Bed G (3.77‰ in sample -2). The highest sample analysed (-1) from Bed H yielded a value of

3.14‰ (Supplementary Table G).

The δ18O data fluctuate in the -3.58 to -5.12‰ range in the lower succession. Minimum negative

values for the section occur in the Central Limestone (-3.03‰ in sample -14; Fig. 4) and overlying

basal Bed B (-3.25‰ in sample -13). Upper Bed B and Beds C to H, all yielded δ18O values more

negative than -4.5‰, with three notable peaks in the samples with > 1% TOC (lower Bed C, lower

Bed E and Bed G; -5.35‰ to -6.46‰).

Stratigraphy

Micropalaeontology

A study of foraminifera at Melton Ross has not yet been undertaken. The Cenomanian planktonic

marker taxon Rotalipora cushmani (Morrow) has been recorded from the Ferriby Chalk Formation in

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 21: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

eastern England, at South Ferriby and Elsham, but has to date not been recorded from above the

erosion surface at the base of the Welton Chalk Formation (Hart et al. 1993). In most southern

England sections, it ranges up into the lower part of the Plenus Marls, e.g. Bed 3 at Dover (Jarvis et

al. 1988a) and Bed 4 at Eastbourne (Jarvis et al. 2006). At Misburg, R. cushmani occurs in

Cenomanian strata and has its last occurrence (LO) in the pre-Plenus ‗black shale‘ succession

(Hilbrecht 1986). Future inspection of our samples from Melton Ross will test for its presence in

localised Beds I–VII.

The regionally correlative Beds B–H in eastern England are dominated by a Hedbergella/Whiteinella

assemblage, with an increase in small buliminids and simple agglutinated foraminifera in the dark

coloured mudstones (Hart & Bigg 1981; Hart et al. 1993; Dodsworth 1996, fig. 6). The assemblage in

the mudstones adjacent to the Black Band is comparable to those of the upper part of the Plenus Marls

in southern England (Beds 4–8). This led Hart & Bigg (1981) to suggest that the Black Band may be

the lateral equivalent of Bed 6, a relatively clay-rich unit. On grounds of event stratigraphy, i.e. the

most argillaceous levels within the CTB successions of the Northern and Southern Provinces, Jeans et

al. (1991) also suggested correlation of eastern England Black Band, Beds C and E, with southern

England Beds 4 and 6, respectively. However, a correlation of the Black Band with a level in the

Plenus Marls is not supported by the palynological and stable isotopes data presented in this study

(see below).

Hart & Bigg (1981) found the first occurrence (FO) of Helvetoglobotruncana praehelvetica (Trujillo)

in Bed H at Elsham. This taxon is an early morphotype of Helvetoglobotruncana helvetica (Bolli), the

diagnostic zonal marker for the Early to Middle Turonian in Tethyan areas. The FO of H. helvetica

was tentatively recorded from a marl seam at South Ferriby (Hart & Leary 1989) that probably

correlates with the marl seam above the Turnus Bed at Melton Ross (Fig. 4). However, its FO is based

on the (highly subjective) evolutionary boundary between praehelvetica and helvetica (Hart & Leary

1989). The rarity of H. helvetica at high latitudes such as Lincolnshire also renders its FO

impracticable as a confident time-diagnostic event in the Northern Province (Hart 2019). In NW

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 22: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Germany, Hilbrecht (1986) reported the FO of H. helvetica within the Fish Shale. At Dover, southern

England, its FO is difficult to locate precisely in the nodular chalk lithology of the Ballard Cliff

Member (above the Plenus Marls), as this is difficult to process for calcareous microfossils (Hart &

Leary 1989). At Eastbourne, the FO of H. helvetica is picked between Mead Marls 3 and 4 of the

Ballard Cliff Member (Jarvis et al. 2006). In Fig. 4, the planktonic foraminiferal zonation is

extrapolated to Melton Ross from Elsham and South Ferriby.

Palynology

Several dinoflagellate cysts have widespread range bases and tops in the Cenomanian and Turonian

stages, and have been used to zone their substages (Clarke & Verdier 1967; Burgess 1971; Foucher

1981; Williams 1977; Williams et al. 2004). Regional zonations have recently been revised for

Central and Northern Europe (Olde et al. 2015a) and adapted for the Western Interior of the USA

(Dodsworth & Eldrett 2019).

The LO of consistent and common Litosphaeridium siphoniphorum in the lower succession at Melton

Ross (sample -20) correlates with its LO in the pre-Plenus sequence of NW Germany, at Wunstorf

outcrop (Marshall & Batten 1988) and core (van Helmond et al. 2015), and Misburg outcrop

(Marshall & Batten 1988). In southern England, the LO of common L. siphoniphorum occurs in

Plenus Marls Bed 6 at Lulworth and Eastbourne (Fig. 1; Dodsworth 2000; Pearce et al. 2009) while

isolated specimens have been recorded from Bed 7 and Bed 8 at Lulworth (Dodsworth 2000). Isolated

specimens have also been sporadically recorded from Turonian and Coniacian deposits at some

localities in southern England, France, NW Germany (Clarke & Verdier 1967; Foucher 1982, 1983;

Marshall & Batten 1988) and Pueblo, Colorado (Dodsworth 2000). These isolated Turonian and

younger occurrences are probably reworked specimens. Previous reports of the LO of consistent L.

siphoniphorum in Lower Turonian deposits (Williams & Bujak 1985; Costa & Davey 1992) probably

derive from former assignment of the Plenus Marls in southern England to this substage (e.g. Jefferies

1962, 1963); the Plenus Marls have subsequently been confidently reassigned to a Late Cenomanian

age (e.g. Jarvis et al. 1988a; Gale et al. 1993). A Late Cenomanian LO of consistent L. siphoniphorum

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 23: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

has been reported worldwide, including other locations in southern England (Davey 1969; Hart et al.

1987; Jarvis et al. 1988b), the Witch Ground Graben, central North Sea (Harker et al. 1987), France

(Foucher 1979, 1980; Courtinat et al. 1991), northern Spain (Mao & Lamolda 1999), Crimea

(Dodsworth 2004a), Poland (Dodsworth 2004b), eastern USA (Aurisano 1989), the Western Interior

of the USA (Courtinat 1993; Dodsworth 2000, 2016; Eldrett et al. 2015a; Dodsworth & Eldrett 2019),

Ocean Drilling Project holes including Demerara Rise (Fig. 1; Leg 207, Site 1260) and Kerguelen

Plateau (Leg 183, Site 1138; Eldrett et al. 2017), Australia (Morgan 1980; McMinn 1988) and New

Zealand (Hasegawa et al. 2013; Schiøler & Crampton 2014).

At Melton Ross, the LO of Pterodinium crassimuratum (Fig. 5.7) occurs in sample -21, close to the

LO of consistent/common L. siphoniphorum in sample -20. Davey & Williams (1966) and Clarke &

Verdier (1967) recorded it from beds of Middle and Late Cenomanian age in southern England, its LO

coinciding with that of L. siphoniphorum within the Plenus Marls on the Isle of Wight and at a

slightly higher level at Eastbourne, in Plenus Marls Bed 8 (Pearce et al. 2009). At Pueblo, Colorado,

the LO of P. crassimuratum also occurs in the Upper Cenomanian (Dodsworth & Eldrett 2019).

The FO of abundant Cyclonephelium compactum–membraniphorum in lower Bed B (sample -13),

immediately above the Central Limestone at Melton Ross, correlates with the base of its abundant

occurrence immediately above the Plenus Bed in NW Germany, at Wunstorf outcrop (Marshall &

Batten 1988) and core (van Helmond et al. 2015), and Misburg outcrop (Marshall & Batten 1988).

The event is less clearly defined in southern England, but can be tentatively picked near the top of the

Plenus Marls, in Bed 7 at Eastbourne (Pearce et al. 2009) and Bed 8 at Lulworth (Dodsworth 2000).

Cyclonephelium membraniphorum is rare or absent in coeval Tethyan sections farther south, in Spain

(Lamolda & Mao 1999; Peyrot 2011; Peyrot et al. 2011, 2012) and Morocco (Prauss 2012a, b).

At Melton Ross, an influx of Oligosphaeridium totum from Bed B to lower Bed C can be correlated

with an influx of O. totum from the same levels at Caistor (samples CL-11 to CL-9; 1–3%) and from

upper Bed B at Market Weighton (samples MW-20 to MW-18; 3–6%; see Supplementary Table I for

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 24: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

sample positions). Duane (in Hart et al. 1993) recorded its highest relative abundance (3.4%) in the

lower part of the Black Band (their sample SFE-18) at South Ferriby (the underlying Bed B samples

SFE-19 and SFE-20 at South Ferriby were palynologically barren). In NW Germany, Marshall &

Batten (1988) only recorded O. totum from above the Plenus Bed at Wunstorf (their sample 38) and at

Misburg (their samples 27, 22 and 20). At Misburg, an influx of O. totum was indicated, with its base

in sample 22, collected from a bed that contains the only occurrence of the latest Cenomanian zonal

ammonite Neocardioceras juddii in the section. An acme of O. totum occurs in overlying sample 20.

In central Poland, O. totum occurs in sample Pul-17 of the Pulawy borehole (Fig. 1), within the δ13C

excursion that spans the CTB (Peryt & Wyrwicka 1993), one sample above the FO of abundant C.

compactum–membraniphorum (sample Pul-16). This, in turn, is stratigraphically higher than the LO

of consistent/common L. siphoniphorum (sample Pul-15). Thus, the three bio-events also occur in the

same relative order over a condensed (< 1 m thick) interval in that section (Dodsworth 2004b).

The LO of Adnatosphaeridium tutulosum (Fig. 5.4) occurs in Bed F at Melton Ross (sample -5). It

was recorded from Bed C at Flixton (Dodsworth 1996). At other European and North America

locations, the LO of A. tutulosum also occurs within the uppermost Cenomanian, above the LO of

consistent/common L. siphoniphorum and the FO of abundant C. compactum–membraniphorum: in

NW Germany, above the Plenus Bed at Wunstorf and Misburg outcrops (Marshall & Batten 1988); in

southern England, at Hooken Cliffs, Devon (Jarvis et al. 1988b), Eastbourne (Plenus Marls Bed 8;

Pearce et al. 2009) and 0.5 m above the Plenus Marls at Lulworth (Dodsworth 2000); and in southern

France (cf. Courtinat et al. 1991 & Jarvis et al. 2011) and the Western Interior of the USA

(Dodsworth 2000; Harris & Tocher 2003; Dodsworth & Eldrett 2019). The LO of A. tutulosum is at

the same level as the LO of L. siphoniphorum in northern France (Foucher 1983) and Crimea

(Dodsworth 2004a).

The LO of Carpodinium obliquicostatum (Fig. 5.3) occurs in Bed C (sample -11.5) at Melton Ross. It

was recorded from Bed E at South Ferriby and Flixton (Dodsworth 1996). In NW Germany, Marshall

& Batten (1988) reported it above the Plenus Bed in the Wunstorf and Misburg outcrops. In southern

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 25: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

England, it occurs 0.5 m above the Plenus Marls at Lulworth (Dodsworth 2000). Although rare and

sporadic in occurrence, its range top in the Western Interior is reported to coincide approximately

with that of A. tutulosum at Pueblo, Colorado (Dodsworth 2000).

Microdinium setosum (Fig. 6.6) occurs fairly consistently in the Black Band at Melton Ross, with its

LO in Bed E (sample -6). In southern England, it occurs throughout the Cenomanian (Clarke &

Verdier 1967) and is sporadic in the Plenus Marls with an LO in Bed 2 at Eastbourne (Jarvis et al.

2011) and Bed 7 at Lulworth (Dodsworth 2000). However, it occurs sporadically as high as Middle

Turonian in the North Sea Basin (Costa & Davey 1992).

At Melton Ross, Adnatosphaeridium? chonetum (Fig. 5.8) was recorded from local Bed II (sample -

21), with an isolated, questionable specimen also present in Bed E (sample -6). It has previously been

recorded from Bed C at Flixton (Dodsworth 1996). The occurrence of A.? chonetum in Cenomanian

deposits was noted by Davey (1969) in Northern Europe and by Cookson & Eisenack (1962) and

Backhouse (2006) in Australia. In the Western Interior of the USA, it is common in the Upper

Cenomanian Substage, with an uppermost Cenomanian LO (Harris & Tocher 2003; Dodsworth &

Eldrett 2019). In the Shetland Group, offshore Norway, it reappears in Coniacian and Lower

Santonian deposits (PD, personal observation).

The FOs of Canningia glomerata, Heterosphaeridium difficile, Florentinia buspina, F.? torulosa and

Senoniasphaera turonica are intra-Lower Turonian biostratigraphic marker events in Europe (Davey

& Verdier 1976; Foucher 1980, 1981, 1983; Tocher & Jarvis 1987; Jarvis et al. 1988a; Costa &

Davey 1992; FitzPatrick 1995; Pearce et al. 2003, 2009, 2011). These taxa have not been recorded at

Melton Ross or in any of the other eastern England localities/samples indicated in Supplementary

Table I. Thus, Beds C–G are probably stratigraphically lower than their range bases.

In terms of dinoflagellate cyst zonation, the palynologically productive samples from the lower

succession at Melton Ross are assigned to the Litosphaeridium siphoniphorum Interval Zone of Olde

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 26: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

et al. (2015a) and Dodsworth & Eldrett (2019). The productive part of the upper succession (Bed B to

Bed G) is assigned to the Cyclonephelium membraniphorum Zone of these authors, with Bed B to

lowermost Bed F belonging to the Adnatosphaeridium tutulosum Subzone of Dodsworth & Eldrett

(2019; Fig. 4).

At Wunstorf, Prauss (2006) reported acanthomorph acritarchs. Higher relative abundances of

Veryhachium (ca. 1–2%) occur in darker lithologies from the pre-Plenus and lower post-Plenus

succession, relative to lighter coloured inter-beds. There is a large influx of Micrhystridium in the

upper part of the Fish Shale (ca. 15–70%). Following Wall (1965) and Downie et al. (1971), Prauss

(2006) attributed the Veryhachium occurrences to a distal offshore, hydrodynamically quiet

environment, and the prominent acritarch peak dominated by Micrhystridium within the upper part of

the Fish Shale to the influence of a relatively near-shore turbulent water environment. An up-section

change from common Veryhachium (Beds C to E) to abundant Micrhystridium (Bed G) at Melton

Ross may correlate with that reported from Wunstorf.

The consistent presence of the bisaccate pollen Rugubivesiculites rugosus in the upper succession at

Melton Ross is noteworthy. It has recently been reported throughout the Upper Cenomanian–Middle

Turonian section exposed at Pueblo, Colorado (Dodsworth & Eldrett 2019) and in Turonian cores

from Texas (e.g. Iona-1; Eldrett et al. 2017). In the Greenland area, Nøhr-Hansen (2012, appendix 1),

Fensome et al. (2016) and Nøhr-Hansen et al. (2016) indicated that Rugubivesiculites spp. / R.

rugosus has a consistent LO within the Turonian.

Calcareous nannofossils

A study of calcareous nannofossils at Melton Ross has not yet been undertaken, but Bralower (1988)

reported their ranges from the CTB interval at other quarries, including the nearby sections at South

Ferriby and Elsham (Fig. 2). The LOs of the coccoliths Axopodorhabdus albianus (Black), Helenea

chiastia Worsley and Rhagodiscus asper Stradner occur in Bed H at South Ferriby. At Elsham, the

LO of R. asper also occurs in Bed H but the LOs of H. chiastia and A. albianus occur respectively 20

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 27: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

cm and 15 cm lower, within undifferentiated variegated marls. At Misburg, NW Germany, all three

events occur within Lower Turonian deposits above the Fish Shale (cf. Bralower 1988, fig. 21, and

Hilbrecht & Hoefs 1986, fig. 2), with the LOs of H. chiastia and A. albianus in a ‗triple band‘ of

stratigraphically higher, dark-coloured mudstones, and the LO of R. asper ca. 5 m above the ‗triple

band‘. At Dover, southern England, the three events occur stratigraphically lower, within Upper

Cenomanian deposits, i.e. the highest 0.55 m of the Plenus Marls which is 4.45 m thick at the sampled

location. At Pueblo, Colorado, the three events occur over a ca. 4.5 m interval that spans the CTB, in

the stratigraphic order of LO A. albianus, LO H. chiastia and LO R. asper (Fig. 10). Bralower (1988)

suggested that slow sedimentation rates, a probable hiatus between Bed H and the Turnus Bed, and

subsequent bioturbation may have ‗smeared‘ these events at South Ferriby and Elsham, while at

Dover they occur in proximity to a lithological change to the hard, condensed limestone of the Ballard

Cliff Member that overlies the Plenus Marls.

Stable isotopes

Before discussing the possible stratigraphic significance of the measured changes in the δ13C stable

isotope record, diagenetic or local lithological influences have to be excluded. Diagenetic alteration is

commonly observed in fine-grained carbonate sedimentary rocks. Deep burial cementation and

recrystallization can result in the addition of isotopically depleted calcite to the bulk carbonate pool,

shifting the bulk δ18O record towards lower values (e.g. Jarvis et al. 2011, 2015). However, during

burial diagenesis, carbon isotope values are less prone to diagenetic alteration than oxygen isotope

values as the carbon isotope system is rock-dominated and δ13C is subject to a much smaller

temperature-controlled fractionation (Marshall 1992). A cross-plot of δ13C and δ18O allows the

elucidation of the depositional and post-depositional controls on stable isotope values, with a strong

co-variance between δ13C and δ18O indicating potential diagenetic influence (Fig. 11).

In general, the stable isotope results for the Melton Ross section show poor co-variance (R2= 0.0076),

indicating limited diagenetic overprint. However, the samples from Beds I–VII and Beds A and B

that contain low TOC (< 0.2%) do show a relatively strong co-variance (R2= 0.84), indicating

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 28: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

potential dissolution and re-precipitation of isotopically light 13C cements through interaction with

meteoric pore fluids during burial in this interval. Conversely, the δ13C and δ18O co-variance remains

poor in samples with low TOC from Beds C–H. It is uncertain why the lower part of the succession

may be preferentially affected by diagenesis. The range of δ13C and δ18O values from Melton Ross

appear consistent with bulk stable isotope values from coeval European Cenomanian–Turonian

sections that record a 2–5‰ positive carbon isotope excursion (CIE) associated with OAE-2 (Fig. 11).

The δ13C profile from Beds A–H at Melton Ross broadly correlates with that assigned to OAE-2 at

South Ferriby (Schlanger et al. 1987; Hart et al. 1991; Clarkson et al. 2018), which also shows

highest δ13C values in Beds A and B and an overall decreasing trend in values through Beds C to H,

with a relatively minor, terminal peak around Bed G/H. The South Ferriby studies sampled metres of

chalk below and above Beds A–H, and indicated ca. 1‰ lower δ13C background values below and ca.

0.5‰ lower background values above. Therefore, although the δ18O record may have been influenced

by diagenesis, the δ13C signal appears less impacted and possibly reflects a primary

palaeoceanographic signal. A similar conclusion was drawn by Hu et al. (2012) and Mitchell (2019)

for the chalks and marls from the Northern Province.

The detailed profile of the CIE associated with OAE-2 has been shown to correlate across the

Southern Province in England, including Dover and Eastbourne (Jarvis et al. 2006), with Eastbourne

being taken as the principal reference section (Paul et al. 1999). It has also been correlated inter-

regionally, particularly with the USA (e.g. Gale et al. 1993; Joo & Sageman 2014; Eldrett et al.

2015a; Fig. 10), Morocco (Tsikos et al. 2004; Jenkyns et al. 2017), Japan (Uramoto et al. 2013) and

other areas of Europe (Jarvis et al. 2015). In Misburg outcrop (Hilbrecht 1986), Gröbern core (Voigt

et al. 2006) and Wunstorf core (van Helmond et al. (2015), NW Germany, background δ13C values

are recorded in the pre-Plenus sequence while peak values associated with OAE-2 occur within the

Plenus Bed. The new record of peak δ13C values in the Central Limestone at Melton Ross supports its

correlation with the Plenus Bed, as initially predicted by Wood & Mortimore (1995). The magnitude

and top of the δ13C excursion vary across NW Germany (Hilbrecht et al. 1992), but it shows an

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 29: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

overall decreasing trend through the Fish Shale at Misburg and Wunstorf, possibly correlating with

that seen through Beds C to H at Melton Ross and South Ferriby.

In the Southern Province of England, a ‗build-up phase‘ of overall increasing δ13C values through

Beds 1–2 of the Plenus Marls precedes a lower peak (‗a‘) around Bed 3, a relatively thick limestone

unit. Praeactinocamax plenus occurs in the overlying part of the Plenus Marls (Beds 4–8),

particularly in Bed 4 (Jefferies 1962, 1963). Above a trough in δ13C values through Beds 4–8, a

second and maximum peak (‗b‘) occurs close to the boundary between the Plenus Marls and overlying

Ballard Cliff Member. Above peak ‗b‘, δ13C values remain high but there is an overall decline in

values across the CTB, through the Ballard Cliff Member, including some prominent marl seams

(Mead Marls 1–6), below a final OAE-2 peak ‗c‘ (Jarvis et al. 2006).

In terms of correlation between the Northern Province and Southern Province in England, our new

δ13C data from Melton Ross could be interpreted as follows. The three (> 3‰) peaks of δ13C in Beds

I–VII may be ‗precursor‘ / build-up events to OAE-2, correlating with southern England Beds 1–2;

the base of the > 4‰ interval in Bed A (Central Limestone) equals peak ‗a‘ of Jarvis et al. (2006;

southern England Bed 3); lower Bed B may correlate with part of Beds 4–8 in southern England,

supported by the record of P. plenus at Melton Ross, Site 1 (Wood et al. 1997; Fig. 4); the top of the >

4‰ interval in upper Bed B equals peak ‗b‘ of Jarvis et al. (2006); the trough in Beds C–F and peak

in Bed G may correlate with the trough through the Ballard Cliff Member and peak ‗c‘ in southern

England (Fig. 10). A comparable correlation has been made between the δ13C profile from Beds A–H

at South Ferriby and that of the Southern Province (Wood & Mortimore 1995; Clarkson et al. 2018).

Notwithstanding concerns about diagenetic alteration of the eastern England Chalk Group oxygen

isotope signal discussed above, the δ18O curve from Melton Ross (Fig. 4) is comparable to those from

other parts of Europe (cf. Jarvis et al. 2011, fig. 8). The minimum values in the Central Limestone and

overlying lower Bed B may correlate with the PCE. The higher values from Beds C–H could be

considered consistent with subsequent warming around the CTB. The peaks in δ18O values from the

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 30: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

most organic-rich layers could reflect temperature maxima and/or increased run-off during their

deposition (see below).

Discussion

Organic geochemistry

In an overview of OAE-2 in North European shelf settings, Jenkyns (1985) reported TOC values

mainly within the range of 1–3% and HI from 150–300 mg HC/g TOC (mixed marine and terrigenous

sources) in Cenomanian–Turonian organic-rich deposits, including eastern England (South Ferriby,

0.79–3.15% TOC) and NW Germany (1.2–2.8% TOC; Schlanger et al. 1987). The new HI data from

the > 1% TOC dark mudstones at Melton Ross plot around the lower end of this range (78–203 mg

HC/g TOC; Fig. 9), consistent with a relatively large terrigenous component. Jenkyns (1985) noted

that there appears to be a general trend of an increase in preservation of organic carbon as the

depositional environment deepened from stable, relatively shallow shelf to rifted graben and deeper

continental margin. Higher TOC values have since been reported from the Black Band farther north in

eastern England, from Speeton (13% TOC, Farrimond et al. 1990) and Flixton (4.5–10.2% TOC,

Jeans et al. 1991) at the edge of the Cleveland Basin (Fig. 2). Herbin et al. (1986) reported 30% TOC

from the Black Band Bed in the Central Graben of the North Sea, including large quantities of Type

III organic matter, consistent with a substantial terrigenous component. In palynological assemblages

from Melton Ross and other onshore Black Band localities, marine organic-walled phytoplankton

dominate over terrigenous pollen and spores (Supplementary Table I), although at Flixton, the latter

compose approximately one half of the assemblages from Bed C.

The Cyclonephelium compactum–membraniphorum issue

The southward incursion of abundant Cyclonephelium compactum–membraniphorum into Central

Europe has been attributed to a stressed marine environment associated with oceanic anoxia (Marshall

& Batten 1988; Courtinat et al. 1991; Hart et al. 1991). However, its initial migration has more

recently been correlated with the PCE, including sites outside Europe. The latter include a proto-

Atlantic coastal setting in New Jersey (Bass River; van Helmond et al. 2014; Fig. 1) and the Western

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 31: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Interior Seaway of the USA (e.g. Portland-1 core, central Colorado; Iona-1 core, SW Texas; Eldrett et

al. 2014, 2017; van Helmond et al. 2016; Fig. 1). At Melton Ross, its basal abundant occurrence in

sample -13 is at the same level as records of Praeactinocamax plenus (Fig. 4) and it is consistently

abundant up to the top of the sampled section (sample -1). While the initial southward migration may

have been linked to the PCE, its subsequent abundance in latest Cenomanian and Early Turonian

deposits occurred during a time of extreme warmth. Its post-PCE prosperity during the OAE-2

interval may be related to the relative tolerance of areoligeracean dinoflagellate cysts in general to

stressed marine environments, where their abundant occurrence is often associated with nearshore

environments / falls in relative sea level (Brinkhuis & Zachariasse 1988; Harker et al. 1990, p. 202-

204; Li & Habib 1996; Olde et al. 2015b). In this case, the environmental stress is possibly associated

with oxygen-deficient conditions in the water column and photic zone (Marshall & Batten 1988)

and/or increased runoff (van Helmond et al. 2015).

Samples with prominent Dinoflagellate? type D of Ioannides (1986)

Dinoflagellate? type D occurs consistently throughout the sampled section at Melton Ross, and in

dominant relative abundance (48–100% of palynological assemblages) in the poorly preserved

material from samples -20 to -17, from the local Beds III?–VII. However, inspection of the absolute

abundance data (Table 3) indicates that its occurrences in these beds are in the single units to tens per

gramme range, less than its hundreds per gramme concentration estimates from beds with richer

recovery. It is suggested here that its dominance in Beds III?–VII is due to its preferential

preservation, probably related to a relatively thick and resilient wall structure. This interpretation is

supported by the data of Batten (in Wood et al. 1997), who reported diversified dinoflagellate cyst

assemblages with L. siphoniphorum up to local Bed VI at Melton Ross, Site 2, i.e. in correlative strata

a few tens of metres away (Fig. 3).

Weathering of exposed quarry sections probably destroys palynomorphs and other organic matter

over a short time, of months to a few years, with only central levels within the Black Band being

unaffected at some other sampled localities, e.g. South Ferriby and Caistor. Weathered samples

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 32: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

contain lower concentrations of poorly preserved palynomorphs, mainly those with relatively thick

walls: Dinoflagellate? type D, Kalyptea spp., prasinophyte phycomata and indeterminate peridinioid

endocysts. Samples worst affected are marked with an asterisk in Supplementary Table I.

Terrigenous input

In eastern England, the higher T/M ratio in relatively organic-rich (> 1% TOC) lithologies

documented here from Melton Ross (Table 2) has previously been noted in Black Band samples from

Flixton (Dodsworth 1996, fig. 9; T/M ratio 0.27–0.54 in 4.5–10.2% TOC samples, and 0–0.04 in 0.1–

0.4% TOC samples). Analyses from other Yorkshire sections (Fig. 2) appear to confirm this pattern:

East Knapton (T/M ratio average 0.11 in dark lithologies and 0.02 in light lithologies), Market

Weighton (T/M ratio average 0.08 in dark lithologies and 0.01 in light lithologies) and, to a lesser

extent, Bishop Wilton (T/M ratio average 0.06 in dark lithologies and 0.04 in light lithologies;

Supplementary Table I). This trend cannot be discerned at other Lincolnshire localities (South

Ferriby, Bigby and Caistor), where Bed B and Bed F/G samples are heavily weathered or barren and

Bed D is relatively argillaceous and tentatively picked on slightly less dark marls within the Black

Band.

At Wunstorf, Prauss (2006) calculated an average T/M ratio of ca. 0.2 for the CTB succession. Prauss

(2006) and van Helmond et al. (2015) also reported elevated relative and absolute abundances of

spores and pollen in dark coloured mudstones compared with interbedded organic-poor lithologies,

particularly within the Fish Shale (T/M ratio ca. 0.2–0.46, in ca. 1–3% TOC rocks). Van Helmond et

al. (2015) interpreted this pattern as indicative of an intensified hydrological cycle during OAE-2, i.e.

increased evaporation, precipitation and run-off during ‗black shale‘ deposition, associated with

global warming after the PCE (see also Heimhofer et al. 2018). Reworking of mud to the basin (e.g.

Jenkyns 1980) and/or the formation of swamps due to drowning of land masses (cf. Ioannides et al.

1976) during phases of CTB transgression could also have contributed to the increase in spores and

pollen (Dodsworth 2000).

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 33: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

In eastern England, gymnosperm bisaccate pollen are invariably the dominant terrigenous

palynomorph group in the Black Band and Bed G (Dodsworth 1996; this paper). The productive

samples from the lower succession at Melton Ross (samples -22 to -20; Table 2) contain few

terrigenous palynomorphs (T/M ratio 0.01 or less) which are also mainly bisaccate pollen. At

Wunstorf, however, saccate pollen constitute less than one half of the terrigenous palynomorphs in

most samples from the CTB succession, with trilete spores and non-saccate pollen often composing

most of the assemblage (Prauss 2006; van Helmond et al. 2015). This contrast between the regions

probably reflects greater proximity to sources of detrital input in the Lower Saxony Basin, as reflected

by a much thicker succession there. Prauss (1993) noted that trilete spores show higher relative

abundances than saccate pollen in marine deposits with greater fluvial input because the latter are

more readily transported by wind (e.g. Mudie & McCarthy 1994). Different hydrodynamic properties,

i.e. the relative buoyancy of saccate pollen, is probably also a factor in their differential sorting with

distance from fluvial-deltaic sources (cf. Muller 1959). In the Cenomanian deposits of Texas, spores

form much larger proportions of terrigenous palynomorph assemblages in the proximal East Texas

Basin than in the relatively distal Maverick Basin (Dodsworth 2016).

Pollen provincialism

Herngreen & Chlonova (1981), Herngreen et al. (1996) and Traverse (2007) described Late

Cretaceous palynofloral provinces. Costa & Davey (1992) reported the southerly limit of triprojectate

pollen (the Aquilapollenites Province) to correlate approximately with the northerly siliciclastic

Shetland Group – southerly calcareous Chalk Group transition in Northern Europe, around modern

latitude 59°N. The rare but consistent occurrences of Normapolles pollen from the CTB succession at

Melton Ross complement previous records of their presence in the Black Band at Flixton, South

Ferriby and North Sea well 47/10-1 (Fig. 2; Dodsworth 1996). These may be the most northerly

published Cenomanian–Turonian records in Europe (cf. Peyrot et al. 2008, fig. 4) and indicate

assignment of the eastern England Chalk Group to the Normapolles Province. Normapolles pollen are

a more common component of palynological assemblages from coeval deposits farther south in

Europe (e.g. Bulgaria, Pavlishina & Minev 1998; Romania, Ion et al. 2004; SE France, Heimhofer et

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 34: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

al. 2018; Spain, Peyrot et al. 2008, 2011). In SE France, Heimhofer et al. (2018) suggested that the

PCE may have fostered a first spread of Normapolles-type angiosperms. Climatic conditions would

have been cooler and less humid during the PCE, possibly resulting in an open, savanna-type

vegetation community with increased abundances of Normapolles-producing angiosperms.

Phytoplankton productivity and preservation

The P-cyst Palaeohystrichophora infusorioides and the G-cyst Spiniferites ramosus constitute the

main components of European dinoflagellate cyst assemblages from Late Cretaceous offshore/deeper

water environments and upwelling zones (Pearce et al. 2003; Olde et al. 2015b). The P-cyst genera

Palaeohystrichophora and Subtilisphaera have been suggested as heterotrophic taxa. High numbers

may reflect increased nutrient availability from upwelling in the Atlantic Ocean off the NW African

(Prauss 2012b) and European margins, and subsequent transport of nutrients via a proto-Gulf Stream

to the Anglo-Paris Basin (Prince et al. 1999, 2008; Pearce et al. 2009). Subtilisphaera pontis-mariae

and P. infusorioides mainly compose the high P/G ratio in Melton Ross local Beds I–?III (average

0.61). A source for enhanced nutrient supply during this interval is uncertain. The low T/M ratio is

inconsistent with increased run-off. Upwelling as a nutrient source requires a thermal difference and

sea floor topographical variation but there is no evidence for these in eastern England.

Phytoplankton productivity during OAE-2 ‗black shale‘ deposition has previously been discussed

from Wunstorf. Linnert et al. (2010) studied assemblages of calcareous nannofossils, showing a shift

from a generally oligotrophic ecosystem during deposition of lighter coloured beds to more

mesotrophic or even eutrophic conditions during deposition of dark coloured mudstones. Van

Helmond et al. (2014, 2015) suggested that an acceleration of the hydrological cycle during the

warmer intervals of OAE-2 may have played a key role in supplying nutrients offshore and enhancing

stratification, thus contributing to the development of ocean anoxia. Increased precipitation, run-off

(reflected by a higher T/M ratio) and associated nutrients may have contributed to productivity of

organic-walled phytoplankton during deposition of the dark coloured layers in eastern England.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 35: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Within the Black Band and Bed G at Melton Ross, common Eurydinium saxoniense and Bosedinia cf.

sp. 1 of Prauss (2012b) and rare Bosedinia laevigata, make a greater contribution to the P/G ratio in

the > 1% TOC samples (Table 4). Eurydinium saxoniense was initially described in higher relative

abundances in dark coloured mudstone layers from Wunstorf and Misburg by Marshall & Batten

(1988), who associated it with a stressed marine environment with a low level of oxygen extending

high up the water column.

High relative abundances of Bosedinia spp. in Late Cretaceous marine settings have been interpreted

as reflecting reduced salinity surface water and enhanced density stratification (at Tarfaya; Prauss

2012b, fig. 5), and/or increased nutrient (nitrite/nitrate) availability from oxygen-deficient waters

encroaching into the photic zone (in the Western Interior of the USA and Demerara Rise; Dodsworth

2016; Eldrett et al. 2017). In the Black Band and Bed G, higher numbers of E. saxoniense and

Bosedinia spp. occur in the absence of fresh/brackish-water algae and may record some stimulation of

dinoflagellate productivity by increased supply of reduced nitrogen chemo-species in photic-zone

waters (cf. Prauss 2007). This is supported by the relative increase in prasinophyte phycomata in the

darker layers at Melton Ross, Flixton (5.7–9.3% in Bed C, 14% in Bed E; Dodsworth 1996) and

Wunstorf (up to 15%; Prauss 2006). Prasinophyte prosperity may be mainly related to productivity

stimulated by introduction of nitrogen/ammonium-enriched waters of the denitrification zone into the

photic zone (Prauss 2006, 2007). However, the average P/G ratio at Melton Ross (0.33) in the > 1%

TOC samples is much lower than that discussed above for Bosedinia-dominated organic-rich

lithologies from, for example, the Cenomanian of the Western Interior (ca. 0.75–0.85), and may

therefore, reflect a relatively moderate increase in peridinioid productivity. Likewise, prasinophyte

increases are moderate relative to uppermost Cenomanian organic-rich mudstones in Aksudere,

Crimea (Fig. 1; up to 7.2% TOC; Naidin 1993), where they typically constitute over half the

palynomorph assemblage (Dodsworth 2004a). Regardless of the relative contribution of organic-

walled phytoplankton productivity to OAE-2, the higher proportions of E. saxoniense, Bosedinia and

prasinophytes in the > 1% TOC lithologies at Melton Ross and Flixton are consistent with increased

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 36: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

quality of sea floor preservation in a low O2 environment. They may indicate periodic

hydrographically restricted and eutrophic conditions (cf. Eldrett et al. 2017).

Palynomorph concentration

Most palynological studies of Cenomanian and Turonian deposits report relative abundances only.

Records of specimen numbers per microscope slide from Dover, Eastbourne and the Isle of Wight,

southern England (Jarvis et al. 1988a; Fitzpatrick 1996), have given a semi-quantitative indication of

recovery. Until the last ten years, there have been few reports of fully quantitative concentration,

counts per gramme (cpg) data. Duane (in Paul et al. 1994), using a volumetric method, reported

dinoflagellate cysts in the hundreds (mainly 100–800 cpg) range from Middle Cenomanian chalks and

marls in the Dover–Folkestone Warren area, southern England. Pearce et al. (2003), using a method

involving adding a known number of Lycopodium spores to samples (Stockmarr 1971; Mertens et al.

2009), reported dinoflagellate cysts in the tens to 600 cpg range from Turonian chalks and marls at

Banterwick Barn, southern England.

Much higher concentrations of palynomorphs (mainly dinoflagellate cysts) were reported from Upper

Cenomanian deposits in southern England, at Eastbourne and Lulworth. Pearce et al. (2009), using the

Lycopodium method, reported ca. 50,000–75,000 cpg from six marl and chalk samples over a 10 m

interval below the Plenus Marls at Eastbourne. Dodsworth (2000), using this study‘s volumetric

method, reported nearly 38,000 cpg from a marl sample one metre below the Plenus Marls at

Lulworth. Pearce et al. (2009) noted that assemblages from below the Plenus Marls at both localities

are dominated by the P-cyst Palaeohystrichophora infusorioides. They attributed its high

concentrations to upwelling in the Atlantic Ocean off the European and NW African margins, with

associated nutrients being transported to the southern England area by dominantly south-westerly

surface winds and ocean surface currents. Both Dodsworth (2000) and Pearce et al. (2009)

documented overall decreasing concentrations of palynomorphs through the Plenus Marls (to just over

100 cpg in Bed 8 at Lulworth) and barren or impoverished recovery in the overlying Ballard Cliff

Member. Jarvis et al. (1988a), FitzPatrick (1996) and Pearce et al. (2003) also reported a

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 37: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

palynologically barren or impoverished interval above the Plenus Marls at other southern England

localities, with better palynological recovery from Turonian chalks and marls above the Ballard Cliff

Member and Holywell Member.

Dodsworth (2000) attributed poor palynological recovery from the upper part of the Plenus Marls and

the Ballard Cliff Member to upward coarsening through the succession (Hancock 1989; Jeans et al.

1991; Lamolda et al. 1994) and the associated reduced palynomorph preservation potential of

lithologies with a relatively high coarse silt and sand grade component. Pearce et al. (2009)

acknowledged this but, following Jarvis et al. (1988a), FitzPatrick (1996) and Lamolda & Mao

(1999), suggested that a primary decline in phytoplankton productivity across the CTB in the Anglo-

Paris Basin may have been an additional cause of the low recovery. The Eastbourne and Lulworth

localities record a marked decrease in the relative and absolute abundance of P. infusorioides above

Bed 4 of the Plenus Marls, consistent with a decrease in nutrient supply (Pearce et al. 2009). An

overall decline in the nannofossil fertility index also occurs between Bed 4 and Ballard Cliff Member

(Gale et al. 2000).

Palynomorph concentration data are becoming more frequently available from areas outside southern

England. From the CTB succession at Bass River, New Jersey, van Helmond et al. (2014), using the

Lycopodium method, reported dinoflagellate cysts in the 2,000–30,000 cpg range, and terrigenous

spores and pollen in the 400–11,000 cpg range, from silty clay to clayey silt lithologies (0.61–1.51%

TOC) that were deposited in an inner neritic setting. From the CTB succession in Wunstorf core, van

Helmond et al. (2015), also using the Lycopodium method, reported dinoflagellate cysts mainly in the

1,000–15,000 cpg range and terrigenous spores and pollen in the 2,500–5,000 cpg range, from

samples with 0.5–3% TOC. Using this study‘s volumetric method, comparable values for organic-

walled phytoplankton were obtained from the CTB succession at Aksudere, Crimea, 3,520–20,400

cpg, from samples with 0.3–7.2% TOC (Dodsworth 2004a). Comparable values have also been

obtained from relatively un-weathered dark coloured mudstones of the Black Band at South Ferriby

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 38: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

and Caistor; average phytoplankton is in the range of 10,000–15,000 cpg with pollen and spores at

200–900 cpg (Supplementary Table I).

Significantly higher concentrations of palynomorphs have been recorded from the dark coloured

mudstones of the Black Band and Bed G elsewhere in Lincolnshire, at Melton Ross (average

phytoplankton ca. 84,000 cpg; pollen and spores ca. 10,000 cpg; Table 2) and Bigby (average

phytoplankton ca. 60,000 cpg; pollen and spores ca. 3,500 cpg). These values may be more

representative than those from South Ferriby and Caistor, given that samples from the latter sections

are apparently more affected by quarry wall weathering. However, the relatively expanded sections at

Melton Ross and Bigby could have been deposited in local topographic troughs in the seafloor (Gaunt

et al. 1992; Wood & Mortimore 1995; Wiese et al. 2009) in which organic matter preservation may

have been enhanced (cf. Jenkyns 1985). Farther north in Yorkshire, average phytoplankton

concentrations from relatively un-weathered dark coloured mudstones of the Black Band and Bed G

are ca. 40,000–50,000 cpg at Market Weighton, Bishop Wilton and East Knapton, with pollen and

spores at ca. 3,500–5,000 cpg; i.e. values relatively comparable to Melton Ross and Bigby. The

highest concentrations of palynomorphs in the region are from sections on the margins of the

Cleveland Basin (e.g. Jeans et al. 1991; Mitchell 2000) at Flixton and Speeton, with average

phytoplankton concentrations of ca. 95,000–100,000 cpg, and pollen and spores ca. 10,000 cpg at

Speeton and ca. 80,000 cpg at Flixton (Supplementary Table I). These regional data appear to be

consistent with the trend of an increase in preservation of organic matter as the depositional

environment deepened from relatively shallow shelf to deeper continental margin (Jenkyns 1985).

Unlike the samples below the Plenus Marls at Eastbourne and Lulworth, which also contain high

concentrations of dinoflagellate cysts (ca. 38,000–75,000 cpg; see above), P. infusorioides is a

relatively minor component of assemblages from dark coloured mudstones in eastern England (Table

4), suggesting that nutrients originating from upwelling might not have been the cause of high

concentrations of dinoflagellate cysts there.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 39: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

The fact that the Wunstorf and Misburg sections are at least ten times as thick as the correlative

eastern England sections, yet show a succession of dinoflagellate cyst and acritarch bio-events

comparable to those at Melton Ross, is consistent with a high degree of stratigraphic condensation in

eastern England. The greater palynomorph concentrations documented from the dark coloured

mudstones in eastern England may partly reflect this, along with increased quality of sea floor

preservation (in a low O2 environment), particularly in potential local topographic troughs in the sea

floor and in deeper areas of the continental shelf. A loss of biogenic carbonate sediment during the

CTB mass extinction interval, and a reduction of siliciclastic input during maximum flooding

associated with the base of the Black Band (eastern England; Wood et al. 1997) and the correlative

base of the Fish Shale (NW Germany; Ernst et al. 1984), may also have contributed to stratigraphic

condensation and enhanced palynomorph concentration.

Palynomorphs are recovered from most levels within the Chalk Group of the Southern Province

(Anglo-Paris Basin; see examples above) and the Transitional Province, e.g. the Trunch borehole (Fig.

2) which has palynological recovery throughout the Upper Cretaceous section apart from a Middle

Cenomanian to Middle Turonian barren interval (Pearce 2010, 2018; Olde et al. 2015a; Pearce et al.

2020). All chalk samples analysed to date from the Cenomanian and Lower Turonian of the Northern

Province, below and above the Black Band, and from the Central Limestone at Melton Ross (this

study), are palynologically barren. This is also the case for chalk samples from the CTB interval in

NW Germany (Marshall & Batten 1988; Prauss 2006; van Helmond et al. 2015). In the North Sea

Basin, poor or patchy palynological recovery from the Chalk Group has led to palynology not being

routinely used in stratigraphic studies of hydrocarbon wells (with exceptions such as Maastrichtian

chalk in the Dan Field, Danish sector; Schiøler & Wilson 1993), whereas it is a primary stratigraphic

discipline in the coeval northerly, siliciclastic Shetland Group, which has consistent rich palynological

recovery (Costa & Davey 1992). In England, a different diagenetic history of Cenomanian and

Turonian chalks in the Northern Province, which are hardened relative to those of the Transitional and

Southern Provinces (Jeans 1980; Jeans et al. 2014), probably contributed to post-depositional loss of

palynomorphs.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 40: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Correlation with Norwegian sector well 35/6-2 S

Ditch cuttings samples from the Grosso exploration well 35/6-2 S in the Norwegian sector, North

Viking Graben (Shetland Group facies; Figs 1, 12) were analysed for palynology by one of us (PD) in

2009. The well has a relatively thick (37 m) Blodøks Formation gamma log profile that probably

correlates with the more expanded outcrop lithological successions discussed in this paper.

A sharp rise in gamma log values marks the base of the Blodøks Formation at 2657 m (log depth).

The LO of consistent / common Litosphaeridium siphoniphorum, and rare Pterodinium

crassimuratum at 2655 m, indicate the probable presence of the pre-Plenus succession. A fall in

gamma log values from 2650.4 m to 2648.6 m may correlate with the Plenus Bed in NW Germany

and Central Limestone / Bed A in eastern England. This is supported by the FO of common

Cyclonephelium compactum–membraniphorum, Eurydinium saxoniense and Alterbidinium daveyi in

the next up-hole sample from 2646 m, in an interval of rising gamma log values from 2648.6 m to

2641.7 m (likely equivalent of eastern England Bed B). The remainder of the Blodøks Formation from

2641.7 m to 2620.0 m contains two gamma log plateaux, separated by a lower gamma interval,

probably correlative with the relatively clay- and organic-rich Fish Shale in NW Germany and Beds C

to H in eastern England. This interpretation is supported by the LO of Carpodinium obliquicostatum

from a cuttings sample near the top of the formation at 2616 m. Common C. compactum–

membraniphorum, E. saxoniense and A. daveyi persist higher than the log pick for the top of the

Blodøks Formation. The highest common occurrence of C. compactum–membraniphorum is often

used as a proxy for the top of the Lower Turonian (Dodsworth & Eldrett 2019), occurring in this well

at 2466 m. Rugubivesiculites rugosus occurs sporadically in eight samples in the Turonian interval,

from 2256 m, but is more consistent in Cenomanian deposits from 2646 m and below.

Conclusions

A sequence of dinoflagellate cyst and acritarch bio-events supports the proposed correlation of the

Melton Ross CTB succession with sections in NW Germany. The top of consistent/common

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 41: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Litosphaeridium siphoniphorum occurs in the lower succession at Melton Ross and the pre-Plenus

beds at Misburg and Wunstorf; the base of abundant Cyclonephelium compactum–membraniphorum

occurs immediately above the Central Limestone at Melton Ross, in lower Bed B, and immediately

above the Plenus Bed at Misburg and Wunstorf; a regional influx of Oligosphaeridium totum, with an

acme in upper Bed B at Melton Ross, occurs at the same level as a record of the latest Cenomanian

zonal ammonite Neocardioceras juddii at Misburg; rare specimens of the dinoflagellate cysts

Adnatosphaeridium tutulosum and Carpodinium obliquicostatum have their range tops in or just

above the Black Band in eastern England, and within the Fish Shale at Misburg and Wunstorf. An

influx of the acanthomorph acritarch Micrhystridium spp. in Bed G at Melton Ross may correlate with

that recorded from the upper parts of the Fish Shale at Wunstorf.

Correlation between eastern and southern England is less straightforward. Carbon isotope data from

Melton Ross and South Ferriby, and the distribution of dinoflagellate cyst bio-events, support the

interpretation that the Black Band is stratigraphically higher than the Plenus Marls. The restriction of

Praeactinocamax plenus (belemnite) to the Plenus Marls and Neocardioceras juddii (ammonite) to

the overlying Ballard Cliff Member are consistent with this. The correlation points to separate

depositional histories and marked differences in siliciclastic input between the Northern and Southern

provinces during post-Plenus Marls times. However, the foraminiferal and calcareous nannofossil data

can alternatively be interpreted in terms of the Black Band being equivalent to the upper part of the

Plenus Marls. Poor preservation and/or recovery of all three microfossil groups in the Ballard Cliff

Member hampers confident assessment of correlative bio-events around the CTB in southern England.

Palynological assemblages from the > 1% TOC samples at Melton Ross contain higher relative

abundances of the P-cysts Eurydinium saxoniense and Bosedinia spp., along with prasinophyte

phycomata, consistent with periodic hydrographically restricted and eutrophic conditions during

deposition of the darker mudstone layers. Exceptionally high concentrations of palynomorphs suggest

enhanced preservation in such a low O2 environment at the sediment-water interface coupled with a

high degree of stratigraphic condensation. This is in agreement with the view of Hart & Leary (1989),

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 42: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

who suggested that the Black Band was deposited during ‗sluggish‘ oceanic conditions that would

have been ideal for the concentration and preservation of organic-rich sediments, with or without high

surface-water productivity.

Systematic Palaeontology

Division DINOFLAGELLATA (Bütschli, 1885) Fensome et al., 1993

Class DINOPHYCEAE Pascher, 1914

Order PERIDINIALES Haeckel, 1894

Suborder PERIDINIINEAE (Autonym)

Family PERIDINIACEAE Ehrenberg, 1831

Genus Alterbidinium Lentin & Williams, 1985

Type species: Alterbidinium “recticorne” (Vozzhennikova, 1967) Harker et al., 1990

Alterbidinium daveyi basionym nov.

Derivation of name. In honour of the palynologist Roger J. Davey.

Designation of holotype. Davey (1970, plate 1, fig. 3). Location, International Yarbo Borehole no.

17, south-east Saskatchewan, Canada (coordinates supplied by Davey, 1969, p. 115, fig. 8, are; ―east

of Regina at Lsd. 1, Sec. 24, Twp. 20, Rg. 33, W1st Meridian‖). Sample depth 254.5 m (835 ft) below

Kelly Bushing, Second White Specks Shale, Colorado Group. The slide is curated at the Natural

History Museum, London (slide/specimen reference number V.51979).

Description. See Davey (1970, p. 338).

Discussion. Davey (1970, pl. 1, figs 3, 4, p. 338) described specimens from Cenomanian deposits in

Saskatchewan that he assigned to Deflandrea (now Subtilisphaera) pirnaensis. He pointed out

differences with the type material of S. pirnaensis, as described by Alberti (1959) from Turonian

deposits in Germany, including a smaller size, length 46 (62.7) 87 μm (compared with 80–106 μm)

and width, 34 (45.5) 63 μm (compared with 58–64 μm), and presence of an archaeopyle in many

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 43: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Saskatchewan specimens but absence of one in the type specimens. It is also noted here that the

holotype of S. pirnaensis (Alberti 1959, pl. 8, fig. 1) has possible pre- and postcingular tabulation that

is absent from the Saskatchewan specimens. Stover & Evitt (1978) proposed a new species, Alterbia

daveyi, based on the specimens illustrated in Davey (1970) but did not designate one of the specimens

as a holotype. The name was therefore not validly published (Fensome et al. 2019). Subsequently, the

genus Alterbia has been considered an illegitimate name and its species have been transferred to the

genus Alterbidinium (Fensome et al. 2019). Alterbidinium “daveyi” has become widely accepted as a

separate but informal species. The designation herein of one of the specimens illustrated in Davey

(1970) as a holotype, gives it formal status.

Alterbidinium daveyi and Eurydinium saxoniense have comparable morphology, and both possess a

single intercalary (2a) plate archaeopyle which is steno- to iso-deltaform. Alterbidinium daveyi may

be slightly larger (cf. length, 52–66 μm, and width, 37–54 μm for E. saxoniense; Marshall & Batten

1988) and appears to have a more strongly-developed apical horn and cingulum than E. saxoniense,

but intergradations may occur. The cingulum in A. daveyi is marked by low ridges that sometimes

possess pustules distally and is occasionally crossed by low ridges delimiting plate boundaries (Davey

1970, p. 338); the latter features are not reported in E. saxoniense. The holotype of A. daveyi appears

to show some anterior dorsal intercalary tabulation (2a and 3a plate boundaries). Dorsal tabulation is

usually restricted to the archaeopyle in E. saxoniense (Marshall & Batten 1988), though our illustrated

specimen (Fig. 5.6) may possess sutures around both 2a and 3a intercalary plates.

The eastern England specimens inspected in this paper and previous studies have been assigned to E.

saxoniense, although the additional presence of A. daveyi cannot be ruled out. The dinoflagellate cyst

distribution charts of Marshall & Batten (1988) indicate an absence of A. daveyi and the common to

abundant occurrence of E. saxoniense from the (post-Plenus) CTB interval in NW Germany.

Conversely, in coeval deposits from North America, A. daveyi (sometimes recorded as S. pirnaensis)

is prominent (Bloch et al. 1999, fig. 23; Dodsworth 2000, 2016; Harris & Tocher 2003; Dodsworth &

Eldrett 2019). In the Shetland Group of Northern Europe, both A. daveyi and E. saxoniense are

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 44: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

common, e.g. in Norwegian well 35/6-2 S, where confident differentiation between the species can be

problematic for some specimens.

Genus Bosedinia He, 1984, emend. Chen et al., 1988, Prauss, 2012c

Type species: Bosedinia granulata (He & Qian, 1979) He, 1984

Bosedinia laevigata (Jiabo, 1978, ex He & Qian, 1979) He, 1984

Figs 7.11, 7.12, 7.13, 7.17, 7.21

Description. Small to intermediate-sized, sub-spheroidal shaped, smooth and thin-walled proximate

autocysts with an ornament of ca. 10–50 sparsely distributed, non-tabular, solid verrucae, each of

which is ca. 1–3 μm in diameter. Tabulation is indicated by an archaeopyle only. A flap-like

operculum, probably involving fused apical and anterior intercalary plates, is often attached.

Inclusions (omphali) are present in some specimens.

Dimensions. Diameter (central body, 12 measured specimens): 35.7 (43.5) 56.0 µm.

Discussion. The type specimens of Bosedinia laevigata (He et al. 2009, pl. 17, figs 11–15), from

lacustrine Neogene deposits in China, are of comparable size (35–60 μm diameter) to the specimens

measured from Melton Ross, and contain a comparable number, size and distribution of small

verrucae ornament. Bosedinia laevigata has been recorded from marine Upper Cretaceous deposits at

Tarfaya, Morocco (Prauss 2012c, 2015) and the Abu Gharadig Basin in Egypt (Ahmed et al. 2020;

Fig. 1), but not previously in Europe.

Stratigraphic range/occurrence. At Melton Ross, Bed C (sample -11.5) to Bed E (sample -8).

Bosedinia cf. sp. 1 of Prauss (2012b)

Figs 7.5, 7.9, 7.10

Discussion. The abundant occurrence of mainly enclosed spheres, containing omphali, in organic-rich

Cenomanian–Coniacian deposits at Tarfaya, has been discussed by Prauss (2012a, b, c, 2015). Prauss

(2012b) provided evidence for their assignment to the genus Bosedinia from occasional specimens

that have an archaeopyle/operculum. Eldrett et al. (2017) documented the abundant occurrence of

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 45: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

comparable specimens at other proto-Atlantic sites (Demerara Rise, DSDP Sites 1260 and 1261; Fig.

1) and within the Western Interior Seaway of the United States (SW Texas and central Colorado).

Specimens from Melton Ross constitute the first record of their presence in Europe. However, there

are issues with consistent identification. While the specimen illustrated in Fig. 7.9 shows a rare

example of discernible anterior intercalary and apical plates being involved in operculum formation,

consistent with Bosedinia, the fully enclosed specimen in Fig. 7.10 could alternatively be interpreted

as an endocyst of Subtilisphaera pontis-mariae. (cf. the specimens in Figs 5.14, 5.15). The two

overlapping specimens in Fig. 7.5, resemble Eyrea nebulosa, which also possesses omphali (cf. plate

11, in Cookson & Eisenack 1971), but may be distinguished by the presence of a kalyptra, when

preserved. In the Melton Ross data, specimens encountered are mainly single enclosed spheres with

omphali and have been assigned to ―Bosedinia cf. sp. 1 / peridinioid endocysts‖ in the supplementary

distribution chart.

Stratigraphic range/occurrence. At Melton Ross, local Bed II (sample -21) to Bed H (sample -1).

Genus Ginginodinium Cookson & Eisenack, 1960, emend. Lentin & Williams, 1976

Type species: Ginginodinium spinulosum Cookson and Eisenack, 1960

Ginginodinium? sp. A of Prauss, 2006, 2012a

Figs 6.3, 6.7

Discussion. In Cenomanian and Turonian deposits from NW Germany (Wunstorf) and Morocco

(Tarfaya), Prauss (2006, 2012a) differentiated specimens that are morphologically close to

Trithyrodinium suspectum but with a more complex archaeopyle, tentatively assigning them to the

genus Ginginodinium. The archaeopyle in Ginginodinium? sp. A involves from one intercalary plate

(2a) to three intercalary plates (1a–3a), often attached, and sutures are usually present between

precingular plates which are attached at the cingular border (3I3Pa). Prauss (2006) noted that

Ginginodinium? sp. A probably includes specimens informally described as ‗Cyst form A‘ by

Marshall & Batten (1988). The type species, Ginginodinium spinulosum (Cookson & Eisenack 1960,

p. 7, pl. 2, fig. 9), and other accepted species, sometimes exhibit accessory sutures on dorsal

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 46: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

precingular plates, in addition to a three plate intercalary archaeopyle (Stover & Evitt 1978). In the

Melton Ross material, consistent differentiation of this taxon from T. suspectum was problematic.

Stratigraphic range/occurrence. At Melton Ross, local Bed I (sample -22) to Bed H (sample -1).

Genus Subtilisphaera Jain & Millepied, 1973, emend. Lentin & Williams, 1976

Type species: Subtilisphaera senegalensis Jain & Millepied, 1973

Subtilisphaera pontis-mariae (Deflandre, 1936) Lentin & Williams, 1976

Figs 5.14, 5.15

Description. Small to intermediate-sized, elongate to ovoidal-shaped, smooth-walled, bicavate to

circumcavate peridinioid dinoflagellate cysts with pericoels developed into a sub-conical apical horn

and one similar antapical horn which is positioned asymmetrically. The epicyst is larger than the

hypocyst. The cingulum is delimited by two low ridges. An archaeopyle has not been observed. The

endocyst is sub-spherical and slightly thicker walled than the periphragm. The horns are

approximately 1/5 to 1/4 of the diameter of the endocyst.

Dimensions (seven measured specimens). Length 47.3 (52.7) 57.0 µm; width 32.6 (36.0) 42.2 µm.

Discussion. The Subtilisphaera specimens encountered in the present study have comparable

morphology, with some variation in the degree of cavation and the length of horns. They are recorded

here as Subtilisphaera pontis-mariae. The holotype (Deflandre 1936, pl. 2, fig.7) is closely

comparable to the Melton Ross specimen illustrated in Fig. 5.14. Some subsequently illustrated

specimens (e.g. Davey 1970, pl.1, figs 10–11; Dodsworth 2016, pl. 1, fig. 7) differ slightly in having

longer horns, ca. 1/2 to 1/3 the diameter of the endocyst. While Dodsworth (1996, 2000) used

Subtilisphaera spp. for eastern and southern England specimens from the CTB interval, other workers

in eastern England (Marshall & Batten 1988; Hart et al. 1993), southern England (FitzPatrick 1995;

Pearce et al. 2009) and northern France (Foucher 1979) have considered the range of morphological

variation of Subtilisphaera specimens to fall within acceptable intra-specific limits for S. pontis-

mariae in material of this age. However, Foucher (1980) assigned latest Cenomanian specimens to S.

cf. pontis-mariae in the Boulonnais region of northern France.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 47: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Stratigraphic range/occurrence. At Melton Ross, local Bed I (sample -22) to Bed H (sample -1).

Genus Trithyrodinium Drugg, 1967, emend. Davey, 1969, Lentin & Williams, 1976, Marheinecke,

1992

Type species: Trithyrodinium evittii Drugg, 1967

Trithyrodinium maculatum sp. nov.

Figs 7.14, 7.15, 7.18, 7.19

Derivation of name. From maculate, meaning spotted or stained, with reference to the distinctive

surface markings.

Designation of holotype. Fig. 7.18. Sample MR3. Slide MR3(A). Coordinates S32/3. Location,

Melton Ross Quarry, Lincolnshire, UK. Stratum, Bed F, Flixton Member, Welton Chalk Formation.

The specimen is curated in the MPK collection of the British Geological Survey, Keyworth,

Nottingham, U.K., specimen number MPK 14662.

Diagnosis. A species of Trithyrodinium possessing an ornament of ring-shaped indentations on the

endophragm and periphragm.

Description. Small to intermediate-sized, spheroidal to ovoidal-shaped, circumcavate, peridinioid

dinoflagellate cysts. The endophragm (ca. 1 μm thick) and periphragm (ca. 0.5 μm thick) are smooth

to finely-granular, both possessing regular ornament elements that are circular to sub-circular, ring-

shaped indentations (ca. 3–6 μm diameter), which are at least half the thickness of the surrounding

wall areas. Cavation is ca. 2–3 μm wide at the lateral margins and up to 10 μm in the antapical region.

An archaeopyle involving three anterior intercalary plates, type 3I(1–3a), is formed in the endocyst, with

operculum plates attached or detached. The 2a intercalary plate is isodeltaform, hexa-type (Fig. 7.19;

cf. text-figure 6 in Bujak & Davies 1983). The periarchaeopyle type has not been confirmed. There is

no evidence of tabulation, other than that indicated by the archaeopyle.

Dimensions (length x width of the central endocyst). Holotype, 39.6 x 42.5 µm. Other specimens (5

measured): length 38.0 (50.0) 66.9 µm, width 39.5 (45.6) 55.3 µm.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 48: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Discussion. Only six specimens of this taxon have been recorded to date. However, its distinctive

morphology makes it easy to identify and warrants the erection of a formal species. The endocyst

archaeopyle type is consistent with the genus Trithyrodinium but an outer wall layer (periphragm) has

only been observed on two specimens, including the holotype. The circular indentations on the cyst

walls are ring-shaped, as opposed to the whole area within the circle being thinned. The ring shapes

are reminiscent of those attributed to impressions made by coccoliths on the surface of specimens of

an alga (Campenia sp.) by Prauss (2012b, figs 10A, 11H). However, the T. maculatum specimens

occur in assemblages of abundant dinoflagellate cysts, with other taxa not exhibiting comparable

markings. The markings are a key identifying feature but could be a preservation artefact. This type of

surface ornament has not previously been reported on other species of P-cysts, including those

belonging to the genus Trithyrodinium. In the G-cyst species Apteodinium maculatum, comparable

ring-shaped thinning surrounds small thickened circular areas (Eisenack & Cookson 1960).

Stratigraphic range/occurrence. At Melton Ross, Bed B (sample -12) to Bed F (sample -3).

Trithyrodinium? sp. A (this paper)

Figs 7.1, 7.2

Description. Small to intermediate-sized, sub-polygonal to peridinioid-shaped, smooth and thin-

walled cysts that lack ornament. Peridinioid tabulation is indicated by sutures or breaks between some

plates on the epicyst, indicating three anterior intercalary plates. An archaeopyle has not been

observed. A periphragm has not been observed. There is no evidence of tabulation, other than

indicated by the sutures.

Dimensions (length x width). First specimen, 46.5 x 40.7 µm; second specimen, 54.3 x 50.0 µm.

Discussion. Prauss (2012a, fig. 15) illustrated an unpublished smooth, thin-walled species of

Trithyrodinium at Tarfaya that usually has sutures around three intercalary plates and rarely possesses

a periphragm (Prauss, 2012a, fig. 15J). It is uncertain whether this taxon is the same as the Melton

Ross specimens described here.

Stratigraphic range/occurrence. At Melton Ross, Bed C (sample -11) to Bed G (sample -2).

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 49: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Order GONYAULACALES Taylor, 1980

Suborder GONYAULACINEAE (Autonym)

Family GONYAULACACEAE Lindemann, 1928

Genus Dissiliodinium Drugg, 1978

Type species: Dissiliodinium globulus Drugg, 1978

?Dissiliodinium globulus Drugg, 1978

Figs 7.4, 7.8

Description. Intermediate-sized, spheroidal-shaped, smooth and thin-walled proximate autocysts

lacking ornament. Tabulation is indicated by sutures between precingular plates on the epicyst.

Archaeopyle formation involves one or more precingular plate. There is no evidence of tabulation on

the hypocyst.

Dimensions (four specimens measured). Diameter: 57.0 (60.7) 65 µm.

Discussion. The observed features of the four available specimens allow tentative assignment to

Dissiliodinium globulus. The taxon has been documented from the Upper Jurassic of Central Europe

(Drugg 1978) and Northern Europe (e.g. Bailey et al. 1997) and is present in higher proportions

within the Lower Cretaceous of SW Morocco (Below 1981). Prauss (2012b, fig 9a–i) reported

specimens from around the CTB at Tarfaya.

Stratigraphic range/occurrence. At Melton Ross, Bed E (sample -8 only).

Genus Leptodinium Klement, 1960, emend. Stover & Evitt, 1978, Sarjeant, 1982

Type species: Leptodinium subtile Klement, 1960

Leptodinium? aff. delicatum (this paper)

Fig. 6.12

Description. Intermediate-sized, smooth and thin-walled autocysts with sub-polygonal shape and

septae (entire; ca. 1.5 to 3.5 μm high) which probably correspond to tabulation. The cingulum appears

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 50: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

to have strongly offset ventral ends. An archaeopyle has not been observed on the two specimens

recorded.

Dimensions. Diameter: illustrated specimen, 61.8 µm; second specimen, 60.8 µm.

Discussion. In Leptodinium? delicatum, the wall is very thin (less than 0.5 μm thick) and only attains

a thickness of 0.5 μm when forming plate boundaries (Davey 1969, p. 123).The Melton Ross

specimens resemble L.? delicatum but differ in possessing sutural septae. The detailed plate formula is

not yet determined. The probable offset sulcus is consistent with the genus Leptodinium though the

presence of septae is more compatible with the genus Impagidinium (Stover & Evitt 1978).

Stratigraphic range/occurrence. At Melton Ross, Bed C (samples -11 and -11.5).

Genus Oligosphaeridium Davey & Williams, 1966, emend. Davey, 1982

Type species: Oligosphaeridium complex (White, 1842) Davey & Williams, 1966

Oligosphaeridium totum Brideaux, 1971

Figs 5.9, 5.10, 8.9

Discussion. The published taxon Oligosphaeridium totum is considered here to accommodate a

skolochorate dinoflagellate cyst from NW Germany and the North Sea that was informally assigned

by Marshall & Batten (1988) to Litosphaeridium sp. A. The latter name was adopted in subsequent

studies of the CTB interval in eastern England (Hart et al. 1993; Dodsworth 1996), southern England

(FitzPatrick 1995) and Poland (Dodsworth 2004b). The processes of O. totum are tubular, sometimes

slightly narrower medially, with apices usually flaring, buccinate and possessing an entire margin

(Brideaux 1971). Two subspecies are differentiated, based mainly on process length relative to central

body, O. totum subsp. totum (relatively long processes, approximately two-thirds to one central body

diameter in length; comparable to most specimens encountered in this study, e.g. Figs 5.9, 8.9) and O.

totum subsp. minus (relatively short processes, approximately one half of the central body diameter;

cf. Fig. 5.10). The Melton Ross specimens have the same process formula as those described by

Brideaux (1971) for O. totum and by Marshall & Batten (1988) for Litosphaeridium sp. A; 4I, 6II, 0c,

5III, 1p, 1IIII, 1s. This is a process formula common to all species of Oligosphaeridium and some

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 51: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

species of Litosphaeridium (Lucas-Clark 1984). Litosphaeridium differs from Oligosphaeridium in

having dome-shaped processes that are typically not expanded distally (Stover & Evitt 1978). The

type material of O. totum has a slightly thicker, more scabrate central body wall than the smooth walls

observed on Melton Ross specimens and those described and illustrated from NW Germany (Marshall

& Batten 1988, p. 92, pl. I, fig. 9; Marshall 1983, pl. 13, figs 16, 18). However, Singh (1971) reported

a smooth cyst wall on Oligosphaeridium diastema, a junior synonym of O. totum (as agreed by Singh

and Brideaux; Brideaux & McIntyre 1975, p. 29). Some variation in thickness and ornament of the

central body may therefore be attributed to intra-specific variation.

Stratigraphic range/occurrence. At Melton Ross, local Bed VII (sample -17) to Bed F (sample -3).

Genus Pterodinium Eisenack, 1958, emend. Yun, 1981, Sarjeant, 1985

Type species: Pterodinium aliferum Eisenack, 1958

Pterodinium crassimuratum (Davey & Williams, 1966) Thurow et al., 1988

Fig. 5.7

Discussion. Kjellström (1973) and Pavlishina (1990) considered Pterodinium? pterotum to be a senior

synonym of Pterodinium crassimuratum and Pterodinium cingulatum subsp. polygonale. In

Dodsworth & Eldrett (2019, plate 1, fig.4), specimens encountered were assigned on the basis of this

proposed synonymy to P.? pterotum. However, subsequent inspection of the holotype photographs

and original descriptions for the three taxa indicates that our specimens, both from the United States

and Europe, are closest to P. crassimuratum and P. cingulatum polygonale, having thickened intra-

plate areas and ―a linear depressed area on each side of the ledges [septae] thereby separating a more

elevated central plate area from the ledges‖ (Clarke & Verdier 1967, p. 47). We agree with Clarke et

al. (1968, p. 181) that the latter subspecies is a junior synonym of P. crassimuratum. However,

Cookson & Eisenack (1958, p. 50) make no reference to comparable thickened intra-plate areas in P.?

pterotum, and none are visible on their holotype photograph (plate 11, fig. 7). We therefore conclude

that P.? pterotum is not a senior synonym of P. crassimuratum.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 52: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Family AREOLIGERACEAE Evitt, 1963

Genus Aptea Eisenack, 1958

Type species: Aptea polymorpha Eisenack, 1958

Aptea? spongireticulata (Prössl, 1990 ex Prössl, 1992) Fensome et al., 2019

Figs 6.1, 6.2

Description. Intermediate-sized, lenticular-shaped cysts with rounded or flattened apical and

antapical areas, lacking well-developed horns. The autophragm is smooth to lightly pitted and is

covered with a coarse reticulum. Rod-like structures with expanded to bifurcating tips support the

reticulum muri and are ca. 6–10 μm long. The muri are fibrous or finely fenestrate. The lacuna

between the muri vary in size, from ca. 5–13 μm across. A slight indentation in the reticulum in the

antapical region on one specimen (Fig. 6.1) may represent a gap between antapical horns. Tabulation

is indicated by an apical archaeopyle only.

Dimensions (five measured specimens, length x width without operculum). Length 61.0 (66.9) 73.0

µm; width 83.8 (86.3) 89.0 µm.

Discussion. The Melton Ross specimens conform to the type material of A.? spongireticulata,

illustrated from NW Germany by Prössl (1990). Canningia macroreticulata differs in having smaller

lacuna between its muri of ca. 2–7 μm and has shorter (ca. 2 μm high), entire muri without

fibrous/fenestrate structure (Lebedeva, in Ilyina et al. 1994). The Melton Ross specimens may be the

first published record of A.? spongireticulata outside Germany.

Stratigraphic range/occurrence. At Melton Ross, Bed B (sample -13) to Bed F (sample -5). In NW

Germany, Prössl (1990) reported an Albian to Middle Turonian range for A.? spongireticulata.

Genus Canninginopsis Cookson & Eisenack, 1962

Type species: Canninginopsis denticulata Cookson & Eisenack, 1962

Canninginopsis? lindseyensis sp. nov.

Figs 6.8, 6.9, 7.7

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 53: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Derivation of name. From Lindsey, a former Anglo-Saxon kingdom and current district of

Lincolnshire.

Designation of holotype. Figures 6.8, 6.9. Sample MR11. Slide MR11(B). Coordinates P44/3.

Location, Melton Ross Quarry, Lincolnshire, UK. Stratum, Bed C, Flixton Member, Welton Chalk

Formation. The specimen is curated in the MPK collection of the British Geological Survey,

Keyworth, Nottingham, U.K., specimen number MPK 14663.

Diagnosis. An areoligeracean dinoflagellate cyst possessing tabular to penitabular thickenings that

form the bases of thinner, entire septae.

Description. Small to intermediate-sized, lenticular-shaped autocysts with differential development of

a larger left antapical horn. A sulcal notch is present. The cyst wall is ca. 1 μm thick and has a smooth

to slightly reticulate surface. Tabulation is indicated by an apical archaeopyle (operculum detached)

and tabular to penitabular thickenings, ca. 1–2 μm high and 1–3 μm wide, that form the bases of

thinner septae. The septae are ca. 2–5 μm high and 1 μm wide, entire, smooth to slightly reticulate and

often exhibit some longitudinal folding (‗creases‘). Proximal thickenings/septae are adjacent (tabular)

or separated by gaps of up to 4 μm (penitabular). The gonyaulacoid tabulation pattern is (operculum

not observed); 6II, 6c, 5–6III, ?1P, 1IIII, ?s.

Dimensions (length x width, without operculum). Holotype: length 44.4 µm, width 54.1 µm. Other

specimens (2 measured): length 43.0 (46.3) 49.6 µm, width 58.9 (64.3) 69.6 µm.

Discussion. The shape, presence of a sulcal notch and a probable small postcingular 1III plate (Fig.

6.8) indicate an areoligeracean affinity (cf. Evitt 1985, text-figs 10.2, 10.6). Strongly developed

tabulation is consistent with the genus Canninginopsis. Although the type species and most other

accepted species often possess discontinuous ornament elements reflecting tabulation such as spines

or grana, continuous septae are present on the species Canninginopsis maastrichtiensis from the

Maastrichtian of Belgium and Netherlands (Slimani 1994). However, the development of gaps

between septae, giving rise to penitabular ornament, is inconsistent with Canninginopsis and warrants

questionable assignment to the genus. Penitabulation has been noted in the related genus Canningia,

e.g. Canningia transitoria (Stover & Helby 1987, fig. 4) but its septae are coarsely perforate.

Schematophora possesses entire to occasionally perforate penitabular ridges but has a spherical, non-

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 54: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

areoligeracean shape and lacks any tabulation in the cingulum region (Deflandre & Cookson 1955;

Stover & Evitt 1978). On the holotype of C.? lindseyensis, the dorsal precingular plate (3II) appears to

be absent but this could be due to specimen damage. Likewise, a precingular 6II plate may be

damaged. A second specimen with less well-developed or preserved septae (Fig. 7.7) indicates that six

precingular plates may usually be present. The highly distinctive morphology warrants erection of a

formal species, although it is acknowledged that doing so is unusual for just three specimens.

Recovery and inspection of further specimens is required to fully elucidate the taxon and its generic

assignment.

Stratigraphic range/occurrence. At Melton Ross, Bed C (sample -11 only).

Dinoflagellate cysts of uncertain supra-generic affinity

Dinoflagellate? type D of Ioannides, 1986

Figs 6.13, 6.14

Dimensions (27 specimens). Diameter (autophragm central body): 45.7 (60.7) 75.2 µm.

Discussion. Ioannides (1986, p. 42, pl. 25, figs 1– 4, 6) described comparable though slightly larger

taxa from Upper Cretaceous (Santonian–Maastrichtian) deposits in Arctic Canada (size range 70–79

μm length and 70–87 μm width; 11 measured specimens). He noted that ornament varied from

granular to minutely verrucate/rugulate, and that, ―in some specimens, a number of ‗slits‘ have been

observed along the autophragm. Although a weak plate attachment may be suggested, no definite

regular pattern has been determined. Occasionally, six paraplates (precingular or postcingular) may be

visualised‖, as shown here in Figs 6.13, 6.14. A similar but larger form (140 x 119 μm), surrounded

by a kalyptra, was described from Western Australia by Cookson & Eisenack (1971, p. 223, plate 11,

fig. 1) as Eyrea sp. Occasional Melton Ross specimens possess a kalyptra, comparable to Eyrea sp.,

but most specimens lack an outer layer. Comparable specimens from Poland were assigned to Eyrea?

spp. by Dodsworth (2004b). In poorly preserved material from eastern England, it can be difficult to

distinguish specimens of Dinoflagellate? type D from degraded and/or broken specimens of

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 55: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Sentusidinium ringnesiorum (cf. Fig. 7.20) and the granular endocysts of Trithyrodinium suspectum

and Ginginodinium? sp. A.

Stratigraphic range/occurrence. At Melton Ross, local Bed I to Bed H. A broad Late Cretaceous

range is suggested by the records of Ioannides (1986) and Cookson & Eisenack (1971).

Acknowledgements

Richard Stansfield, CEO of the Singleton Birch Group, is thanked for allowing access to Melton Ross

Quarry in 1997 and encouraging publication of new data. In 2019, quarry staff members Ben Hyde

and Dave Alliss provided an update on the sampled sites. PD undertook palynological laboratory

processing of the Melton Ross samples at the former Aberdeen office of Robertson Research

International Ltd while employed there as a biostratigrapher in 1997. The company and its former

management, Keith L. Marshall and Grant Heath, are thanked for providing the laboratory facilities.

KLM is also thanked for permission to refer to his unpublished PhD thesis. Tim Absalom and James

Quinn of Plymouth University drafted Figure 2. Jennie Houlgrave of RPS Ichron prepared Figure 3

and assisted with Figure 4. Natalia Lebedeva, Klaus Prössl and staff at the Geological Society library

assisted with gathering literature. Organic geochemistry and stable isotope work were commissioned

by StrataSolve Ltd in 2019. TOC and Rock-Eval pyrolysis analyses were undertaken by Adam

Fermor and Patrick Barnard of APT (U.K.) Ltd. Stable isotope analyses were conducted by

Christopher Day, Department of Earth Sciences, Oxford University. StatoilHydro AS (now Equinor

Energy AS) are thanked for permission in 2010 to publish the CTB section from their Norwegian well

35/6-2 S. James P. Fenton and Martin A. Pearce (reviewers), and Stewart G. Molyneux (editor) made

helpful suggestions for improving the manuscript.

References

Ahmed, M., Tahoun, S., Gentzis, T. & Elewa, A.M.T. 2020. The marine palynology of the Upper

Cretaceous Abu Roash ‗A‘ Member in the BED 2-3 borehole, Abu Gharadig Basin, Egypt.

Palynology, 44, 167–186.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 56: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Alberti, G.A. 1959. Zur Kenntnis der Gattung Deflandrea Eisenack (Dinoflag.) in der Kreide und im

Alttertiär Nord- und Mitteldeutschlands. Mitteilungen aus dem Geologischen Staatsinstitut in

Hamburg, 28, 93–105.

Arthur, M.A., Bottjer, D.J., Dean, W.E., Fischer, A.G., Hattin, D.E., Kauffman, E.G., Pratt, L.M. &

Scholle, P.A. 1986. Rhythmic bedding in Upper Cretaceous pelagic carbonate sequences: Varying

sedimentary response to climatic forcing. Geology, 14, 153–156.

Aurisano, R.W. 1989. Upper Cretaceous dinoflagellate biostratigraphy of the subsurface Atlantic

coastal plain of New Jersey and Delaware. Palynology, 13, 143–179.

Backhouse, J. 2006. Albian (Lower Cretaceous) dinoflagellate cyst biostratigraphy of the Lower

Gearle Siltstone, southern Carnarvon Basin, Western Australia. Palynology, 30, 43–68.

Bailey, D., Milner, P. & Varney, T. 1997. Some dinoflagellate cysts from the Kimmeridge Clay

Formation in North Yorkshire and Dorset, UK. Proceedings of the Yorkshire Geological Society, 51,

235–243.

Below, R. 1981. Dinoflagellaten-Zysten aus dem Oberen Hauterive bis Unteren Cenoman Süd-West

Marokkos. Palaeontographica, Abteilung B, 176, 1–145.

Bengtson, P., Cobban, W.A., Dodsworth, P., Gale, A.S., Kennedy, W.J., Lamolda, M.A., Matsumoto,

T., Reyment, R.A., Seibertz, E. & Tröger, K.-A. 1996. The Turonian stage and substage boundaries.

In: Rawson, P.F. et al. (eds) Proceedings, "Second International Symposium on Cretaceous Stage

Boundaries", Brussels 8–16 September 1995. Bulletin de l'Institut Royal des Sciences Naturelles de

Belgique–Sciences de la Terre, 66, 69–79.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 57: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Bloch, J.D., Schroder-Adams, C.J., Leckie, D.A., Craig, J. & McIntyre, D.J. 1999. Sedimentology,

micropaleontology, geochemistry, and hydrocarbon potential of shale from the Cretaceous Lower

Colorado Group in western Canada. Geological Survey of Canada, Bulletin 531. ISBN 0-660-17729-

3.

Boulila, S., Charbonnier, G., Spangenberg, J.E., Gardin, S., Galbrun, B., Briard, J. & Le Callonnec, L.

2020 (online). Unravelling short- and long-term carbon cycle variations during the Oceanic Anoxic

Event 2 from the Paris Basin Chalk. Global and Planetary Change, 186,

Doi.org/10.1016/j.gloplacha.2020.103126. hal-02459697.

Bowman, A.R. & Bralower, T.J. 2005. Paleoceanographic significance of high-resolution carbon

isotope records across the Cenomanian–Turonian boundary in the Western Interior and New Jersey

coastal plain, USA. Marine Geology, 217, 305–321.

Bralower, T.J. 1988. Calcareous nannofossil biostratigraphy and assemblages at the Cenomanian–

Turonian boundary interval: implications for the origin and timing of oceanic anoxia.

Paleoceanography, 3, 275–316.

Bralower, T.J. & Bergen, J.A. 1998. Cenomanian–Santonian calcareous nannofossil biostratigraphy of

a transect of cores drilled across the western interior seaway. In: Dean, W.E. & Arthur, M.A. (eds)

Stratigraphy and paleoenvironments of the Cretaceous western interior seaway, U.S.A. SEPM,

Concepts in Sedimentology and Paleontology, 6, 59–77.

Brett, C.E., Pratt, B.R. & Landing, E. 2018. North American Commission on Stratigraphic

Nomenclature Note 68 – Application for addition of submembers to the North American stratigraphic

code: A case for formalizing lithostratigraphic units of intermediate rank. Stratigraphy, 15, 103–108.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 58: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Brideaux, W.W. 1971. Palynology of the Lower Colorado Group, central Alberta, Canada. I.

Introductory remarks. Geology and microplankton studies. Palaeontographica, Abteilung B, 135, 53–

114, pl.21–30.

Brideaux, W.W. & McIntyre, D.J. 1975. Miospores and microplankton from Aptian–Albian rocks

along Horton River, District of Mackenzie. Geological Survey of Canada, Bulletin, 252, 1–85, pl.1–

14.

Brinkhuis, H. & Zachariasse, W.J. 1988. Dinoflagellate cysts, sea level changes and planktonic

foraminifers across the Cretaceous–Tertiary boundary at El Haria, N.W. Tunisia. Marine

Micropaleontology, 13, 153–191.

Bujak, J.P. & Davies, E.H. 1983. Modern and fossil Peridiniineae. AASP Contribution Series No. 13.

ISSN 0160-8843.

Burgess, J.D. 1971. Palynological Interpretation of frontier environments in central Wyoming.

Geoscience and Man, 3, 69–82.

Burnhill, T.J. & Ramsay, W.V. 1981. Mid-Cretaceous Palaeontology and Stratigraphy, Central North

Sea. In: Illing, L.V. & Hobson, G.D. (eds) Petroleum Geology of the Continental Shelf of North-West

Europe. Institute of Petroleum, London, 245–254.

Caron, M., Dall‘Agnolo, S., Accarie, H., Barrera, E., Kauffman, E.G., Amédro, F. & Robaszynski, F.

2005. High-resolution stratigraphy of the Cenomanian–Turonian boundary interval at Pueblo (USA)

and Wadi Bahloul (Tunisia): Stable isotope and bio-events correlation. Geobios, 39, 171–200.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 59: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Charbonnier, G., Boulila, S., Spangenberg, J.E., Adatte, T., Föllmi, K.B. & Laskar, J. 2018. Obliquity

pacing of the hydrological cycle during the Oceanic Anoxic Event 2. Earth and Planetary Science

Letters, 499, 266–277.

Clarke, R.F.A., Davey, R.J., Sarjeant, W.A.S. & Verdier, J.-P. 1968. A note on the nomenclature of

some Upper Cretaceous and Eocene dinoflagellate taxa. Taxon, 17, 181–183.

Clarke, R.F.A. & Verdier, J.-P. 1967. An investigation of microplankton assemblages from the Chalk

of Isle of Wight, England. Verhandelingen der Koninklijke Nederlandsche Akademie van

Wetenschappen, Afdeling Natuurkunde, Eerste Reeks, 24, 1–96.

Clarkson, M.O., Stirling, C.H., Jenkyns, H.C., Dickson, A.J., Porcelli, D., Moy, C.M., Pogge von

Strandmann, P.A.E., Cooke, I.R. & Lenton, T.M. 2018. Uranium isotope evidence for two episodes of

deoxygenation during Oceanic Anoxic Event 2. Proceedings of the National Academy of Sciences,

115, 2918–2923.

Cobban, W.A. 1993. Diversity and distribution of Late Cretaceous ammonites, Western Interior, U.S.

In: Kauffman, E.G. & Caldwell, W.G.E. (eds) Evolution of the Western Interior Basin. Geological

Association of Canada Special Paper, 39, St Johns, NL, Canada, 435–452.

Cookson, I.C. & Eisenack, A. 1958. Microplankton from Australian and New Guinea Upper

Mesozoic sediments. Proceedings of the Royal Society of Victoria, 70, 19–79.

Cookson, I.C. & Eisenack, A. 1960. Microplankton from Australian Cretaceous sediments.

Micropaleontology, 6, 1–18.

Cookson, I.C. & Eisenack, A. 1962. Additional microplankton from Australian Cretaceous sediments.

Micropalaeontology, 8, 485–507.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 60: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Cookson, I.C. & Eisenack, A. 1971. Cretaceous microplankton from Eyre No.1 Bore Core 20,

Western Australia. Proceedings of the Royal Society of Victoria, 84, 217–226.

Coplen, T. B. 1996. New guidelines for reporting stable hydrogen, carbon, and oxygen isotope-ratio

data. Geochimica et Cosmochimica Acta, 60, 3359–3360.

Costa, L.I. & Davey, R.J. 1992. Dinoflagellate cysts of the Cretaceous System. In: Powell, A.J. (ed) A

Stratigraphic Index of Dinoflagellate Cysts. Chapman & Hall, London, 99–154.

Courtinat, B, Crumière, J.-P., Méon, H. & Schaaf, A. 1991. Les associations de kystes de

dinoflagellés du Cénomanien-Turonien de Vergons (Bassin Vocontien France). Geobios, 24, 649–

666.

Courtinat, B. 1993. The significance of palynofacies fluctuations in the Greenhorn Formation

(Cenomanian–Turonian) of the Western Interior basin, USA. Marine Micropaleontology, 21, 249–

257.

Crittenden, S., Cole, J. & Harlow, C. 1991. The Early to ‗Middle‘ Cretaceous lithostratigraphy of the

Central North Sea (UK sector). Journal of Petroleum Geology, 14, 387–416.

Davey, R.J. 1969. Non-calcareous microplankton from the Cenomanian of England, northern France

and North America, Part I. Bulletin of the British Museum (Natural History) Geology, 17(3), 107–180.

Davey, R.J. 1970. Non-calcareous microplankton from the Cenomanian of England, northern France

and North America, Part II. Bulletin of the British Museum (Natural History) Geology, 18(8), 333–

397.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 61: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Davey, R.J. & Verdier, J.-P. 1976. A review of certain non-tabulate Cretaceous hystrichosphaerid

dinocysts. Review of Palaeobotany and Palynology, 22, 307–335.

Davey, R.J. & Williams, G.L. 1966. IV. The genera Hystrichosphaera and Achomosphaera. In:

Davey, R.J., Downie, C., Sarjeant, W.A.S. & Williams, G.L. Studies on Mesozoic and Cainozoic

dinoflagellate cysts. British Museum (Natural History) Geology, Bulletin, Supplement, 3, 28–52.

Day, C.C. & Henderson, G.M. 2011. Oxygen isotopes in calcite grown under cave-analogue

conditions. Geochimica et Cosmochimica Acta, 75, 3956–3972.

Deegan, C.E. & Scull, B.J. 1977. A standard lithostratigraphic nomenclature for the central and

northern North Sea. Institute of Geological Sciences, London, Report 77/25.

Deflandre, G. 1936. Microfossiles des silex crétacés. Première partie. Généralités. Flagellés. Annales

de paléontologie, 25, 151–191; pl.1–10.

Deflandre, G. & Cookson, I.C. 1955. Fossil microplankton from Australian Late Mesozoic and

Tertiary sediments. Australian Journal of Marine and Freshwater Research, 6, 242–313.

Dodsworth, P. 1995. A note of caution concerning the application of quantitative palynological data

from oxidized preparations. Journal of Micropalaeontology, 14, 6.

Dodsworth, P. 1996. Stratigraphy, microfossils and depositional environments of the lowermost part

of the Welton Chalk Formation (late Cenomanian to early Turonian, Cretaceous) in eastern England.

Proceedings of the Yorkshire Geological Society, 51, 45–64.

Dodsworth, P. 2000. Trans-Atlantic dinoflagellate cyst stratigraphy across the Cenomanian–Turonian

(Cretaceous) Stage boundary. Journal of Micropalaeontology, 19, 69–84.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 62: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Dodsworth, P. 2004a. The palynology of the Cenomanian–Turonian (Cretaceous) boundary

succession at Aksudere in Crimea, Ukraine. Palynology, 28, 129–141.

Dodsworth, P. 2004b. The distribution of dinoflagellate cysts across a Late Cenomanian carbon

isotope (δ13C) anomaly in the Pulawy borehole, central Poland. Journal of Micropalaeontology, 23,

77–80.

Dodsworth, P. 2016. Palynostratigraphy and palaeoenvironments of the Eagle Ford Group (Upper

Cretaceous) at the Lozier Canyon outcrop reference section, west Texas, USA. Palynology, 40, 357–

378.

Dodsworth, P. & Eldrett, J.S. 2019. A dinoflagellate cyst zonation of the Cenomanian and Turonian

(Upper Cretaceous) in the Western Interior, United States. Palynology, 43, 701–723.

Downie, C., Hussain, M.A. & Williams, G.L. 1971. Dinoflagellate cyst and acritarch associations in

the Paleogene of southeast England. Geoscience and Man, 3, 29–35.

Drugg, W.S. 1978. Some Jurassic dinoflagellate cysts from England, France and Germany.

Palaeontographica, Abteilung B, 168, 61–79, pl.1–8.

Duane, A.M. 1992. Palynological investigations of Cenomanian chalks and marls from England. PhD

thesis (unpublished), Polytechnic South West (now Plymouth University).

Du Vivier, A. D. C., Selby, D., Condon, D. J., Takashima, R. & Nishi, H. 2015. Pacific 187Os/188Os

isotope chemistry and U–Pb geochronology: Synchroneity of global Os isotope change across OAE 2.

Earth and Planetary Science Letters, 428, 204–216.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 63: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Eisenack, A. & Cookson, I.C. 1960. Microplankton from Australian Lower Cretaceous sediments.

Proceedings of the Royal Society of Victoria, 72, 1–11; pl.1–3.

Eldrett, J.S., Minisini, D. & Bergman, S.C. 2014. Decoupling of the carbon cycle during Oceanic

Anoxic Event 2. Geology, 42(7), 567–570.

Eldrett, J.S., Ma, C., Bergman, S.C., Lutz, B., Gregory, J.G., Dodsworth, P., Phipps, M., Hardas, P.,

Minisini, D., Ozkan, A., Ramezani, J., Bowring, S.A., Kamo, S.L., Ferguson, K., Macaulay, C. &

Kelly, A. 2015a. An astronomically calibrated stratigraphy of the Cenomanian, Turonian and earliest

Coniacian from the Cretaceous Western Interior Seaway, USA: Implications for global

chronostratigraphy. Cretaceous Research, 56, 316–344.

Eldrett, J.S., Ma, C., Bergman, S.C., Ozkan, A., Minisini, D., Lutz, B., Jackett, S.J., Macaulay, C. &

Kelly, A.E. 2015b. Origin of limestone–marlstone cycles: astronomic forcing of organic-rich

sedimentary rocks from the Cenomanian to early Coniacian of the Cretaceous Western Interior

Seaway, USA. Earth and Planetary Science Letters, 423, 98–113.

Eldrett, J.S., Dodsworth, P., Bergman, S.C., Wright, M. & Minisini, D. 2017. Water-mass evolution in

the Cretaceous Western Interior Seaway of North America and equatorial Atlantic. Climate of the

Past, 13, 855–878.

Ernst, G., Schmid, F. & Seibertz, E. 1983. Event-Stratigraphie im Cenoman und Turon von NW-

Deutschland. Zitteliana, 10, 531–554.

Ernst, G., Wood, C.J. & Hilbrecht, H.H. 1984. The Cenomanian–Turonian boundary problem in NW-

Germany with comments on the north–south correlation to the Regensburg Area. Bulletin of the

Geological Society of Denmark, 33, 103–113.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 64: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Ernst, R.E. & Youbi, N. 2017. How Large Igneous Provinces affect global climate, sometimes cause

mass extinctions, and represent natural markers in the geological record. Palaeogeography,

Palaeoclimatology, Palaeoecology, 478, 30–52.

Evitt, W.R. 1985. Sporopollenin dinoflagellate cysts. Their morphology and interpretation. American

Association of Stratigraphic Palynologists Foundation, Dallas. 333 pp. ISBN 0-931871-00-X.

Farrimond, P., Eglington, G., Brassell, S.C. & Jenkyns, H.C. 1990. The Cenomanian/Turonian anoxic

event in Europe: an organic geochemical study. Marine and Petroleum Geology, 7, 75–89.

Fensome, R.A., Nøhr-Hansen, H. & Williams, G.L. 2016. Cretaceous and Cenozoic dinoflagellate

cysts and other palynomorphs from the western and eastern margins of the Labrador–Baffin Seaway.

Geological Survey of Denmark and Greenland Bulletin, 36, 143 pp.

Fensome, R.A., Williams, G.L. & McRae, R.A. 2019. The Lentin and Williams index of fossil

dinoflagellates 2019 edition. AASP Contribution Series, No. 50. American Association of

Stratigraphic Palynologists Foundation. ISSN 0160-8843.

FitzPatrick, M.E.J. 1995. Dinoflagellate cyst biostratigraphy of the Turonian (Upper Cretaceous) of

southern England. Cretaceous Research, 16, 757–791.

FitzPatrick, M.E.J. 1996. Recovery of Turonian dinoflagellate cyst assemblages from the effects of

the oceanic anoxic event at the end of the Cenomanian in southern England. In: Hart, M.B. (ed) Biotic

Recovery from Mass Extinction Events. Geological Society, London, Special Publications, 102, 279–

297.

Forster, A., Schouten, S., Moriya, K., Wilson, P.A. & Sinninghe Damsté, J.S. 2007 (online). Tropical

warming and intermittent cooling during the Cenomanian/Turonian oceanic anoxic event 2: sea

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 65: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

surface temperature records from the equatorial Atlantic. Paleoceanography, 22, PA1219,

Doi.org/10.1029/2006PA001349.

Foucher, J-C. 1979. Distribution stratigraphique des kystes de dinoflagellés et des acritarches dans le

Crétacé Supérieur du Bassin de Paris et de l‘Europe septentrionale. Palaeontographica Abteilung B,

169, 78–105.

Foucher, J-C. 1980. Dinoflagellés et acritarches dans le Crétacé du Bolonnais. Pp. 233, 288–297,

310–311. In: Robaszynski et al., Synthèse biostratigraphique de l‘Aptien au Santonien du Boulonnais

à partir de sept groups paléontologiques: Foraminifères, nannoplancton, dinoflagellés et macrofaunes

– Zonations micropaléontologiques integrées dans le cadre du Crétacé boreal nord-européen. Revue de

Micropaléontologie, 22(4), 195–321, 28 fig., 20pl.

Foucher, J.-C, 1981. Kystes de dinoflagellés dans le Crétacé Moyen Européen: Proposition d‘une

Echelle Biostratigraphique pour le Domaine Nord-occidental. Cretaceous Research, 2, 331–338.

Foucher, J.-C. 1982. Dinoflagellés et Acritarches du Turonien stratotypique (affleurments du

Saumurois, sondage de Civray-de-Touraine). In: Robaszynski, F., Alcaydé, G., Amedro, F., Badillet,

G., Damotte, R., Foucher, J.-C., Jardiné, S., Legoux, O., Manvit, H., Monciardini, C. & Sornay, J.

Turonien de la Region-type: Saumurois et Touraine. Stratigraphie, biozonations, sédimentologie.

Bulletin des Centres de Recherches Exploration-Production Elf-Aquitaine, 6, 147–150, 171–173, 176.

Foucher, J.-C. 1983. Distribution des kystes de dinoflagellés dans le Crétacé Moyen et Supérieur du

Bassin de Paris. Cahiers de Micropaleontologie, 4, 23–41.

Gale, A.S. 1995. Cyclostratigraphy and correlation of the Cenomanian Stage in Western Europe. In:

House, M.R. & Gale, A.S. (eds) Orbital Forcing Timescales and Cyclostratigraphy. Geological

Society, London, Special Publications, 85, 177–197.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 66: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Gale, A.S., Jenkyns, H.C., Kennedy, W.J. & Corfield, R.M. 1993. Chemostratigraphy versus

biostratigraphy: data from around the Cenomanian–Turonian boundary. Journal of the Geological

Society, London, 150, 29–32.

Gale, A. S. & Christensen, W.K. 1996. Occurrence of the belemnite Actinocamax plenus in the

Cenomanian of SE France and its significance. Bulletin of the Geological Society of Denmark, 43, 68–

77.

Gale, A.S., Smith, A.B., Monks, N.E.A., Young, J.A., Howard, A., Wray, D.S. & Huggett, J.M. 2000.

Marine biodiversity through the Late Cenomanian–Early Turonian: palaeoceanographic controls and

sequence stratigraphic biases. Journal of the Geological Society, London, 157, 745–757.

Gangl, S.K., Moy, C.M., Stirling, C.H., Jenkyns, H.C., Crampton, J.S., Clarkson, M.O., Ohneiser, C.

& Porcelli, D. 2019. High-resolution records of Oceanic Anoxic Event 2: Insights into the timing,

duration and extent of environmental perturbations from the palaeo-South Pacific Ocean. Earth and

Planetary Science Letters, 518, 172–182.

Gaunt, G.D., Fletcher, T.P. & Wood, C.J. 1992. Geology of the country around Kingston upon Hull

and Brigg. British Geological Survey Memoir, England and Wales, Sheets 80, 89. British Geological

Survey, Keyworth, Nottingham.

Gharaie, M.H.M. & Kalanat, B. 2018. Enhanced chemical weathering and organic carbon burial as

recovery factors for the OAE2 environmental conditions: a case study from Koppeh-Dagh Basin, NE

Iran. Geopersia, 8, 233–244.

Gradstein, F.M. & Waters, C.N. 2016 Stratigraphic Guide to the Cromer Knoll, Shetland and Chalk

Groups, North Sea and Norwegian Sea. Newsletters on Stratigraphy, 49, 71–280.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 67: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Grosheny, D., Ferry, S., Lécuyer, C., Thomas, A. & Desmares, D. 2017. The Cenomanian–Turonian

Boundary Event (CTBE) on the southern slope of the Subalpine Basin (SE France) and its bearing on

a probable tectonic pulse on a larger scale. Cretaceous Research, 72, 39–65.

Hancock, J.M. 1976. The Petrography of the Chalk. Proceedings of the Geologists’ Association, 86,

499–535, for 1975.

Hancock, J.M. 1989. Sea level changes in the British region during the Late Cretaceous. Proceedings

of the Geologists’ Association, 100, 565–594.

Hancock, J.M. 1991. Ammonite scales for the Cretaceous System. Cretaceous Research, 12, 259–

291.

Haq, B.U., Hardenbol, J. & Vail, P.R. 1988. Mesozoic and Cenozoic chronostratigraphy and cycles of

sea level change. In: Wilgus C.W. et al. (eds) Sea-Level Changes: An Integrated Approach. SEPM

Special Publication, 42, 71–108. ISBN 0-918985-74-9.

Haq, B.U. & Huber, B.T. 2016 (online). Anatomy of a eustatic event during the Turonian (Late

Cretaceous) hot greenhouse climate. Science China Earth Sciences. Doi: 10.1007/s11430-016-0166-y.

Harker, S.D., Gustav, S.H. & Riley, L.A. 1987. Triassic to Cenomanian stratigraphy of the Witch

Ground Graben. In: Brooks, S.J. & Glennie, K. (eds) Petroleum Geology of North-West Europe.

Graham & Trotman, London, 809–818.

Harker, S.D., Sarjeant, W.A.S. & Caldwell, W.G.E. 1990. Late Cretaceous (Campanian) organic-

walled microplankton from the Interior Plains of Canada, Wyoming and Texas: biostratigraphy,

palaeontology and environmental interpretation. Palaeontographica Abteilung B, 219, 1–243.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 68: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Harris, A.J. & Tocher, B.A. 2003. Palaeoenvironmental analysis of Late Cretaceous dinoflagellate

cyst assemblages using high-resolution sample correlation from the Western Interior Basin, USA.

Marine Micropaleontology, 48, 127–148.

Hart, M.B. 1996. Recovery of the food chain after the Late Cenomanian extinction event. In: Hart,

M.B. (ed) Biotic Recovery from Mass Extinction Events. Geological Society, London, Special

Publications, 102, 265–277.

Hart, M.B. 2005. Conversation with the Earth: A personal view. Report and Transactions of the

Devonshire Association for the Advancement of Science, Literature & the Arts, 137, 1–35.

Hart, M.B. 2019. The ‗Black Band‘: local expression of a global event. Proceedings of the Yorkshire

Geological Society, 62, 217–226.

Hart, M.B. & Bigg, P.J. 1981. Anoxic events in late Cretaceous chalk seas of North-West Europe. In:

Neale, J.W. & Brasier, M.D. (eds) Microfossils of Recent and Fossil Shelf Seas. Horwood, Chichester,

177–185.

Hart, M.B., Dodsworth, P., Ditchfield, P.W., Duane, A.M. & Orth, C.J. 1991. The late Cenomanian

event in eastern England. Historical Biology, 5, 339–354.

Hart, M.B., Dodsworth, P. & Duane, A.M. 1993. The late Cenomanian event in eastern England.

Cretaceous Research, 14, 495–508.

Hart, M.B. & Koutsoukos, E.A.M. 2015. Paleoecology of Cretaceous Foraminifera: Examples from

the Atlantic Ocean and Gulf of Mexico Region. Gulf Coast Association of Geological Societies

Transactions, 65, 175–199.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 69: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Hart, M.B. & Leary, P.N. 1989. The stratigraphic and palaeogeographic setting of the late

Cenomanian ‗anoxic‘ event. Journal of the Geological Society, London, 146, 305–310.

Hart, M.B., Weaver, P.P.E., Clements, R.G., Burnett, J.A., Tocher, B.A., Batten, D.J., Lister, J.K. &

MacLennan, A.M. 1987. The Isle of Wight. Cretaceous. In: Lord, A.R. & Bown, P.R. (eds) Mesozoic

and Cenozoic Stratigraphical Micropalaeontology of the Dorset Coast and Isle of Wight, Southern

England. British Micropalaeontological Society Guide Book 1, 88–149.

Hasegawa, T., Crampton, J.S., Schiøler, P., Field, B., Fukushi, C. & Kakizaki, Y. 2013. Carbon

isotope stratigraphy and depositional oxia through Cenomanian/Turonian boundary sequences (Upper

Cretaceous) in New Zealand. Cretaceous Research, 40, 61–80.

Hay, W.W., DeConto, R.M., de Boer, P.L., Flögel, S., Song, Y. & Stepashko, A. 2018 (online).

Possible solutions to several enigmas of Cretaceous climate. International Journal of Earth Sciences,

Doi.org/10.1007/s00531-018-1670-2

He, C., Zhichen, S. & Youhua, Z., 2009. Fossil dinoflagellates of China. Nanjing Institute of Geology

and Palaeontology, Chinese Academy of Sciences, Nanjing, 737 pp., 200 pl.

Heimhofer, U., Wucherpfennig, N., Adatte, T., Schouten, S., Schneebeli-Hermann, E., Gardin, S.,

Keller, K., Kentsch, S. & Kujau, A. 2018 (online). Vegetation response to exceptional global warmth

during Oceanic Anoxic Event 2. Nature Communications, 9, 3832, Doi: 10.1038/s41467-018-06319-

6.

Helmond, N.A.G.M. van, Sluijs, A., Reichart, G.-J., Sinninghe Damsté, J.S., Slomp, C.P. &

Brinkhuis, H. 2014. A perturbed hydrological cycle during Oceanic Anoxic Event 2. Geology, 42,

123–126.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 70: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Helmond, N.A.G.M. van, Sluijs, A., Reichart, G.-J., Sinninghe Damsté, J.S., Reichart, G.J., Voigt, S.,

Erbacher, J., Pross, J. & Brinkhuis, H. 2015. Freshwater discharge controlled deposition of

Cenomanian–Turonian black shales on the NW European epicontinental shelf (Wunstorf, northern

Germany). Climate of the Past, 11, 495–508.

Helmond, N.A.G.M. van, Sluijs, A., Papadomanolaki, A., Plint, G., Gröcke, D.R., Pearce, M.A.,

Eldrett, J.S., Trabucho-Alexandre, J., Walaszczyk, I., van de Schootbrugge, B. & Brinkhuis, H. 2016.

Equatorward phytoplankton migration during a cold spell within the Late Cretaceous super-

greenhouse. Biogeosciences, 13, 2859–2872.

Herbin, J.P., Montadert, L., Müller, C., Gomez, R., Thurow, J. & Wiedmann, J. 1986. Organic-rich

sedimentation at the Cenomanian–Turonian boundary in oceanic and coastal basins in the North

Atlantic and Tethys. In: Summerhayes, C.P. & Shackleton, N.J. (eds) North Atlantic

Palaeoceanography. Geological Society, London, Special Publications, 21, 389–422.

Herngreen, G.F.W. & Chlonova, A.F. 1981. Cretaceous microfloral provinces. Pollen et Spores, 23,

441–555.

Herngreen, G.F.W., Kedeves, M., Rovina, L.V. & Smirnova, S.B. 1996. Cretaceous palynofloral

provinces: a review. In: Jansonius, J. & McGregor, D.C. (eds) Palynology: Principles and

Applications, Volume 3. Salt Lake City, 1157–1188. American Association of Stratigraphic

Palynologists.

Hilbrecht, H.H. 1986. On the correlation of the Upper Cenomanian and Lower Turonian of England

and Germany (Boreal and N–Tethys). Newsletters on Stratigraphy, 15, 115–138.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 71: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Hilbrecht, H.H. & Dahmer, D.D. 1994. Sediment dynamics during the Cenomanian–Turonian

Oceanic Anoxic Event in Northwestern Germany. Facies, 30, 63–84.

Hilbrecht, H.H. & Hoefs, J. 1986. Geochemical and palaeontological studies of the δ13C anomaly in

boreal and North Tethyan Cenomanian–Turonian sediments in Germany and adjacent areas.

Palaeogeography, Palaeoclimatology, Palaeoecology, 53, 169–189.

Hilbrecht, H.H., Hubberten, H.-W. & Oberhänsli, H. 1992. Biostratigraphy of planktonic foraminifera

and regional isotope variations: productivity and water masses in Late Cretaceous Europe.

Palaeogeography, Palaeoclimatology, Palaeoecology, 92, 407–421.

Hildreth, P. 1999. The Variegated Beds Member of the Welton Chalk Formation of North

Lincolnshire. Humberside Geologist, 12, 13–30.

Hopson, P.M. 2005. A stratigraphical framework for the Upper Cretaceous Chalk of England and

Scotland with statements on the Chalk of Northern Ireland and the UK Offshore Sector. British

Geological Survey, Research Report RR/05/01. 102 pp. ISBN 0 85272 517 5.

Hu, X.-F., Jeans, C.V. & Dickson, T. 2012. Geochemical and stable isotope patterns of calcite

cementation in the Upper Cretaceous Chalk, UK: Direct evidence from calcite-filled vugs in

brachiopods. Acta Geologica Polonica, 62, 143–172.

Huber, B.T., Norris, R.D. & McLeod, K.D. 2002. Deep-sea paleotemperature record of extreme

warmth during the Cretaceous. Geology, 30, 123–126.

Ilyina, V.I., Kulkova, I.A. & Lebedeva, N.K. 1994. Microphytofossils and detailed stratigraphy of

marine Mesozoic and Cenozoic of Siberia. (Mikrofitofossilii i detalnaya stratigrafiya morskogo i

kainozoya Sibiri.) Russian Academy of Sciences, Siberian Branch, United Institute of Geology,

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 72: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Geophysics and Mineralogy, Transactions, Issue no. 818, 1–192, pl.1–56. (In Russian with English

abstract.)

International Commission on Stratigraphy 2019 (online). International Chronostratigraphic Chart, v.

2019/05. IUGS. www.stratigraphy.org. (Updated from: Cohen, K.M., Finney, S.C., Gibbard, P.L. &

Fan, J.-X. 2013. The ICS International Chronostratigraphic Chart. Episodes, 36, 199–204.)

Ioannides, N.S. 1986. Dinoflagellate cysts from Upper Cretaceous–Lower Tertiary sections, Bylot and

Devon Islands, Arctic Archipelago. Geological Survey of Canada, Bulletin, 371, 1–99, pl.1–25.

Ioannides, N.S., Stavrinos, G.N. & Downie, C. 1976. Kimmeridgian microplankton from Clavell‘s

Head, Dorset, England. Micropaleontology, 22, 443–478.

Ion, J., Antonescu, E., Melinte, M.C. & Szasz, L. 2004. Integrated biostratigraphy of the Turonian of

Romania. Acta Palaeontologica Romaniae, 4, 151–161.

Isaksen, D. & Tonstad, K. (eds) 1989. A revised Cretaceous and Tertiary lithostratigraphic

nomenclature for the Norwegian North Sea. Norwegian Petroleum Directorate Bulletin, 5, 59 pp.

Jarvis, I., Carson, G.A., Cooper, K., Hart, M.B., Horne, D., Leary, P.N., Rosenfeld, A. & Tocher,

B.A. 1988a. Chalk microfossil assemblages and the Cenomanian–Turonian (late Cretaceous) oceanic

anoxic event, new data from Dover, England. Cretaceous Research, 9, 3–103.

Jarvis, I., Carson, G.A., Hart, M.B., Leary, P.N. & Tocher, B.A. 1988b. The Cenomanian–Turonian

(late Cretaceous) oceanic anoxic event in SW England: evidence from Hooken Cliffs near Beer, SE

Devon. Newsletters on Stratigraphy, 18, 147–164.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 73: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Jarvis, I., Gale, A.S., Jenkyns, H.C. & Pearce, M.A. 2006. Secular variation in Late Cretaceous carbon

isotopes: a new d13C carbonate reference curve for the Cenomanian–Campanian (99.6–70.6 Ma).

Geological Magazine, 143, 561–608.

Jarvis, I., Lignum, J.S., Gröcke, D.R., Jenkyns, H.C. & Pearce, M.A. 2011 (online). Black shale

deposition, atmospheric CO2 drawdown, and cooling during the Cenomanian–Turonian Oceanic

Anoxic Event. Paleoceanography, 26 (PA3201), 17pp.

Jarvis, I., Trabucho-Alexandre, J., Gröcke, D., Uličnỳ, D. & Laurin, J. 2015. Intercontinental

correlation of organic carbon and carbonate stable isotope records: evidence of climate and sea-level

change during the Turonian (Cretaceous). The Depositional Record, 1, 53–90.

Jeans, C.V. 1980. Early submarine lithification in the Red Chalk and Lower Chalk of eastern England:

a bacterial control model and its implications. Proceedings of the Yorkshire Geological Society, 43,

81–157.

Jeans, C.V., Long, D., Hall, M.A., Bland, D.J. & Cornford, C. 1991. The geochemistry of the Plenus

Marls at Dover, England: evidence of fluctuating oceanographic conditions and of glacial control

during the development of the Cenomanian–Turonian δ13C anomaly. Geological Magazine, 128, 604–

632.

Jeans, C.V., Long, D., Hu, X.-F. & Mortimore, R.N. 2014. Regional hardening of Upper Cretaceous

Chalk in eastern England: trace element and stable isotope patterns in the Upper Cenomanian and

Turonian Chalk and their significance. Acta Geologica Polonica, 64, 419–455.

Jefferies, R.P.S. 1962. The palaeoecology of the Actinocamax plenus Subzone (Lowest Turonian) in

the Anglo-Paris Basin. Palaeontology, 4, 609–647.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 74: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Jefferies, R.P.S. 1963. The stratigraphy of the Actinocamax plenus Subzone (Turonian) in the Anglo-

Paris Basin. Proceedings of the Geologists’ Association, 74, 1–30.

Jenkyns, H.C. 1980. Cretaceous anoxic events: from continents to oceans. Journal of the Geological

Society, London, 137, 171–188.

Jenkyns, H.C. 1985. The Early Toarcian and Cenomanian–Turonian anoxic events in Europe:

comparisons and contrasts. Geologische Rundschau, 74, 505–518.

Jenkyns, H.C. 2010 (online). Geochemistry of oceanic anoxic events. Geochemistry, Geophysics,

Geosystems, 11(3), Q03004, Doi:10.1029/2009GC002788.

Jenkyns, H.C., Dickson, A.J., Ruhl, M. & Boorn, S.H.J.M. van den. 2017. Basalt–seawater

interaction, the Plenus Cold Event, enhanced weathering and geochemical change: deconstructing

Oceanic Anoxic Event 2 (Cenomanian–Turonian, Late Cretaceous). Sedimentology, 64, 16–43.

Johnson, H. & Lott, G.K. 1993. 2. Cretaceous of the Central and Northern North Sea. In: Knox,

R.W.O‘B. & Cordey, W.G. (eds) Lithostratigraphic nomenclature of the UK North Sea. British

Geological Survey, Nottingham.

Joo, Y.J. & Sageman, B.B. 2014. Cenomanian to Campanian carbon isotope chemostratigraphy from

the Western Interior Basin, U.S.A. Journal of Sedimentary Research, 84, 529–542.

Kauffman, E.G. 1984. The fabric of marine extinctions. In: Berggren, W.A. & Van Couvering, J. (eds)

Catastrophes and Earth History: the New Uniformitarianism. Princeton University Press, Princeton,

N.J., 151–246.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 75: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Kauffman, E.G. & Caldwell, W.G.E. 1993. The Western Interior Basin in Space and Time. In:

Kauffman, E.G. & Caldwell, W.G.E. (eds) Evolution of the Western Interior Basin. Geological

Association of Canada Special Paper, 39, St Johns, NL, Canada, 1–30.

Keller, G. & Pardo, A. 2004. Age and paleoenvironment of the Cenomanian–Turonian global

stratotype section and point at Pueblo, Colorado. Marine Micropaleontology, 51, 95–128.

Kennedy, W.J. & Cobban, W.A. 1991. Stratigraphy and inter-regional correlation of the Cenomanian–

Turonian transition in the Western Interior of the United States near Pueblo, Colorado: A potential

boundary stratotype for the base of the Turonian stage. Newsletters in Stratigraphy, 24, 1–33.

Kennedy, W.J. & Gale, A.S. 2006. The Cenomanian Stage. Proceedings of the Geologists’

Association, 117, 187–205.

Kennedy, W.J., Gale, A.S., Lees, J.A. & Caron, M. 2004. The Global Boundary Stratotype Section

and Point (GSSP) for the base of the Cenomanian Stage, Mont Risou, Hautes-Alpes, France.

Episodes, 27, 21–32.

Kennedy, W.J., Walaszczyk, I. & Cobban, W.A. 2000. Pueblo, Colorado, USA, Candidate Global

Boundary Stratotype Section and Point for the base of the Turonian Stage of the Cretaceous and for

the Middle Turonian Substage, with a revision of the Inoceramidae (Bivalvia). Acta Geologica

Polonica, 50, 295–334.

Kennedy, W.J., Walaszczyk, I., Cobban, W.A., Dodsworth, P., Elder, W.P., Gale, A.S., Scott, G.R.,

Hancock, J.M., Voigt, S. & Kirkland, J.I. 2005. The Global Boundary Stratotype Section and Point for

the base of the Turonian Stage of the Cretaceous: Pueblo, Colorado, U.S.A. Episodes, 28, 93–104.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 76: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Kjellström, G. 1973. Maastrichtian microplankton from the Höllviken Borehole No.1 in Scania,

southern Sweden. Sveriges Geologiska Undersökning, Serie C, no.688, 67(8), 1–59.

Košták, M., Čhek, S., Ekrt, B., Mazuch, M., Wiese, F., Voigt, S. & Wood, C.J. 2004. Belemnites of

the Bohemian Cretaceous Basin in a global context. Acta Geologica Polonica, 54, 511–533.

Lamolda, M.A., Gorostidi, A. & Paul, C.R.C. 1994. Quantitative estimates of calcareous nannofossil

changes across the Plenus Marls (latest Cenomanian), Dover, England: implications for the generation

of the Cenomanian–Turonian Boundary Event. Cretaceous Research, 15, 143–164.

Lamolda, M.A. & Mao, S. 1999. The Cenomanian–Turonian boundary event and dinocyst record at

Ganuza (northern Spain). Palaeogeography, Palaeoclimatology, Palaeoecology, 150, 65–82.

Li, H. & Habib, D. 1996. Dinoflagellate stratigraphy and its response to sea level change in

Cenomanian–Turonian sections of the Western Interior of the United States. Palaios, 11, 15–30.

Linnert, C., Mutterlose, J. & Erbacher, J. 2010. Calcareous nannofossils of the Cenomanian/Turonian

boundary interval from the Boreal Realm (Wunstorf, northern Germany). Marine Micropaleontology,

74, 38–58.

Lucas-Clark, J. 1984. Morphology of species of Litosphaeridium (Cretaceous, Dinophyceae).

Palynology, 8, 165–193.

Mao, S. & Lamolda, M.A. 1999. Quistes de dinoflagelados del Cenomaniense superior y Turoniense

inferior de Ganuza, Navarra, II. – Biostatigafía. Revista Española de Paleontología, no. extr.

Homenaje al Prof. J. Truylos, 195–203.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 77: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Marcinowski, R., Walaszczyk, I. & Olszewska-Neijbert, D., 1996. Stratigraphy and regional

development of the mid-Cretaceous (Upper Albian through Coniacian) of the Mangyshlak Mountains,

western Kazakhstan. Acta Geologica Polonica, 46, 1–60.

Marshall, J.D. 1992. Climatic and oceanographic isotopic signals from the carbonate rock record and

their preservation. Geological Magazine, 129, 143–160.

Marshall, K.L. 1983. Dinoflagellate cysts from the Cenomanian, Turonian and Coniacian of Germany

and England. PhD thesis (unpublished), University of Aberdeen.

Marshall, K.L. & Batten, D.J. 1988. Dinoflagellate cyst associations in Cenomanian–Turonian ‗Black

Shale‘ sequences of northern Europe. Review of Palaeobotany and Palynology, 54, 85–103.

McCrea, J. M. 1950. On the isotopic chemistry of carbonates and a paleotemperature scale. The

Journal of Chemical Physics, 18, 849–857.

McLachlan, S.M.S., Pospelova, V. & Hebda, R.J. 2018. Dinoflagellate cysts from the upper

Campanian (Upper Cretaceous) of Hornby Island, British Columbia, Canada, with implications for

Nanaimo Group biostratigraphy and paleoenvironmental reconstructions. Marine

Micropalaeontology, 145, 1–20.

McMinn, A. 1988. Outline of a Late Cretaceous dinoflagellate zonation of northwest Australia.

Alcheringa, 12, 137–156.

Mertens, K.N., Verhoeven, K., Verleye, T., Louwye, S., Amorim, A., Ribeiro, S., Deaf, A.S.,

Harding, I.C., Schepper, S. de, Gonzalez, C., Kodrans-Nsiah, M., Vernal, A. de, Henry, M., Radi, T.,

Dybkjaer, K., Poulsen, N.E., Feist-Burkhardt, S., Chitolie, J., Heilmann-Clausen, C., Londeix, L.,

Turon, J.-L., Marret, F., Matthiessen, J., McCarthy, F.M.G., Prasad, V., Pospelova, V., Kyffin

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 78: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Hughes, J.E., Riding, J.B., Rochon, A., Sangiorgi, F., Welters, N., Sinclair, N., Thun, C., Soliman, A.,

Nieuwenhove, N. van, Vink, A. & Young, M. 2009. Determining the absolute abundance of

dinoflagellate cysts in recent marine sediments: The Lycopodium marker-grain method put to the test.

Review of Palaeobotany and Palynology, 157, 238–252.

Miller, K.G., Sugarman, P.J., Browning, J.V., Kominz, M.A., Hernández, J.C., Olsson, R.K., Wright,

J.D., Feigenson, M.D. & Sickel, W. van. 2003. Late Cretaceous chronology of large, rapid sea-level

changes: Glacioeustasy during the greenhouse world. Geology, 31, 585–588.

Miller, K.G., Wright, J.D. & Browning, J.V. 2005. Visions of ice sheets in a greenhouse world.

Marine Geology, 217, 215–231.

Milne, D., Raup, D., Billingham, J., Niklaus, K. & Padian, K. (eds) 1985. The evolution of complex

and higher organisms. NASA SP-478. US Government Printing Office, Washington, DC.

Minisini, D., Eldrett, J.S., Bergman, S.C. & Forkner, R. 2018. Chronostratigraphic framework and

depositional environments in the organic-rich, mudstone-dominated Eagle Ford Group, Texas, USA.

Sedimentology, 65, 1520–1557.

Mitchell, S.F. 2000. The Welton Formation (Chalk Group) at Speeton, NE England: implications for

the late Cretaceous evolution of the Market Weighton Structure. Proceedings of the Yorkshire

Geological Society, 53, 17–23.

Mitchell, S.F. 2019. The Chalk Group (Upper Cretaceous) of the Northern Province, eastern England

– a review. Proceedings of the Yorkshire Geological Society, 62, 153–177.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 79: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Mitchell, S.F., Paul, C.R.C. & Gale, A.S. 1996. Carbon isotopes and sequence stratigraphy. In:

Howell, J.A. & Aitken, J.F. (eds) High Resolution Sequence Stratigraphy: Innovations and

Applications. Geological Society, London, Special Publications, 104, 11–24.

Molen, A.S. van der & Wong, T. E. 2007. Towards an improved lithostratigraphic subdivision of the

Chalk Group in the Netherlands North Sea area—A seismic stratigraphic approach. Netherlands

Journal of Geosciences (Geologie en Mijnbouw), 86, 131–143.

Monteiro, F., Pancost, R., Ridgwell, A. & Donnadieu, Y. 2012 (online). Nutrients as the dominant

control on the spread of anoxia and euxinia across the Cenomanian–Turonian Oceanic Anoxic Event

(OAE 2): Model‐data comparison. Paleoceanography, 27, PA4209, Doi: 10.1029/2012PA002351.

Moore, C.H. 1989. Diagenetic environments of porosity modification and tools for their recognition in

the geologic record. Chapter 3. In: Moore, C.H. (ed.) Carbonate Diagenesis and Porosity.

Developments in Sedimentology, 46. 338 pp. eBook ISBN 9780080869605.

Morgan, R. 1980. Palynostratigraphy of the Australian Early and Middle Cretaceous. Geological

Survey of New South Wales, Palaeontology Memoir, 18, 153 pp.

Mortimore, R.N. 2014. Logging the Chalk. Whittles Publishing, Dunbeath, Scotland.

Mudie, P.J. & McCarthy, F.M.G. 1994. Late Quaternary pollen transport processes, western north

Atlantic: Data from box models, cross-margin and N-S transects. Marine Geology, 118, 79–105.

Muller, J. 1959. Palynology of Recent Orinoco delta and shelf sediments: Reports of the Orinoco

Shelf Expedition; Volume 5. Micropaleontology, 5, 1–32.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 80: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Naidin, D.P. 1993. Late Cretaceous events in the East-European Paleobiogeographic Province. 2.

Cenomanian/Turonian and Maastrichtian/Danian events. Bulletin of Moscow Society of Naturalists.

Geological Series, 68, 33–53. (In Russian, English abstract.)

Nøhr-Hansen, H. 2012. Palynostratigraphy of the Cretaceous–lower Palaeogene sedimentary

succession in the Kangerlussuaq Basin, southern East Greenland. Review of Palaeobotany and

Palynology, 178, 59–90.

Nøhr-Hansen, H., Costa, L., Pearce, M.A. & Alsen, P. 2018. New Albian to Cenomanian (Cretaceous)

dinoflagellate cyst taxa of ovoidinioid affinities from East Greenland, the Barents Sea and England.

Palynology, 42, 366–391.

Nøhr-Hansen, H., Williams, G.L. & Fensome, R.A. 2016. Biostratigraphic correlation of the western

and eastern margins of the Labrador–Baffin Seaway and implications for the regional geology.

Geological Survey of Denmark and Greenland Bulletin, 37, 74 pp.

O‘Brien, C.L., Robinson, S.A., Pancost, R.D., Sinninghe Damsté, J.S., Schouten, S., Lunt, D. J.

Alsenz, H., Bornemann, A., Bottini, C., Brassell, S.C., Farnsworth, A., Forster, A., Huber, B.T., Inglis, G.N.,

Jenkyns, H.C., Linnert, C., Littler, K., Markwick, P., McAnena, A., Mutterlose, J., Naafs, B.D.A., Püttmann,

W., Sluijs, A., van Helmond, N.A.G.M., Vellekoop, J., Wagner, T. & Wrobel, N.E. 2017. Cretaceous sea-

surface temperature evolution: Constraints from TEX86 and planktonic foraminiferal oxygen isotopes.

Earth-Science Reviews, 172, 224–247.

O‘Connor, L.K., Jenkyns, H.C., Robinson, S.A., Remmelzwaal, S.R.C., Batenburg, S.J, Parkinson,

I.J. & Gale, A.S. 2020 (online). A Re‐evaluation of the Plenus Cold Event, and the Links Between

CO2, Temperature, and Seawater Chemistry During OAE 2. Paleoceanography and

Paleoclimatology, 35. Doi.org/10.1029/2019PA003631.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 81: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Olde, K., Jarvis, I., Pearce, M., Uličnỳ, D., Tocher, B., Trabucho-Alexandre, J. & Gröcke, D. 2015a.

A revised northern European Turonian (Upper Cretaceous) dinoflagellate biostratigraphy: integrating

palynology and carbon isotope events. Review of Palaeobotany and Palynology, 213, 1–15.

Olde, K., Jarvis, I., Uličnỳ, D., Pearce, M., Trabucho-Alexandre, J., Čech, S., Gröcke, D., Laurin, J.,

Švábenická, L. & Tocher, B. 2015b. Geochemical and palynological sea-level proxies in hemipelagic

sediments: A critical assessment from the Upper Cretaceous of the Czech Republic.

Palaeogeography, Palaeoclimatology, Palaeoecology, 435, 222–243.

Paul, C.R.C., Lamolda, M.A., Mitchell, S.F., Vaziri, M.R., Gorostidi, A. & Marshall, J.D. 1999. The

Cenomanian–Turonian boundary at Eastbourne (Sussex, UK): a proposed European reference section.

Palaeogeography, Palaeoclimatology, Palaeoecology, 150, 83–121.

Paul, C.R.C. & Mitchell, S.F. 1994. Is famine a common factor in marine mass extinctions? Geology,

22, 679–682.

Paul, C.R.C., Mitchell, S.F., Marshall, J.D., Leary, P.N., Gale, A.S., Duane, A.M. & Ditchfield, P.W.

1994. Palaeoceanographic events in the Middle Cenomanian of northwest Europe. Cretaceous

Research, 15, 707–738.

Pavlishina, P. 1990. Early Cenomanian palynomorphs near the village of Sanadinovo, central north

Bulgaria. Review of the Bulgarian Geological Society, 51, 89–101; pl.1–3.

Pavlishina, P. & Minev, V. 1998. Turonian and Coniacian Normapolles from southwest to northeast

Bulgaria and their calibration against the standard ammonite zones. Zentralblatt für Geologie und

Paläontologie, Teil I, 1217–1223.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 82: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Pearce, M.A. 2010. New organic-walled dinoflagellate cysts from the Cenomanian to Maastrichtian of

the Trunch borehole, UK. Journal of Micropalaeontology, 29, 51–72.

Pearce, M.A. 2018. Additional new organic-walled dinoflagellate cysts from two onshore UK Chalk

boreholes. Journal of Micropalaeontology, 37, 73–86.

Pearce, M.A., Jarvis, I., Ball, P. & Laurin, J. 2020 (in press). Palynology of the Cenomanian to

lowermost Campanian (Upper Cretaceous) Chalk of the Trunch Borehole (Norfolk, UK) and a new

dinoflagellate cyst bioevent stratigraphy for NW Europe. Review of Palaeobotany and Palynology,

278. Doi: 10.1016/j.revpalbo.2020.104188.

Pearce, M.A., Jarvis, I., Swan, A.R.H., Murphy, A.M., Tocher, B.A. & Edmunds, W.M. 2003.

Integrating palynological and geochemical data in a new approach to palaeoecological studies: Upper

Cretaceous of the Banterwick Barn Chalk borehole, Berkshire, UK. Marine Micropaleontology, 47,

271–306.

Pearce, M.A., Jarvis, I. & Tocher, B.A. 2009. The Cenomanian–Turonian boundary event, OAE2 and

palaeoenvironmental change in epicontinental seas: new insights from the dinocyst and geochemical

records. Palaeogeography, Palaeoclimatology, Palaeoecology, 280, 207–234.

Pearce, M.A., Lignum, J.S. & Jarvis, I. 2011. Senoniasphaera turonica (Prössl, 1990 ex Prössl, 1992)

comb. nov., senior synonym of Senoniasphaera rotundata alveolata Pearce et al., 2003: an important

dinocyst marker for the Lower Turonian chalk of NW Europe. Journal of Micropalaeontology, 30,

91–93.

Peryt, D. & Wyrwicka, K. 1993. The Cenomanian/Turonian boundary event in Central Poland.

Palaeogeography, Palaeoclimatology, Palaeoecology, 104, 185–197.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 83: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Peyrot, D. 2011. Late Cretaceous (Late Cenomanian–Early Turonian) dinoflagellate cysts from the

Castilian Platform, northern Spain. Palynology, 35, 267–300.

Peyrot, D., Barrón, E., Comas-Rengifo, M.J., Barroso-Barcenilla, F. & Feist-Burkhardt, S. 2008.

Palynological characterisation of the Cenomanian/Turonian boundary at the Puentedey section

(Burgos, Spain). Coloquios de Paleontología, 58, 101–161.

Peyrot, D., Barroso-Barcenilla, F., Barrón, E. & Comas-Rengifo, M.J. 2011. Palaeoenvironmental

analysis of Cenomanian–Turonian dinocyst assemblages from the Castilian Platform (Northern-

Central Spain). Cretaceous Research, 32, 504–526.

Peyrot, D., Barroso-Barcenilla, F. & Feist-Burkhardt, S. 2012. Palaeoenvironmental controls on late

Cenomanian–early Turonian dinoflagellate cyst assemblages from Condemios (Central Spain).

Review of Palaeobotany and Palynology, 180, 25–40.

Pomerol, B. & Mortimore, R.N. 1993. Lithostratigraphy and correlation of the Cenomanian–Turonian

boundary sequence. Newsletters on Stratigraphy, 28, 59–78.

Prauss, M.L. 1993. Sequence-palynology – Evidence from Mesozoic sections and conceptual

framework. Neues Jahrbuch für Geologie und Paläontologie – Abhandlungen, 190, 143–163.

Prauss, M.L. 2006. The Cenomanian/Turonian Boundary Event (CTBE) at Wunstorf, north-west

Germany, as reflected by marine palynology. Cretaceous Research, 27, 872–886.

Prauss, M.L. 2007. Availability of reduced nitrogen chemospecies in photic-zone waters as the

ultimate cause of fossil prasinophyte prosperity. Palaios, 22, 489–499.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 84: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Prauss, M.L. 2012(a). The Cenomanian/Turonian Boundary event (CTBE) at Tarfaya, Morocco:

Palaeoecological aspects as reflected by marine palynology. Cretaceous Research, 34, 233–256.

Prauss, M.L. 2012(b). The Cenomanian/Turonian Boundary Event (CTBE) at Tarfaya, Morocco,

northwest Africa: Eccentricity controlled water column stratification as major factor for total organic

carbon (TOC) accumulation: Evidence from marine palynology. Cretaceous Research, 37, 246–260.

Prauss, M.L. 2012(c). Potential freshwater dinocysts from marine upper Cenomanian to upper

Coniacian strata of Tarfaya, northwest Africa: Three new species of Bosedinia. Cretaceous Research,

37, 285–290.

Prauss, M.L. 2015. Marine palynology of the Oceanic Anoxic Event 3 (OAE3, Coniacian–Santonian)

at Tarfaya, Morocco, NW Africa – transition from preservation to production controlled accumulation

of marine organic carbon. Cretaceous Research, 53, 19–37.

Prince, I.M., Jarvis, I., Pearce, M.A. & Tocher, B.A. 2008. Dinoflagellate cyst biostratigraphy of the

Coniacian–Santonian (Upper Cretaceous): New data from the English Chalk. Review of Palaeobotany

and Palynology, 150, 59–96.

Prince, I.M., Jarvis, I. & Tocher, B.A. 1999. High resolution dinoflagellate cyst biostratigraphy of the

Santonian – basal Campanian (Upper Cretaceous): new data from Whitecliff, Isle of Wight, England.

Review of Palaeobotany and Palynology, 105, 143–169.

Prössl, K.F. 1990. Dinoflagellaten der Kreide - Unter-Hauterive bis Ober-Turon - im

niedersächsischen Becken. Stratigraphie und Fazies in der Kernbohrung Konrad 101 sowie einiger

anderer Bohrungen in Nordwestdeutschland. Palaeontographica, Abteilung B, 218, 93–191.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 85: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Raup, D.M. & Sepkoski, J.J. Jr. 1982. Mass extinctions in the marine fossil record. Science, 215,

1501–1503.

Raup, D.M. & Sepkoski, J.J. Jr. 1984. Periodicity of extinctions in the geological past. Proceedings of

the National Academy of Sciences, 81, 801–805.

Rhys, G.H. 1974. A proposed standard lithostratigraphic nomenclature for the southern North Sea

and an outline structural nomenclature for the whole of the (UK) North Sea. Institute of Geological

Sciences, Report No. 74/8. H.M.S.O., London.

http://pubs.bgs.ac.uk/publications.html?pubID=B00992

Robinson, S.A., Dickson, A.J., Pain, A., Jenkyns, H.C., O‘Brien, C.L., Farnsworth, A. & Lunt, D.J.

2019. Southern Hemisphere sea-surface temperatures during the Cenomanian–Turonian: Implications

for the termination of Oceanic Anoxic Event 2. Geology, 47(2), 131–134.

Sahagian, D., Pinous, O., Olferiev, A. & Zakharov, V. 1996. Eustatic Curve for the Middle Jurassic–

Cretaceous Based on Russian Platform and Siberian Stratigraphy: Zonal Resolution. AAPG Bulletin

80, 1433–1458.

Schiøler, P. & Crampton, J.S. 2014. Dinoflagellate biostratigraphy of the Arowhanan Stage (upper

Cenomanian–lower Turonian) in the East Coast Basin, New Zealand. Cretaceous Research, 48, 205–

224.

Schiøler, P. & Wilson, G.J. 1993. Maastrichtian dinoflagellate zonation in the Dan Field, Danish

North Sea. Review of Palaeobotany and Palynology, 78, 321–351.

Schlanger, S.O., Arthur, M.A., Jenkyns, H.C. & Scholle, P.A. 1987. The Cenomanian–Turonian

Oceanic Anoxic Event, I: Stratigraphy and distribution of organic carbon-rich beds and the marine

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 86: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

δ13C excursion. In: Brooks, J. & Fleet, A.J. (eds) Marine Petroleum Source Rocks. Geological

Society, London, Special Publications, 26, 371–399.

Schlanger, S.O. & Jenkyns, H.C. 1976. Cretaceous Oceanic Anoxic Events: Causes and

Consequences. Geologie en Mijnbouw, 55, 179–184.

Singh, C. 1971. Lower Cretaceous microfloras of the Peace River area, northwestern Alberta.

Research Council of Alberta, Bulletin, 28, 301–542, pl. 39–80.

Sinninghe Damsté, J.S. & Köster, J. 1998. A euxinic southern North Atlantic Ocean during the

Cenomanian/Turonian oceanic anoxic event. Earth and Planetary Science Letters, 158, 165–173.

Sinninghe Damsté, J.S., Bentum, E.C. van, Reichart, G.J., Pross, J. & Schouten, S. 2010. A CO2

decrease-driven cooling and increased latitudinal temperature gradient during the mid-Cretaceous

Oceanic Anoxic Event 2. Earth and Planetary Science Letters, 293, 97–103.

Slimani, H. 1994. Les dinokystes des craies du Campanien au Danien à Halembaye, Turnhout

(Belgique) et à Beutenaken (Pays-Bas). Mémoires pour servir à l'explication des cartes géologiques et

minières de la Belgique, 37, 173 pp., 18 pl.

Stockmarr, J., 1971. Tablets with spores used in absolute pollen analysis. Pollen et Spores, 13, 615–

621.

Stover, L.E. & Evitt, W.R. 1978. Analyses of Pre-Pleistocene Organic-walled Dinoflagellates.

Stanford University Publications, Geological Series, Volume XV. Stanford, California.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 87: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Stover, L.E. & Helby, R. 1987. Some Early Cretaceous dinoflagellates from the Houtman-1 well,

Western Australia. In: Jell, P.A. (ed) Studies in Australian Mesozoic palynology. Memoir of the

Association of Australasian Palaeontologists, 4, 261–295.

Surlyk, F., Dons, T., Clausen, C.K. & Higham, J. 2003. Upper Cretaceous. In: Evans, D., Graham, C.,

Armour, A. & Bathurst, P. (eds) The Millennium Atlas: Petroleum Geology of the Central and

Northern North Sea. The Geological Society, London, 213–233. ISBN 1 86239 119 X.

Tocher, B.A. & Jarvis, I. 1987. Dinoflagellate cysts and stratigraphy of the Turonian (Upper

Cretaceous) Chalk, near Beer, south-east Devon, England. In: Hart, M.B. (ed) Micropalaeontology of

Carbonate Environments. Special Publication, British Micropalaeontological Society, Ellis Horwood,

Chichester, 138–175.

Traverse, A. 2007. Paleopalynology (2nd Edition). Springer. ISBN 978-1-4020-6684-9.

Tröger, K.-A., Kennedy, W.J., Burnett, J.A., Gale, A.S. & Robaszynski, F. 1996. The Cenomanian

Stage. In: Rawson et al. (eds) Proceedings, "Second International Symposium on Cretaceous Stage

Boundaries", Brussels 8–16 September 1995. Bulletin de l'Institut Royal des Sciences Naturelles de

Belgique – Sciences de la Terre, 66, 57–68.

Tsikos, H., Jenkyns, H.C., Walsworth-Bell, B., Petrizzo, M.R., Forster, A., Kolonic, S., Erba, E.,

Premoli Silva, I., Baas, M., Wagner, T. & Sinninghe Damsté, J.S. 2004. Carbon-isotope stratigraphy

recorded by the Cenomanian–Turonian Oceanic Anoxic Event: Correlation and implications based on

three localities. Journal of the Geological Society, London, 161, 711–719.

Uramoto, G.-I., Tahara, R., Sekiya, T. & Hirano, H. 2013. Carbon isotope stratigraphy of terrestrial

organic matter for the Turonian (Upper Cretaceous) in northern Japan: Implications for ocean-

atmosphere δ13C trends during the mid-Cretaceous climatic optimum. Geosphere, 9, 355–366.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 88: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Veizer, J. & Prokoph, A. 2015. Temperatures and oxygen isotopic composition of Phanerozoic

oceans. Earth-Science Reviews, 146, 92–104.

Voigt, S., Gale, A.S. & Flögel, S. 2004 (online). Midlatitude shelf seas in the Cenomanian–Turonian

greenhouse world: Temperature evolution and North Atlantic circulation. Paleoceanography, 19,

PA4020, Doi:10.1029/2004PA001015.

Voigt, S., Gale, A.S. & Voigt, T. 2006. Sea-level change, carbon cycling and palaeoclimate during the

Late Cenomanian of northwest Europe; an integrated palaeoenvironmental analysis. Cretaceous

Research, 27, 836–858.

Wall, D. 1965. Microplankton, pollen and spores from the Lower Jurassic in Britain.

Micropalaeontology, 11, 151–190.

Weiss, H.M., Wilhelms, A., Mills, N., Scotchmer, J., Hall, P.B., Lind, K. & Brekke, T. 2000.

NIGOGA - The Norwegian Industry Guide to Organic Geochemical Analyses [online]. Edition 4.0.

Published by Norsk Hydro, Statoil, Geolab Nor, SINTEF Petroleum Research and the Norwegian

Petroleum Directorate. 102 pp. http://www.npd.no/engelsk/nigoga/default.htm.

Wiese, F., Košták, M. & Wood, C.J. 2009. The Upper Cretaceous belemnite Praeactinocamax plenus

(Blainville, 1827) from Lower Saxony (Upper Cenomanian, northwest Germany) and its distribution

pattern in Europe. Paläontologische Zeitschrift, 83, 309–321.

Whitham, F. 1991. The stratigraphy of the Upper Cretaceous Ferriby, Welton and Burnham

formations north of the Humber, north-east England. Proceedings of the Yorkshire Geological

Society, 48, 227–254.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 89: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Williams, G.L. 1977. Dinocysts: their palaeontology, biostratigraphy and palaeoecology. In: Ramsey,

A.T.S. (ed.) Oceanic Micropalaeontology. Academic Press, London, 1231–1325.

Williams, G.L., Brinkhuis, H., Pearce, M.A., Fensome, R.A. & Weegink, J.W. 2004. Southern Ocean

and global dinoflagellate cyst events compared: index events for the Late Cretaceous–Neogene. In:

Exon, N.F., Kennett, J.P. & Malone M.J. (eds) Proceedings of the Ocean Drilling Program, Scientific

Results, 189, 1–98.

Williams, G.L. & Bujak, J.P. 1985. Mesozoic and Cenozoic dinoflagellates. In: Bolli, H.M., Saunders,

J.B. & Perch-Nielsen, K. (eds) Plankton Stratigraphy. Cambridge Earth Sciences. Cambridge

University Press, 847–964.

Wood, C.J., Batten, D.J., Mortimore, R.N. & Wray, D.S. 1997. The stratigraphy and correlation of the

Cenomanian–Turonian boundary interval succession in Lincolnshire, eastern England. Freiberger

Forschungsheft, C468, 333–346.

Wood, C.J. & Mortimore, R.N. 1995. An anomalous Black Band succession (Cenomanian–Turonian

boundary interval at Melton Ross, Lincolnshire, eastern England and its international significance.

Berliner Geowissenschaft Abhandlungen, E16 (Gundolf Ernst Festschrift), 277–287.

Wood, C.J. & Smith, E.G. 1978. Lithostratigraphical classification of the Chalk in North Yorkshire,

Humberside and Lincolnshire. Proceedings of the Yorkshire Geological Society, 42, 263–287.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 90: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Table captions

Table 1. Lowermost part of the Welton Chalk Formation: A comparison of bed nomenclature. Sample

positions are indicated.

Table 2. Summary of palynomorph concentration (counts per gramme), total organic carbon (TOC),

palynomorph groups (gymnosperm pollen; pteridophyte/bryophyte spores; angiosperm pollen;

dinoflagellate cysts; prasinophyte phycomata; acanthomorph acritarchs), separate > 10 µm kerogen

counts of 100 particles (AOM = amorphous organic matter; phytoclasts; palynomorphs) and

dinoflagellate cyst species richness (‗diversity‘). Shaded rows highlight samples with > 1 wt.% TOC.

T/M ratio = terrestrial/marine ratio. P/G ratio = the ratio between peridinioid (P) and gonyaulacoid

(G) dinoflagellate cysts.

Table 3. Relative and absolute abundance of selected gonyaulacoid dinoflagellate cysts (G-cysts), and

Dinoflagellate? type D of Ioannides (1986). P = Present, not quantified (rare occurrence, outside the

300 palynomorph count). The calculation of absolute abundance per gram of sample (cpg) is

explained in the text.

Table 4. Relative and absolute abundance of common/abundant peridinioid dinoflagellate cysts (P-

cysts). P = Present, not quantified (rare occurrence, outside the 300 palynomorph count). The

calculation of absolute abundance per gram of sample (cpg) is explained in the text.

Figure captions

Fig. 1. Turonian palaeogeographic reconstruction with the main Northern Hemisphere site locations

discussed in the text: brown shaded area, landmass; light blue, palaeo-shelf; CLIP, Caribbean large

igneous province; HALIP, high-Arctic large igneous province. Modified from Eldrett et al. (2014,

2017) and Du Vivier et al. (2015).

Fig. 2. The Cretaceous outcrop in eastern England and the location of sections discussed in the text

(modified from Hart 2019). The Black Band wedges out to the south of Louth (line A–A). It appears

to represent the ‗feather-edge‘ of OAE-2 when traced towards a palaeo-high located in the region of

the Wash. Between lines A–A and B–B the chalks and nodular chalks of latest Cenomanian – earliest

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 91: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Turonian age are dull red in colour, while south of line B–B the chalks at the same stratigraphic level

are pale green/grey in colour.

Fig. 3. Sketch map of Melton Ross Quarry, north Lincolnshire, showing the position of the

Cenomanian–Turonian boundary (CTB) succession excavations in 1997 (after Wood et al. 1997). Site

1 and Site 4 sampling exposures (located around 53º 35‘ 14‖ N, 0º 22‘ 32‖ W at 19 m above sea level)

are currently (2019) buried approximately 10 m below the restored surface of the quarry.

Fig. 4. Composite lithological log of the sampled Melton Ross Quarry section with carbon and

oxygen isotope curves. Biostratigraphy: 1, planktonic foraminiferal zones extrapolated from Elsham

and South Ferriby (Hart & Bigg 1981; Hart & Leary 1989); 2, 3, dinoflagellate cyst zones and

subzone of Olde et al. (2015a) and Dodsworth & Eldrett (2019). Formations of Wood & Smith

(1978). Members of Jeans (1980), Mitchell (2000) and Hopson (2005; new sub-member rank). The

Buckton Member is ‗The Inoceramite‘ of Hart et al. (1991, 1993) and Hart (2019). Local beds I–VII

of Wood et al. (1997), regional Beds A–H of Dodsworth (1996), Turnus Bed of Pomerol &

Mortimore (1993), ‗Adrian‘s Pair of Marls‘ of Mortimore (2014), Central Limestone of Wood et al.

(1997) and Hildreth (1999), Black Band sensu Wood & Smith (1978). Distribution of the belemnite

Praeactinocamax plenus was reported from Site 1 at Melton Ross by Wood et al. (1997). OAE-2 is

picked based on relatively high δ13C values of greater than 3‰ VPDB in this section. Points ‗a‘, ‗b‘

and ‗c‘ are tentatively correlated with the corresponding southern England picks of Jarvis et al.

(2006), allowing tentative positioning of CTB around the Bed F/G boundary.

Fig. 5. Dinoflagellate cysts and miospores from Melton Ross. All specimens were photographed at x

400. Sample/slide numbers, specimen England Finder co-ordinates, and BGS specimen registration

numbers (MPK prefix) are given. Specimens 1–16 are dinoflagellate cysts; 17–19 are gymnosperm

pollen; 20–21 are angiosperm (Normapolles) pollen. The 50 μm scale bar applies to all specimens.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 92: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

1. Litosphaeridium siphoniphorum (Cookson & Eisenack, 1958) Davey & Williams, 1966, local Bed

I, MR97-22(A), H43/0, MPK 14664;

2. Wrevittia cassidata (Eisenack & Cookson, 1960) Helenes & Lucas-Clark, 1997, local Bed II,

MR97-21(A), S43/3, MPK 14665;

3. Carpodinium obliquicostatum Cookson & Hughes, 1964, Bed C, MR97-11.5(B), X47/1, MPK

14666;

4. Adnatosphaeridium tutulosum (Cookson & Eisenack, 1960) Morgan, 1980, local Bed II, MR97-

21(B), W31/2, MPK 14667;

5. Cyclonephelium compactum Deflandre & Cookson, 1955 – Cyclonephelium membraniphorum

Cookson & Eisenack, 1962, transitional ‗complex‘ of Marshall & Batten (1988), Bed C, MR97-

11.5(D), Q36/2, MPK 14668;

6. Eurydinium saxoniense Marshall & Batten, 1988, Bed E, MR97-7(B), J42/4, MPK 14669;

7. Pterodinium crassimuratum (Davey & Williams, 1966) Thurow et al., 1988, local Bed I, MR97-

22(B), F47/3, MPK 14670;

8. Adnatosphaeridium? chonetum (Cookson & Eisenack, 1962) Davey, 1969, local Bed II, MR97-

21(A), S38/3, MPK 14671;

9. Oligosphaeridium totum Brideaux, 1971, Bed C, MR97-11.5(A), F36/0, MPK 14672. This taxon

was recorded as Litosphaeridium sp. A by Marshall & Batten (1988);

10. O. totum, Bed C, MR97-11.5(A), W42/0, MPK 14673;

11. Stephodinium coronatum Deflandre, 1936, Bed E, MR97-8.5(D), V42/2, MPK 14674;

12. Trithyrodinium suspectum (Manum & Cookson, 1964) Davey, 1969, Bed E, MR97-8(A), T35/2,

MPK 14675;

13. Palaeohystrichophora infusorioides Deflandre, 1935, local Bed II, MR97-21(A), Q28/4, MPK

14676;

14. Subtilisphaera pontis-mariae (Deflandre, 1936) Lentin & Williams, 1976, Bed C, MR97-11.5(D),

R42/1, MPK 14677;

15. S. pontis-mariae, Bed E, MR97-8.5(A), Q25/2, MPK 14678;

16. Sepispinula? huguoniotii (Valensi, 1955) Islam, 1993, Bed C, MR97-10(A), P47/0, MPK 14679;

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 93: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

17. Rugubivesiculites rugosus Pierce, 1961, Bed E, MR97-8(A), R21/1, MPK 14680;

18. Alisporites microsaccus (Couper, 1958) Pocock, 1962, Bed C, MR97-11.5(A), F38/1, MPK

14681;

19. Classopollis spp., Bed G, MR97-2(C), P36/2, MPK 14682;

20. Atlantopollis microreticulatus Krutzsch, 1967, Bed D, MR97-10.5(C), E35/3, MPK 14683;

21. Complexiopollis spp., Bed E, MR97-6(C), N41/4, MPK 14684.

Fig. 6. Dinoflagellate cysts from Melton Ross. All specimens were photographed at x 400.

Sample/slide numbers, specimen England Finder co-ordinates, and BGS specimen registration

numbers (MPK prefix) are given. The 50 μm scale bar applies to all specimens.

1. Aptea? spongireticulata (Prössl, 1990, ex Prössl, 1992) Fensome et al., 2019, Bed B, MR97-12(D),

E32/3, MPK 14685;

2. A.? spongireticulata, Bed D, MR97-9(A), P37/1, MPK 14686;

3. Ginginodinium? sp. A of Prauss (2006, 2012a), Bed D, MR97-10(C), T42/3, MPK 14687. The

arrow indicates a probable suture between two precingular plates;

4. Cyclonephelium membraniphorum Cookson & Eisenack, 1962, Bed E, MR97-8.5(C), U28/0, MPK

14688;

5. Heslertonia striata (Eisenack & Cookson, 1960) Norvick, 1976, Bed H, MR97-1(D), U45/0, MPK

14689;

6. Microdinium setosum Sarjeant, 1966, Bed C, MR97-11(D), K47/3, MPK 14690;

7. Ginginodinium? sp. A, Bed E, MR97-8.5(A), Q29/4, MPK 14691. The arrow indicates a probable

suture between two precingular plates;

8–9. Canninginopsis? lindseyensis sp. nov., Bed C, MR97-11(B), P44/3, MPK 14663. Holotype. 8.

Focus on ventral surface. Note the interpretation of tabulation and sulcal notch (arrow). 9. Same

specimen, focus on distal surface;

10. Ellipsodinium rugulosum Clarke & Verdier, 1967, local Bed II, MR97-21(A), N34/2, MPK

14692;

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 94: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

11. Ginginodinium? sp. A, or an endocyst of Trithyrodinium suspectum, Bed H, MR97-1(C), V29/0,

MPK 14693;

12. Leptodinium? aff. delicata (this paper), Bed C, MR97-11.5(C), R38/1, MPK 14694. The arrows

indicate the possible lateral position of a cingulum;

13. Dinoflagellate? type D of Ioannides (1986), Bed C, MR97-8(B), U29/0, MPK 14695. Note the

presence of three or four possible precingular plates and two intercalary or apical plates on the upper

part of the cyst;

14. Dinoflagellate? type D, local Bed II, MR97-20(A), S47/4, MPK 14696. Note the presence of at

least four ‗plates‘ that resemble a precingular series on the upper part of the cyst and further sutures

on the lower part of the cyst.

Fig. 7. Dinoflagellate cysts from Melton Ross. All specimens were photographed at x 400.

Sample/slide numbers, specimen England Finder co-ordinates, and BGS specimen registration

numbers (MPK prefix) are given. The 50 μm scale bar applies to all specimens.

1. Trithyrodinium? sp. A (this paper), Bed E, MR97-8(C), W29/0, MPK 14697. Note the sub-

polygonal/peridinioid shape, laevigate surface and tabulation interpretation, including a prominent

suture/break above the three intercalary plates;

2. Trithyrodinium? sp. A, Bed E, MR97-8(A), R32/2, MPK 14698. Note the sub-

polygonal/peridinioid shape, tabular sutures and some probable breakage in the upper-right part of the

cyst;

3. Kalyptea spp., Bed C, MR97-11.5(C), V30/0, MPK 14699;

4. ?Dissiliodinium globulus Drugg, 1978, Bed E, MR97-8(C), N39/4, MPK 14700. Polar–dorsal

epicystal compression with sutures between some dorsal precingular plates, plus a possible one plate

archaeopyle with attached operculum;

5. Bosedinia cf. sp.1 of Prauss 2012(b), Bed G, MR97-2(A), M37/4, MPK 14701. Two overlapping

specimens. Note the presence of omphali and absence of a kalyptra;

6. Leberidocysta chlamydata (Cookson & Eisenack, 1962) Stover & Evitt, 1978, local Bed II, MR97-

21(A), J47/1, MPK 14702;

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 95: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

7. Canninginopsis? lindseyensis sp. nov., Bed C, MR97-11(B), W33/2, MPK 14703. Note the

interpretation of precingular tabulation and sulcal notch (arrow);

8. ?D. globulus, Bed E, MR97-8(B), G27/2, MPK 14704. There are two precingular plates with

sutures on the specimen and an archaeopyle involving some detached dorsal precingular plates;

9. Bosedinia cf. sp.1 of Prauss 2012(b), Bed E, MR97-8.5(A), U39/4, MPK 14705. Note the

operculum, probably including apical and three intercalary plates;

10. Bosedinia cf. sp.1 of Prauss 2012(b), or an endocyst of Subtilisphaera spp., Bed G, MR97-2(C),

O39/4, MPK 14706;

11. Bosedinia laevigata (Jiabo, 1978, ex He & Qian, 1979) He, 1984, Bed E, MR97-8(A), R31/0,

MPK 14707;

12. B. laevigata, Bed C, MR97-11(E), H57/3, MPK 14708. Note the operculum, including apical and

probable intercalary 2a and 3a plates (indicated);

13. B. laevigata, Bed C, MR97-11.5(A), W40/1, MPK 14709;

14. Trithyrodinium maculatum sp. nov., Bed C, MR97-11.5(A), E45/0, MPK 14710. Endocyst. Note

the three anterior intercalary plates archaeopyle with at least two attached operculum plates, and the

endocyst wall marked by ring-shaped indentations;

15. T. maculatum, Bed C, MR97-10(A), B49/2, MPK 14711. Endocyst. Focus on probable intercalary

plates at top of specimen;

16. Fromea amphora Cookson & Eisenack, 1958, Bed C, MR97-11.5(A), N32/1, MPK 14712;

17. B. laevigata, Bed C, MR97-11.5(A), J43/3, MPK 14713;

18. T. maculatum, Bed F, MR97-3(A), S32/3, MPK 14662. Holotype. Note comparable ring-shaped

indentations on endophragm and periphragm. The likely position of the three intercalary plate

archaeopyle (operculum plates detached) is indicated;

19. T. maculatum, Bed B, MR97-12(D), K29/2, MPK 14714. Endocyst. Note the two plate intercalary

archaeopyle (2a [isodeltaform hexa-type] and 3a plates detached);

20. Sentusidinium ringnesiorum (Manum & Cookson, 1964) Wood et al., 2016, Bed C, MR97-

11.5(A), K42/2, MPK 14715;

21. B. laevigata, Bed E, MR97-8(B), X34/0, MPK 14716.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 96: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Fig. 8. Kerogen (>10 μm fraction) photographs from Melton Ross. All fields of view are at 200 x 200

μm. Photographs 2, 11 and 12 are from stained kerogen slides.

1. Bed G, MR97-2(K), O42. Note the dominance of Amorphous Organic Matter (AOM);

2. Bed F, MR97-4(A), N42. Note the dominance of dinoflagellate cysts (d.c.) and small phytoclasts;

3. Bed F, MR97-5(K), N42. Note the presence of translucent (brown) phytoclasts and AOM;

4. Bed E, MR97-8(K), O42. Note the dominance of AOM;

5. Bed E, MR97-8.5(K), O42;

6. Bed D, MR97-9(K), O42. Note the prominence of dinoflagellate cysts, including Spiniferites

ramosus (S.r.);

7. Bed C, MR97-11(K), O42. Note the dominance of AOM;

8. Bed C, MR97-11.5(K), N42;

9. Bed B, MR97-12(A), O43. Note dinoflagellate cysts, including Oligosphaeridium totum (O.t.),

phytoclasts and pyritic mineral material;

10. Local Bed VII, MR97-15(A), O43;

11. Local Bed II, MR97-21(A), O42. Note dinoflagellate cysts, including Palaeohystrichophora

infusorioides (P.i.), Subtilisphaera spp. (Subt.) and Leberidocysta chlamydata (L.c.);

12. Local Bed I, MR97-22(A), O41. Note dinoflagellate cysts, including Litosphaeridium

siphoniphorum (L.s.).

Fig. 9. Van Krevelen-type diagram for the Melton Ross samples analysed for Rock-Eval pyrolysis.

Blue-filled circles are data for the > 1 wt.% TOC samples. Open circles are data for the < 1 wt.%

TOC samples; the latter are unreliable, due to probable elevation by occluded CO2 within carbonate of

inorganic origin, in addition to that derived from early diagenesis of organic matter.

Fig. 10. Lithological succession in the Bridge Creek Limestone Member of the Greenhorn Formation,

on the north side of the Pueblo Reservoir State Recreation area, west of Pueblo, Colorado, United

States. This section contains the Global boundary Stratotype Section and Point (GSSP) for the base of

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 97: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

the Turonian Stage (38º 16‘ 56‖ N, 104º 43‘ 39‖ W) and the proposed GSSP for the base of the

Middle Turonian Substage (Bengtson et al. 1996; Kennedy et al. 2000, 2005; Dodsworth & Eldrett

2019). (a) metres above / depth below base of Bridge Creek Member; (b) chronostratigraphy; (c)

lithostratigraphy; (d) lithology; (e) δ13Corg, data from Bowman & Bralower (2005). Peaks ‗a‘, ‗b‘ and

‗c‘ in the OAE-2 excursion follow Jarvis et al. (2006); (f) summary of proposed regional correlation

with sections from NW Europe (this paper); (g) Pueblo ammonite zones after Kennedy & Cobban

(1991), Cobban (1993) and Kennedy et al. (2000). The Watinoceras devonense, Pseudaspidoceras

flexuosum and Vascoceras birchbyi units are often treated as subzones of a Watinoceras coloradoense

Zone or Watinoceras spp. Zone in published literature; (h) Pueblo inoceramid bivalve zonation by

Walaszczyk & Cobban in Kennedy et al. (2000); (i) Pueblo planktonic foraminiferal zones are from

Caron et al. (2005) and Keller & Pardo (2004); (j) subzones are from Keller & Pardo (2004). There is

a difference between Keller & Pardo (2004) and Caron et al. (2005) on the base of

Helvetoglobotruncana helvetica. This difference is marked by the vertical line in the figure; (k)

nannofossil zones are from Bralower (1988) and Bralower & Bergen (1998); (l), (m) dinoflagellate

cyst zones and subzones are from Dodsworth & Eldrett (2019). This diagram is adapted from

Dodsworth & Eldrett (2019). The main sources for NW Europe correlation are: Melton Ross, Wood &

Mortimore 1995, Wood et al. 1997, this paper, C.L. = Central Limestone; Misburg HPCF II quarry,

Hilbrecht 1986, Hilbrecht & Hoefs 1986, Prauss 2006, P.P. = pre-Plenus succession, P.B. = Plenus

Bed, triple = ‗triple band‘; Eastbourne, Jarvis et al. 2006, Pearce et al. 2009, Beds 1–8 compose the

Plenus Marls, the overlying Mead Marls occur within the Ballard Cliff Member and the Holywell

Marls occur within the Holywell Member.

Fig. 11. Scatter diagram plotting 13C -18O co-variance for the Melton Ross section to help elucidate

the primary versus diagenetic signals. Black circles = > 0.2 wt.% TOC; white circles = < 0.2 wt.%

TOC. Reference fields plotted from Moore (1989), with fields of well-preserved foraminifera (after

Veizer & Prokoph 2015; O‘Brien et al. 2017) most likely reflecting unaltered 13C -18O values

indicative of primary Late Cretaceous sea-water. Field of bulk rock 13C -18O values from CTB

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 98: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

sections after Jarvis et al. (2011, 2015). Regression lines and R2 values plotted for all samples (black

line) and low TOC samples between samples MR97-11 and MR97-23 (dashed line).

Fig. 12. The stratigraphy of the Cenomanian–Turonian boundary interval in Norwegian Continental

Shelf, North Viking Graben exploration well 35/6-2 S (‗Grosso‘), drilled by StatoilHydro AS in 2009.

The proposed correlation to onshore outcrops is discussed in the main text. Shetland Group

lithostratigraphy of Gradstein & Waters (2016), dinoflagellate cyst zones and subzone of Olde et al.

(2015a) and Dodsworth & Eldrett (2019). The last occurrence (LO) of Carpodinium obliquicostatum

is taken here as a proxy for the top of the Adnatosphaeridium tutulosum Subzone. FO = First

Occurrence; NWG = NW Germany; GR = Gamma Ray, American Petroleum Institute (API) units.

END OF MANUSCRIPT

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 99: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

CLIP

CentralAtlantic

Western InteriorSeaway

30°N

60°N

0°NTrans-Saharan

Seaway

NeoTethys

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 100: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

0 km 30

0 miles 20

Nor thSea UK

Speeton

Bridlington

Welton

South FerribyElsham

Caistor

Louth

Tetford

SouthThoresby

ClaxbyWelton le Marsh

HunstantonHeacham

Hillington

Marham

HULL

KING’S LYNN

NORWICH

LINCOLN

A

BA

B

Cenozoic

Chalk & Red Chalk

Gault Clay

Lower Cretaceous

Pre-Cretaceous

EastKnapton

Flixton

BishopWilton

MarketWeighton

47/10-1

Bigby MeltonRoss

WashThe

BGS Trunchborehole

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Paul Dodsworth
Typewritten Text
Figure 2
Page 101: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

3

Deepexcavations,as numbered by Wood et al. (1997) Trench

Old Melton Rossquarries

Working areas in 1997 and location of deep excavationsand the trench, exposing the expanded CTB succession; the MR97 samples -1 to -13 were collected from Site 4 while samples -14 to -23 were collected from Site 1

Old Melton Rossquarries

B121

1

railway

0 125 250

metres

375 500

N

4

21

Figure 3

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 102: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

A

B

C

E

F

D

G H

0

1

2

23

222120

19

18.5

18

17

16

15

14

13

12

13

2

45

6 -11.5

m Location: Melton Ross Quarry

Sample MR97-

BlackBand

VII

?VI

III

?III

?V-?IV

Lithostratigraphy Lithology

Figure 4

2 3

δ13C carbonate

a

?b

c

4

-6 -3 -4 -5

1 2 3

Biostrat.Age

Key: A. P. M. = Adrian’s Pair of MarlsF. C. F. = Ferriby Chalk Formation Bivalvia (inoceramid) debris Pyrite nodule Prominent macrobioturbation Prominent microbioturbation

(Chondrites)

Composite log of 1997 deep excavations, Site 1 (base of section to Central Limestone)and Site 4 (Bed B to top of section)

Chondrites II

Chondrites I

δ18O carbonate

5

-7

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 103: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

9 11

50μm

Figure 5

7

21 3

65

4

10

18

12 14 15

17 19 20 21

8

13 16ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 104: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

50μm

Figure 6

9

21

4

10

765

11

12 13 14

8

3

1IlI

1II

1IIII

2IlI

1c2c

lp

3c4c

5c6c

3IlI4IlI

2II4II5II ?6II

5IlI?6IlI5IlI

?s

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 105: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

11

Figure 7

7

21 3

65

4

10

19

12

16 17

18 20 21

13

15

50μm

14

9

8

1a 3a2a2a

3a

2a

3a

2a

1a2a

3a

1a 3a

1IIII

3II2II4II

5II6II1II

2a 3a

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 106: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Figure 8

8

21 3

654

10 11 12

97

O.t.

pyrite

L.s.

L.c.

P.i.

Subt.

pyritic material

brownphytoclast

blackphytoclast

AmorphousOrganic Matter

S.r.

brownphytoclast

d.c.

AOM

blackphytoclast

MR97-2Bed G

MR97-4Bed F

MR97-5Bed F

MR97-8, Bed E MR97-8.5Bed E

MR97-9Bed D

MR97-12Bed B

MR97-11.5Bed C

MR97-11Bed C

MR97-15Bed VII

MR97-21Bed II

MR97-22Bed I

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 107: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

0

100

200

300

400

500

600

700

800

900

1000

0 10 20 30 40 50 60 70 80 90 100

HY

DR

OG

EN

IND

EX

(H

I, m

g H

C/g

TO

C)

OXYGEN INDEX (OI, mg CO2/g TOC)

TYPE I

TYPE II

TYPE IV

TYPE III

l

II

III

IV

Geochemical Services Group, 143 Vision Park Blvd., Shenandoah, Texas 77384 • Phone: 281-681-2200 • Fax: 281-681-0326 • Email: [email protected]

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Paul Dodsworth
Rectangle
Paul Dodsworth
Typewritten Text
Figure 9
Page 108: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

125

122120118116113

109

105

101

97

90

868584

79

69

6463

12

11

10

9

8

7

6

5

4

3

2

1

0

-1

-2

-3

-28 -26 -24 -22

δ13Corg

a. c.b. d. e.

AmmoniteZones

f.

Planktonicforaminifera

Zones

h.

Calcareousnannoplankton

Zones

j.

Nigericas scotti(inferred)

P. flexuosum

V. birchbyi

Mytiloidesmytiloides

Mytiloideskossmati

Mytiloides puebloensis

Mytiloides hattini

Inoceramuspictus

InoceramidZones

g.

Adnatosphaerid. tutulosum

Isabelidinium magnum

Senoniasphaera turonica

Cyclonephelium membraniphorum

Dinoflagellatecyst

Zones

k.

Dinoflagellatecyst

Subzones

l.

LitosphaeridiumsiphoniphorumRotalipora

cushmani

A. tutulosum

Dicarinellahagni

Heterohelixmoremani

G. bentonensisRotaliporaExtinction

?

Mytiloidessubhercynicus

PlanktonicforaminiferaSubzones

i.

Axopodorhab.albianus

Rhagodiscusasper

Eprolithus florialis

Collignoniceraswoollgari

Mammitesnodosoides

Neocardiocerasjuddii

Sciponocerasgracile

Metoicocerasmosbyense

Watinocerasdevonense

Helvetoglobotrunc.helvetica

Whiteinella archaeocretac.

NW Europecorrelation

m.

P.B.a

b

c

Figure 10

C.L.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 109: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

ACCEPTED MANUSCRIPT

by guest on August 28, 2020

http://pygs.lyellcollection.org/D

ownloaded from

Page 110: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

35/6-2 SField: Grosso

Location: Norwegian CSOperator: StatoilHydro AS

Scale: 1:750

Din

ofla

ge

llate

Cys

t S

ub

zon

eA

dn

ato

sp

ha

eri

diu

mtu

tulo

su

m

0 120APIGR

Biostratigraphic Comments

LO Carpodinium obliquicostatum

FO common Cyclonephelium compactum-membraniphorum. Eurydinium saxoniense & Alterbidinium daveyi

LO common / consistent Litosphaeridium siphoniphorum, LO Pterodinium crassimuratum

LO Pterodinium cingulatum granulatum

Pa

lyn

olo

gic

al

Sa

mp

les

2607 D/C

2616 D/C

2625 D/C

2646 D/C

2655 D/C

2664 D/C

2673 D/C

Fo

rma

tion

Try

gg

vass

on

Blo

do

ksS

vart

e

De

pth

, M

etr

es

2600

2610

2620

2630

2640

2650

2660

2670

Co

rre

lativ

eO

utc

rop

Lith

o-U

nit

Fish Shale (NWG) /Beds C - H

(eastern England)

Bed B

Plenus Bed (A)

pre-Plenus

Din

ofla

ge

llate

Cys

t Z

on

eC

yc

lon

ep

he

liu

m m

em

bra

nip

ho

rum

Lit

os

ph

ae

rid

ium

sip

ho

nip

ho

rum

Ag

eE

arly

Tur

onia

nLa

te C

enom

ania

nStratigraphic Summary Chart

Figure 12

ACCEPTED MANUSCRIPT

by guest on August 28, 2020http://pygs.lyellcollection.org/Downloaded from

Page 111: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Locality: Melton Ross. Mortimore Hildreth Wood et Dodsworth Wood & Jeans et al. Jeans et al. Hart et al. JefferiesLithostratigraphy and sampling. This paper 2014 1999 al . 1997 1996 Mortimore 1991 1991 1991 & 1963(Not to scale.) 1995 1993

Melton Ross S. Ferriby Bigby M. Ross S.F./Flixton M.R./S.F. Flixton S. Ferriby S. Ferriby S. FerribyMember Lithology Sample Bed Sub‐unit

limestoneMR97‐1 Bed H gungy marl Unit 8 Bed 10 Bed H Bed 6 Bed 8 Bed vMR97‐2 Bed G Bed 9 Bed G ?15 to ?17 Bed 7

light MR97‐3 Bedsmarl MR97‐4 Bed F Unit 7 Bed 8 Bed F Bed 5 11 to ?14 Bed 6 Bed iv Bed iv

Flixton MR97‐5Member MR97‐6(Black dark MR97‐7 Bed E Bed 7c Bed E Bed 10

Band Sub‐ marl MR97‐8

member) MR97‐8.5 BlackMR97‐9 Bed D Band Unit 6 Bed 7b Bed D Bed 4 Bed 9 Bed 5 Bed iii Bed iiiMR97‐10 C3 C3 Bed 8MR97‐10.5 Bed C C2 Bed 7a Bed C C2 Bed 7

dark MR97‐11 C1 C1 Bed 6marl MR97‐11.5

MR97‐12 B4 Unit 5 Bed 6b (u) Bed 3 Bed 5light silty Bed 6a

Flixton marl MR97‐13 Bed B B3 khaki Unit 4 Bed 5b Bed B Bed 2 Bed 4 Bed 4 Bed ii Bed iiMember B2 marl Unit 3 Bed 5a (l)(Plenus B1 Bed 4

Marls Sub‐ limestone MR97‐14 A3 (CL) Unit 2 (CL) Bed 3 (CL) lst limestone Bed 3 (lst)member) Bed A A2 Unit 1 Bed 2 Bed A Bed 1 Bed 3 Bed 2 Bed i Bed i

A1 Bed 1 Bed 2MR97‐15 to 22 Beds I‐VII Beds I‐VII A. S.

Louth Mbr. limestone MR97‐23 Ferriby Chalk Formation Bed 1 Bed 1

Key: A.S. = 'Anomalous Succession'; CL, = 'Central Limestone'; lst = limestone (chalk); l = lower; u = upper. = Stratigraphic breakTable 1. Lowermost part of the Welton Chalk Formation: A comparison of bed nomenclature.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020

http://pygs.lyellcollection.org/D

ownloaded from

Page 112: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Locality: Melton Ross.(Not to scale.) Kerogen Terrigenous palynomorphs Marine palynomorphsLithostrat. Lithology Sample Counts per TOC AOM P'clasts Palyn. Gymnosp. Pterido. Angiosp. T/M P/G Dinoflag. Prasinoph. Acantho. Dinoflag.

g (total) % % % % % % % ratio ratio % % % diversitylimestone

Bed H MR97‐1 34,320 0.21 71 12 17 9.7 0.3 < 0.3 0.1 0.24 81.7 5.7 2.7 63Bed G MR97‐2 89,760 3.47 92 4 4 11 2.3 0.3 0.137 0.34 68.3 6.3 11.7 62

light MR97‐3 6,688 0.18 58 13 29 3 < 0.3 < 0.3 0.03 0.21 95.3 0.7 0.7 69Bed F marl MR97‐4 1,470 0.12 9 49 42 < 0.3 < 0.3 0 0 0.31 98.7 1.3 0 53

MR97‐5 8,668 0.34 41 33 26 2.7 < 0.3 0 0.027 0.16 96.3 0.7 0.3 67MR97‐6 17,204 0.4 42 30 28 5 1 0.3 0.063 0.25 91.3 1.3 0.7 79

Bed E dark MR97‐7 30,360 0.67 31 33 36 9.7 0.3 < 0.3 0.1 0.36 82.7 4.7 2.3 68marl MR97‐8 85,184 2.18 83 8 9 9.7 2.6 < 0.3 0.123 0.32 77.3 6 4 71

MR97‐8.5 219,648 2.74 71 11 18 12.3 0.3 < 0.3 0.127 0.35 82.7 1.7 2.7 72Bed D MR97‐9 23,518 0.36 19 30 51 < 0.3 0 0 0 0.03 100 0 0 63

MR97‐10 123,024 2.19 81 5 10 10 0.3 0.3 0.107 0.29 86.7 1 1.7 72Bed C MR97‐10.5 53,284 0.7 27 29 44 0.3 0 < 0.3 0.003 0.04 99 0 0.3 65

dark MR97‐11 101,552 1.43 63 16 21 4 < 0.3 0.7 0.047 0.31 92.3 0.7 2.3 79marl MR97‐11.5 87,296 2.2 71 15 14 7.3 < 0.3 0.3 0.077 0.35 90.7 0.7 1 76

Bed B MR97‐12 288 0.2 0.3 0 0 0.003 0.01 99.7 0 0 33MR97‐13 28 0.12 1.4 0.4 0 0.018 0.06 96.5 0 0.7 33

Bed A limestone MR97‐14 <1 0.1 No reliable kerogen  2MR97‐15 0 0.15 counts from sample ‐12 0

Bed VII MR97‐16 <1 0.11 to sample ‐19, due to 1light MR97‐17 27 0.14 sparse recovery and  0 0 0 0 0 100 0 0 4marl MR97‐18 5 0.13 pyritic material. 0 0 0 0 0 100 0 0 1

Bed ?VI MR97‐18.5 8 0.21 0 0 0 0 0 100 0 0 2Bed ?IV‐V MR97‐19 88 0.14 0 0 0 0 0 100 0 0 1Bed ?III MR97‐20 95 0.16 1 89 10 0.3 0 0 0.003 0.72 99.3 0.3 0 32Bed II MR97‐21 12,408 0.28 2 26 72 0.7 < 0.3 < 0.3 0.007 0.45 98.3 1 0 77Bed I MR97‐22 53 0.12 4 90 6 1 0 0 0.01 0.66 97.3 1.7 0 33

limestone MR97‐23 N/A 0.04Table 2. Summary of palynomorph groups, > 10 micron kerogen counts and total organic carbon (TOC). Shaded rows highlight samples with > 1 % TOC.

ACCEPTED MANUSCRIPT

by guest on August 28, 2020

http://pygs.lyellcollection.org/D

ownloaded from

Page 113: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Locality: Melton Ross.(Not to scale.) Gonyaulacoid dinoflagellate cystsLithostrat. Lithology Sample    L. siphoniphorum C. compactum‐memb.    Pterodinium  spp.    Spiniferites  spp.          O. totum Dinoflagell.? type D

% c.p.g. % c.p.g. % c.p.g. % c.p.g. % c.p.g. % c.p.g.limestone

Bed H MR97‐1 0 0 42.7 14655 0.7 240 6 2059 0 0 1.3 458Bed G MR97‐2 0 0 11.7 10502 2.3 2064 9.3 8348 0 0 0.7 598

light MR97‐3 0 0 13 869 10 669 28.3 1893 < 0.3 P 2 134Bed F marl MR97‐4 0 0 19.7 290 3.7 54 18.7 274 < 0.3 P 9 132

MR97‐5 0 0 21.7 1881 4 347 27.7 2401 0.7 58 4 347MR97‐6 0 0 18 3097 3.7 637 22.3 3836 < 0.3 P 1.7 287

Bed E dark MR97‐7 0 0 15.3 4645 3.7 1123 14.7 4463 0 0 1 304marl MR97‐8 0 0 15.7 13374 4 3407 16.7 14226 < 0.3 P 0.3 284

MR97‐8.5 0 0 23.3 51178 2.7 5930 9.3 20427 0 0 0.3 732Bed D MR97‐9 0 0 35 8231 3.3 776 33 7761 1.3 306 1.3 314

MR97‐10 0 0 23.7 29157 3 3691 12 14763 < 0.3 P 0.3 410Bed C MR97‐10.5 0 0 32.0 17050 4.7 2504 25.7 13694 0.3 158 0.7 355

dark MR97‐11 0 0 16 16248 4 4062 25.4 25794 0.7 656 0.7 677marl MR97‐11.5 0 0 13 11348 5 4365 17.3 15102 3.3 2612 1 873

Bed B MR97‐12 ? < 0.3 ?P 13.7 39 < 0.3 P < 0.3 1 40.3 116 16.3 47MR97‐13 0 0 29.3 8 0 0 23.3 7 5 1 1.1 <1

Bed A limestone MR97‐14 0 0 0 0 0 0 0 0 0 0 0 0MR97‐15 0 0 0 0 0 0 0 0 0 0 0 0

Bed VII MR97‐16 0 0 0 0 0 0 0 0 0 0 50 <1light MR97‐17 0 0 0 0 0 0 0 0 0.4 1 98.1 26marl MR97‐18 0 0 0 0 0 0 0 0 0 0 100 5

Bed ?VI MR97‐18.5 0 0 0 0 0 0 0 0 0 0 94.7 8Bed ?IV‐V MR97‐19 0 0 0 0 0 0 0 0 0 0 100 88Bed ?III MR97‐20 1.7 2 0.3 1 1 1 5.4 5 0 0 48 46Bed II MR97‐21 2.7 335 < 0.3 P 4.7 583 29 3598 0 0 1 124Bed I MR97‐22 1 1 0.3 1 4.3 2 13.3 7 0 0 3 2

limestone MR97‐23Table 3. Relative and absolute abundance of selected gonyaulacoid dinoflagellate cysts (G‐cysts), and Dinoflagellate? type D of Ioannides (1986).ACCEPTED M

ANUSCRIPT

by guest on August 28, 2020

http://pygs.lyellcollection.org/D

ownloaded from

Page 114: Proceedings of the Yorkshire Geological Society2020/05/22  · England, U.K. The section is investigated for organic geochemistry and stable isotopes for the first time, while a detailed

Locality: Melton Ross.(Not to scale.)       Peridinioid dinoflagellate cystsLithostrat. Lithology Sample      P. infusorioides   S. pontis‐mariae     E. saxoniense Bosedinia cf. sp. 1     T. suspectum Ginginodinium?  sp.A

% c.p.g. % c.p.g. % c.p.g. % c.p.g. % c.p.g. % c.p.g.limestone

Bed H MR97‐1 < 0.3 P 5 1716 4.7 1613 1 343 7 2402 1.7 583Bed G MR97‐2 0.3 269 4 3590 8 7180 5.7 5116 3 2693 1 898

light MR97‐3 11.3 756 3.3 221 1.7 114 0.3 20 2 134 0.7 47Bed F marl MR97‐4 9.3 137 5.7 84 1.7 25 0.3 4 4 59 0.3 4

MR97‐5 7.7 667 2.7 234 1.7 147 2 173 0.3 26 0 0MR97‐6 9 1548 5.3 912 2.7 465 2 344 2 344 0.7 120

Bed E dark MR97‐7 3.3 1002 8.7 2641 5.3 1609 5 1518 4.7 1427 1.7 516marl MR97‐8 3.7 3152 10.7 9115 2.3 1959 3 2556 2.7 2300 1 852

MR97‐8.5 1.7 3734 18.7 41074 2 4393 2.3 5052 2.3 5052 1.3 2855Bed D MR97‐9 1 235 0.7 165 0.3 71 0 0 0.7 165 0 0

MR97‐10 1 1230 14.7 18085 1 1230 6 7381 1.7 2091 < 0.3 PBed C MR97‐10.5 1.3 693 < 0.3 P 0.3 160 0 0 1.7 906 0 0

dark MR97‐11 5 5078 16.7 16959 0.7 711 1.3 1320 4 4062 0 0marl MR97‐11.5 3 2619 18.7 16324 1 872 2 1746 5 4365 0.3 262

Bed B MR97‐12 0 0 < 0.3 P 0 0 0 0 1 3 0 0MR97‐13 3.2 9 0.7 1 0 0 0.4 1 1.1 1 0 0

Bed A limestone MR97‐14 0 0 0 0 0 0 0 0 0 0 0 0MR97‐15 0 0 0 0 0 0 0 0 0 0 0 0

Bed VII MR97‐16 0 0 0 0 0 0 0 0 0 0 0 0light MR97‐17 0 0 0 0 0 0 0 0 1.1 1 0 0marl MR97‐18 0 0 0 0 0 0 0 0 0 0 0 0

Bed ?VI MR97‐18.5 0 0 0 0 0 0 0 0 0 0 0 0Bed ?IV‐V MR97‐19 0 0 0 0 0 0 0 0 0 0 0 0Bed ?III MR97‐20 13.3 13 17.7 17 0 0 0 0 1.3 1 0 0Bed II MR97‐21 32 3971 10 1241 0 0 < 0.3 P 0.7 87 < 0.3 PBed I MR97‐22 13 7 41.3 22 ?1 ?1 0 0 2.3 1 0.3 1

limestone MR97‐23Table 4. Relative and absolute abundance of common/abundant peridinioid dinoflagellate cysts (P‐cysts).ACCEPTED M

ANUSCRIPT

by guest on August 28, 2020

http://pygs.lyellcollection.org/D

ownloaded from