You are on page 1of 101

(‘wtacxwus Research ( 1988) 9, 3- 103

Microfossil Assemblages and the Cenomanian-


Turonian (late Cretaceous) Oceanic Anoxic Event

I. Jarvis
Sc~hool of Geological Sciences, Kingston Pol>*technic Rood, Kingston
, Pc~nrh~~n upon Thntne~,
Surre_v KTl 2EE, U.K.

G. A. Carson
Department of Earth Sciences, lJniz~ersit_v of Liewpool, Brozonlow Street. PO Box 147.
T,ieterpool. L69 3BX, U.K.

M. K. E. Cooper
Strntigraphic Services International Ltd., Chancellor Court, 20 Priestly Road, Guildford.
.Surre>,, GU2 5 YL, U.h’.

M. B. Hart, P. N. Leary and B. A. Tocher


Depnrtmrnt of Geological Sciences, Pl~~moutk Polytrc.knic, Drake Cirrus, Piyntouth. Dworc.
PI,4 8,424, C. h-.

D. Home
Department of Geology, City of London Pol>stechnic, H’nlhurgh IIozrse, Bigland Street. London
El 2,VG, l.:.h-.

and A. Rosenfeld
Geological Surr,e\l of Israel, 30 ;%falkhe Yisrael Street. Jerusalem 95 501, Israel.

Rec,ceiT,eti28 -4rtgust I987 and in revised form 6 Janunrj, I988

I. Jar\G, G. A. Carson, M. K. I-. Cooper, RI. B. Hart, P. ,X. Lear?, B. .4. ‘rocher, D. Horne
and X. Kosenfeld. Microfossil .4ssemhlages and the Cenomanian-Turonian (late Cretaceous)
Oceanic Anoxic Event. Cretaceous Research (1988) 9, 3-103. The effects of the Cenomanian-
‘I’uronian Oceanic Anoxic Event (OAE) in the Chalk Sea of NW Europe have been
investigated using published macrofossil records combined with new detailed sedimentolog-
ical. foraminiferal, ostracod, calcareous nannofossil, dinoflagellate cyst and stable-isotope
data from Dover, England. The ranges of individual fossil species are displayed against
lithostratigraphic logs, and their relation to the Cenomanian-Turonian houndarv (defined
using macrofaunal data) is discussed. .4 positive carbon stable-isotope excursion, indicating
the stratigraphic extent of the OAE, spans the stage boundary. Correlation with successions
elsewhere in NW Europe suggests that the OAE was isochronous, and that major hiostrati-
graphic marker horizons are characterised by distinctive 6°C values. All microfossil groups
display uppermost Cenomanian abundance and diversity minima which correspond closely to
the peak of the carbon stable-isotope excursion. We propose that the OAE was a phase of
increased upwelling which led to a widespread espansion and intensification of the oxygen-
minimum zone in the oceans. As a result, increasingly dysaerohic bottom waters developed
Lvithin the Chalk Sea, and were responsible for progressive disappearances in the benthonic
microfaun?,including the extinctions of many typical Cenomanian taxa. At the same time the
oxygen-mlmmum zone rose in the water column, causing the extinction of deeper water
planktonic foraminifera and then the gradual loss of intermediate-\vater groups. A temporar!
disappearance of dinoflagellate cysts and a proliferation of calcispheres were associated v ith
these events. As the OAE waned, new species eraduallv evolved to fill niches left vacant
following the extinctions of Cenomanian &a. The appearance of these new species defines the
base of the Turonian. the stage division beinn a direct conseauence
I, .,
of the OAE. It is concluded
the 0.41:s provide a major mechanism for controlling rates of evolution and extinction
throughout the Phanerozoic.

KIX \~ORDS: Cretaceous; Cenomanian; Turonian: Chalk; Oceanic Anoxic Event; Foramini-
fera; Ostracoda; Xannofossils; Dinoflagellates; Correlation: Stratigraph!; Palaeocean-
ography: Oxygen-Minium Zone: Extinction; Evolution.
Contents
1. Introduction (I.J.)

2. Lithostratigraphy and sedimentology (IJ)


2.1 Abbots Cliff Chalk Formation
2.2 Plenus Marl Formation
2.3 Dover Chalk Formation
2.4 Discussion

3. Macrofossil biostratigraphy (IJ)


3.1 Abbots Cliff Chalk Formation
3.2 Plenus Marl Formation
3.3 Dover Chalk Formation
3.4 Definition of the Cenomanian-Turonian boundary

4. Foraminifera (M.B.H. and P.N.L.)


4.1 Material and methods
4.2 Biostratigraphy
4.2.1 Abbots Cliff Chalk Formation
4.2.2 Plenus Marl Formation
4.2.3 Dover Chalk Formation
4.3 Discussion

5. Ostracoda (D.H. and A.R.)


5.1 Material and methods
5.2 Biostratigraphy
5.2.1 Abbots Cliff Chalk Formation
5.2.2 Plenus Marl Formation
5.2.3 Dover Chalk Formation
5.3 Taxonomic notes
5.4 Discussion

6. Calcareous nannofossils (M.K.E.C.)


6.1 Material and Methods
6.2 Biostratigraphy
6.2.1 Abbots Cliff Chalk Formation
6.2.2 Plenus Marl Formation
6.2.3 Dover Chalk Formation
6.3 Discussion

7. Dinof-lagellate cysts (B.A.T.)


7.1 Material and methods
7.2 Biostratigraphy
7.2.1 Abbots Cliff Chalk Formation
7.2.2 Plenus Marl Formation
7.2.3 Dover Chalk Formation
7.3 Discussion

8. Stable-isotope geochemistry (G.A.C. and I.J.)


8.1 Material and methods
8.2 Stratigraphic trends
8.3 Possible diagenetic factors

9. Stratigraphic position and correlation of the OAE (I.J. et al.)


9.1 Integration of stable isotope and biostratigraphic data from Dover
9.2 Correlation with the NE Atlantic continental margin
9.3 Correlation with SE Devon
9.4 Correlation with Humberside and the North Sea
Microfossil Assemblages and the Oceanic Anoxic Event 5

9.5 Correlation with Germany


9.6 Stratigraphic conclusions and lateral variation

10. Palaeoceanography and the OAE (1.J. et al.)


10.1 Evidence for a Cenomanian-Turonian upwelling event
10.2 Causes and consequences of the upwelling event.

11. Effects of the OAE on the Chalk Sea biota (I. J. ef al.)
11.1 Microfossil abundance and diversity
11.2 Oxygenation levels and the biota
11.3 Expansion of the oxygen-minimum zone into the Chalk Sea
11.4 Extinction. evolution and the OAE

12. Conclusions (I.J. et al.)

1. Introduction

Upper Cenomanian to low Turonian marine sediments display anomalous


lithological, geochemical and/or fauna1 characteristics worldwide (see
Graciansky et al., 1986; Herbin et al., 1986; Stein et al., 1986; Schlanger
et al., 1987 for recent reviews). Oceanic areas and many marginal basins
contain dark-coloured (very dark gray to black) laminated organic-rich mud-
rocks of this age. Benthonic faunas are commonly impoverished or absent
in these sediments, which locally contain high concentrations of micro-
plankton. This ‘black shale’ facies is generally absent in shallow shelf
environments, but here, limestones display anomalous carbon stable-
isotope signatures. Cenomanian-Turonian boundary sequences typically
display abnormally high 613C values of +4-S%, (PDB), compared with
+ 2-3x, for the immediately underlying and overlying limestones. These,
and other features, have been interpreted as indicating that significant
parts of the world ocean were intermittently oxygen deficient, and that
oxygen depletion reached a maximum across the Cenomanian-Turonian
boundary (90.5-91.5 Ma ago; Harland et al., 1982). This episode was
termed the Cenomanian-Turonian “Oceanic Anoxic Event” (OAE) by
Schlanger and Jenkyns (1976).
In this paper we describe the Cenomanian-l’uronian boundary succession
at Dover, England (Figure 1). The east Kent sections currently constitute the
best documented Chalk succession in SE England, and are the stratotypes for
the recently proposed lithostratigraphic scheme for the North Downs
(Robinson, 1986). Macrofossil records are reviewed, and the position of the
stage boundary precisely defined on detailed lithostratigraphic logs. We have
studied the sedimentology and the foraminiferal, ostracod, calcareous
nannofossil and dinoflagellate cyst biostratigraphies of the boundary se-
quence. These are described and possible relationships between biostrati-
graphy and the OAE are investigated using detailed 613C and 6180 profiles.

2. Lithostratigraphy and sedimentology (I. J .)

The Upper Cenomanian to lower Turonian succession in SE England


displays several marked lithological changes. These changes have recently
N
t

DOVER

West
Houghton
-I

2km
..
&FOLKESTONE
Microfossil Assemblages and the Oceanic Anoxic Event 7

been expressed in terms of a formal lithostratigraphic nomenclature (Figure


2), erected for the North Downs by Robinson (1986). Rhythmically bedded
marly chalks of the upper Abbots Cliff Chalk Formation (Middle to Upper
Cenomanian) are sharply overlain by a distinctive sequence of marls and
marly chalks (Figure 3), which constitute the Plenus Marl Formation (Upper
Cenomanian). The Plenus Marl is in turn overlain by a succession of nodular
and intraclastic chalks (Figure 3) which form the basal part of the Dover
Chalk Formation (uppermost Upper Cenomanian to high Turonian).

2.1 Abbots Cliff Chalk Formation

The Cenomanian of SE England and northern France characteristically


displays well-developed decimetre-scale alternations of friable marl and
indurated chalk (Jukes-Browne and Hill, 1903; Kennedy, 1969; Carter and
Hart, 1977; Robaszynski et al., 1980). The contrast between the two
lithologies decreases upwards from the base of the sequence, reflecting an
overall decline in the clay content of the sediments. Rhythmicity is less
pronounced in the upper part of the succession, where individual rhythms
are also typically thicker (0.5-l m).
A distinctive sequence of beds containing prominent decimetre-scale
laminated calcarenitic chalk-filled scour structures, occurs in the upper part
of the Middle Cenomanian (Whitaker et al., 1872; Jukes-Browne and Hill,
1903; Kennedy, 1969; Carter and Hart, 1977, Robinson, 1985, 1986;
Robaszynski and Amidro, 1986). The base of these beds corresponds to a
marked reduction in the prominence of rhythmic bedding, and was used by
Robinson (1986) to define the base of his Abbots Cliff Chalk Formation
(Middle-Upper Cenomanian). Robinson (1986, p. 148) distinguished two
units within the Abbots Cliff Chalk (Figure 2), a lower sequence of beds
containing abundant scour-fill structures (Hay Cliff Member), and an upper
succession of weakly rhythmic marly chalks in which these structures are rare
or absent (Capel-le-Ferne Member). The Abbots Cliff Formation is 22m
thick in the formational stratotype at Abbots Cliff path, west of Dover
(Figure 1).
Only the uppermost 6m of the Capel-le-Ferne Member were sampled
during the present study. Petrographically these chalks are fossiliferous
micrites (Figure 4) containing scattered (5lO?i,) silt and very fine sand-sized
bioclasts, principally complete and fragmented benthonic and planktonic
foraminiferal tests, dissociated prisms of calcitic bivalve shells, and calci-
spheres. The non-carbonate content is approximately So/&(Jefferies, 1963a).

2.2 Plenus Marl Formation

The Abbots Cliff Chalk is overlain by a thin (2.5 m at Dover) succession of


marls and marly chalks (Figure 3) which form a prominent marker in the
Upper Cenomanian throughout southern England, northern France and the
North Sea Basin. The distinctive nature of this sequence was first recognised
by Phillips (1818, p. 46-7) who described a “bed of marle” at the base of
Shakespeare Cliff, west of Dover (Figure 1). These beds were studied by
many later workers (Whitaker et al., 1872; Hebert, 1874; Barrois, 1876;
Price, 1877; Hill and Jukes-Browne, 1886; Jukes-Browne and Hill, 1903),
(sled) (sied)
utyuoml ue~uk?uloua~

(Sled)
~ivw kmoa

I I

(wed)
WY3 JaM0-l
blicrofossil Assemblages and the Oceanic Anosic Event 9

Figure 3. The Cenomanian-Turonian boundary sequence on Abbots Cliff path, Dover. The bases of the
Plenus Marl Formation and Melbourn Rock Beds (Shakespeare Cliff Member, Dover Chalk) are
indicated by the lower and upper discontinuous white lines, respectively. Numbered subdivisions of
the Plenus Marl indicated by the white numerals are the bed numbers of Jeff&es (1962, 1963q h).
Locations of samples collected for microfossil and stable isotope analysis are marked by the black-
filled circles and black numerals (samples ABCS-21). The base of the Turonian. defined by the
appearance of &,rytiloides spp. inoceramid bivalves coincides with sample ABC19.
-
Microfossil .&xemblages and the Oceanic .-\nosic Event 11

who referred to them as the “Belemnite Marl” or “Zone of Belemnitella


plena”. The term “Plenus Marls” was introduced by White (1909) and has
now been widely adopted for the unit (Rawson et al., 1978; Wright and
Kennedy, 1981; Figure 2).
The Plenus Marl has been assigned formational status both in the North
Sea Basin (Deegan and Scull, 1977; Burnhill and Ramsay, 1981; Hart and
Bigg, 1981; Hancock, 1986) and, more recently, in SE England (Robinson,
1986). The stratigraphy of the Plenus Marl was studied in detail by Jefferies
(1962, 1963a, 6), who erected the section at Merstham Greystone Limeworks
(Surrey) as the stratotype. Unfortunately, this site has subsequently been
filled, and no exposures are currently available. The Shakespeare Cliff and
Abbots Cliff sections west of Dover (Figure l), however, are virtuall)
identical to those exposed previously at hlerstham (Jefferies, 1962, 1963a),
and are also the localities where the unit was first recognised (Phillips, 18 18).
‘l’hese sections, therefore, are regarded here as the neostratotype for the
formation, although thicker. sequences are de\reloped near Eastbourne, in
Sussex (Jukes-Browne and Hill, 1903; Jefferies, 1962; Mortimore, 1986).
‘rhe base of the Plenus Alar1 is a prominent omission surface, picked-out
b\, the dark gray colour of the overlying clay-rich marls (Figure 3) which fill a
diverse suite of Thalassinoides, Chorzdrites and Plnnolites burrows. In
attenuated successions, significant erosion can be demonstrated at this le\-el
(Jefferies 1962, 1963a; the sub-plenus erosion surface of authors). Jefferies
(1962, 1963a, 6) defined a sequence of eight beds within the Plenus Marl.
These beds were numbered 1 to 8 in ascending order (Figure 3), with
laterally extensive subdivisions being distinguished by letter suffixes
(e.g. 1 a, 1 b). Individual beds were defined (Jefferies op. c-it.) by their litholog)
and macrofossil content. ,411 eight beds can be recognised in basinal chalk
successions throughout the Anglo-Paris Basin (Jefferies, 1963a, Figure lo),
although some beds are absent in thin ‘marginal’ developments of the
formation. In addition to the major omission surface defining the base of the
Plenus Marl, a prominent omission surface is commonly visible at the
bottom of Bed 4. Other omission surfaces occur locally at the bases of Beds 2,
6 and 8, although these are rarely well-developed.
‘I’he base of the Plenus Marl marks a major petrographic change. In
addition to the obvious increase in clay content (a non-carbonate value of

Figure 4. Photomicrographs illustrating major lithofacies within the Cenomanian-Turonian boundary


succession at Dover. Scale bars =0.5 mm (500pm). (a) Abbots Cliff Chalk, ABCS, a fossiliferous
micrite containing scattered silt and very fine sand-sized bioclasts, principally planktonic foramini-
fera and calcispheres; (b) Plenus Marl Bed 1, ABC7, a biomicrite with a wackestone (locally
packstone) texture containing about 40?, fine to medium sand-sized bioclasts, particularly
inoceramid bivalve prisms, foraminifera and ostracods; (c) Plenus Marl Bed 4, ABC1 1, a biomicrite
with a wackestone texture containing around 35 “iI silt to fine sand-sized bioclasts, principally
calcispheres, inoceramid bivalve prisms and a few foraminifera; (d) Plenus Marl Bed 7, ABC13, a
well-sorted packed biomicrite composed predominantly (N 6O’b) of calcispheres, including Pith-
onella spp., a possible calcareous dinoflagellate cyst; (e) Plenus Marl Bed 7, ABC13, a lower
magnification view of (d), illustrating a typical upper Plenus Marl lithofacies; (f) Dover Chalk
Melbourn Rock Beds, AKS4, a less well-sorted packed biomicrite consisting of a calcisphere-rich
matrix containing abundant sand-sized and larger inoceramid bivalve debris and scattered foramini-
fera; (g) Dover Chalk Melbourn Rock Beds, AKS4, a higher magnification view of(f). Note the high
proportion of calcispheres. including Pithonella spp; (h) Dover Chalk, mid Shakespeare Cliff
Member, AKSlO, a sparser calcisphere-biomicrite, typical of the higher beds of the lower Turonian
at Dover.
19”,, was recorded for Bed la at Dover by Jefferies, 1963a), the mean grain-
size of the carbonate fraction is also considerably higher. Lower beds
(1 and 2) consist of biomicrites with wackestone textures (Figure 4) contain-
ing an average of around 300,,, fine to medium sand-sized fossil debris. The
bioclasts are moderately to well-sorted and consist predominantly of com-
plete and fragmented benthonic and planktonic foraminifera, ostracods,
and cornminuted calcitic bivalve-shells. The distribution of the bioclasts
is irregular (some patches contain > SO’,, bioclasts), probably related to
the concentration of sand-sized material in burrow fills.
An increase in carbonate occurs in Bed 3 (< 7O,, noncarbonate; Jefferies,
1963~). This is accompanied by a transition to sparser biomicrites containing
fewer (20°0) and finger-grained (coarse silt-sized) bioclasts. These are
chiefly calcipheres with subordinate foraminifera and calcitic bivalve-shell
prisms. Higher clay-contents in Beds 4-6 (12-29”;,; Jefferies, 1963~) are
accompanied by renewed increases in the proportions (up to 40°sb) and
grades (up to fine sand-size) of the bioclasts, and higher proportions of
planktonic foraminifera and bivalve-shell debris. The summit of the Plenus
Marl (Bed 7; Bed 8 was not analysed petrographically) again contains more
carbonate (8’ o non-carbonate; Jefferies, 1963~) but consists of a distinctive
well-sorted packed biomicrite containing - 60”,, coarse silt-sized bioclasts-
almost entirely calcispheres (Figure 4). Scattered bivalve-shell fragments are
also present, but there are few foraminifera. ‘l’he upper beds (4-8) are
characterised by well-developed compactional and pressure-solution fabrics,
which in thin section produce a strong bedding-parallel alignment to the
bioclasts.

2.3 Doz)er Chalk Formation


A sequence of moderately indurated intraclastic nodular hardgrounds
immediately overlies the Plenus Marl (Figure 3). This dramatic change in
lithology marks the base of the Dover Chalk Formation (uppermost Upper
Cenomanian to high Turonian). The term ‘Dover Chalk’ was introduced by
Dowker (1870) as the exact equivalent of Phillips’s (1818, 1821) “Chalk with
few flints”. This original definition of the formation was extended by
Robinson (1986) to include part of the underlying chalk succession, down to
the top of the Plenus Marl. The Dover Chalk now corresponds approxi-
mately to the “Middle Chalk” (Figure 2) as applied in east Kent by previous
workers (Hill, 1886; Jukes-Browne and Hill, 1903; Rawson rt al., 1978;
Wright and Kennedy, 1981).
The Dover Chalk is 66.9 m thick at Akers Steps west of Dover (Figure l),
the formational stratotype. The formation is characterised by very pale grey
marly chalks with common flaser (stylonodular) structures (cf. Garrison and
Kennedy, 1977), and occasional well-deveioped centimetre-thick marl
seams, which have been traced laterally throughout the North Downs
(Robinson, 1986). The lower part (basal 26.5 m at Akers Steps) of the Dover
Chalk is highly nodular and contains many intraclastic beds, although well-
indurated hardgrounds occur only in the basal 1.2 m. These nodular chalks
form a distinctive unit (Phillips, 1818, 1821; Drew in Whitaker et al., 1872;
Hebert, 1874; Barrois, 1876; Woodroof, 1981), called the Shakespeare Cliff
Microfossil Assemblages and the Oceanic Anosic Event 13
--___ -_

Member (cf. “Craie conglomeree de Shakespeare Cliff” of Barrois, 1876) by


Robinson (1986). We have studied only the basal 21.5 m of the member in
detail.
The bottom 10.8 m of the Shakespeare Cliff Member was termed the “Grit
Bed” by Price (1877) and the “Melbourn Rock” by Hill (1886), terms used
by many subsequent workers (Rowe, 1900; Jukes-Browne and Hill, 1903;
White, 1928; Smart et al., 1966; Woodroof, 1981; Robinson, 1986). These
beds contain abundant limonite-stained chalk-pebble intraclasts and coarse
calcarenitic shell debris, particularly cornminuted inoceramid with occa-
sional echinoderm tests. Their composition corresponds, therefore, to
original definitions of the Melbourn Rock (Jukes-Browne, 1880), and
Robinson (1986). We ha\re studied only the basal 21 ..j m of the member in
detail.
Despite having much higher carbonate contents (probably < 10(, non-
carbonate)pthe basal hardgrounds of the Melbourn Rock are petrographi-
tally very similar to the packed calcisphere-biomicrites which characterise
the top of the Plenus Marl. Cement is predominantly micritic with occasional
patches of neomorphic microspar. Microspar infills of microfossil tests are
also common. The calcisphere content is marginally Iower (- 500/,), princi-
pally due to the absence of a pressure-solution associated concentration of
the coarser carbonate fraction. The remainder of the Melbourn Rock also
consists of calcisphere biomicrites (Figure 4), but additionally contains
around lo”, medium to very coarse sand-sized (and larger) bioclasts,
principally cornminuted inoceramid bival1.e shells. The proportion and size
of the larger bioclasts decreases towards the top of the unit.
Above the Melbourn Rock, the Shakespeare Cliff Member consists of
sparser calcisphere-biomicrites (Figure -1) containing <20°a bioclastic
material. Calcispheres remain dominant but are accompanied by occasional
fine sand-sized foraminifera and scattered bivalve-shell fragments.

2.4 Discztssiort

The Plenus Marl is differentiated from the Abbots Cliff Chalk not only by its
higher clay-mineral content, but also by containing an increased proportion
of silt-sized and coarser grained bioclasts. These observations, together with
the prominence of omission surfaces in the sequence, suggest that removal of
micritic (clay-sized) carbonate by current winnowing was probably a major
process during the deposition of the Plenus Marl. VVinnowing would not
necessarily lead to the removal of clay minerals to the same extent as clay-
grade carbonate. The former possess charged surfaces which cause them to
form silt-sized floes that behave as coarser grained aggregates rather than
individual particles. Winnowing could, therefore, concentrate a range ofsilt-
sized components, both clay minerals and bioclasts.
An alternative hypothesis is that both the proportions of clay minerals and
the coarser-grained carbonate fraction have been increased by the disso-
lution of clay-grade carbonate. Dissolution is certainly not a significant
process in the lower Plenus Marl (Beds l-3), where calcareous microfossils
(including nannofossils) are very well preserved. Above this, however, in
Beds 4-8 and in the basal metre of the Melbourn Rock Beds, microfossils
display evidence of etching. We interpret this dissolution as having been
dominantly a seafloor process (Section 11.3), but later diagenetic dissolution
processes may have had some effect on the composition of the lithofacies.
A major increase in the proportion of calcispheres is seen towards the top
of the Plenus Marl, and calcispheres remain a major component throughout
the Melbourn Rock. The increase in the proportion of calcispheres is a result
of several factors. Firstly, winnowing and dissolution have led to the removal
of clay-sized carbonate. Secondly, there is a significant decline in the
abundance of foraminifera, which form a major component of the coarse-
grained fraction of the lower beds. Finally, calcispheres must have been more
abundant at that time, since the proportion of other bioclasts (e.g. bivalve and
echinoderm debris) is also low.
There are three principal features which differentiate the Melbourn Rock
from the summit of the Plenus Marl, below. Firstly, there is a major decline
in the percentage of clay minerals. Secondly, there is the addition of micritic
cements, both within hardgrounds and isolated nodules. Thirdly, there is an
increase in the mean size of the bioclastic fraction due to the addition of
significant proportions of coarse-grained inoceramid bivalve shell debris.

3. Macrofossil biostratigraphy (I. J .)

Fauna1 records for the Dover successions are available from several sources,
particularly Jefferies (1962, 1963a, b) and Kennedy (1969) for the upper
Abbots Cliff Chalk, Jefferies (op. cit.) and Wright and Kennedy (1981) for the
Plenus Marl, and Jukes-Browne and Hill (1903) and Woodroof (1981) for the
basal Dover Chalk. These records are used to define the position of the
Cenomanian-Turonian boundary in the succession, and provide a biostrati-
graphic framework with which our new microfossil and stable-isotope data
can be compared.

3.1 Abbots Cliff Chalk Formation

Macrofossils are uncommon in the upper beds of the Abbots Cliff Form-
ation, although abundant inoceramid bivalve debris and fragments of thin-
shelled echinoids (Holaster sp.) occur at some levels. Fauna1 records
(Jefferies, 1962, 1963a, b; Kennedy, 1969; A. S. Gale, pers. comm., 1987)
indicate that Holaster trecensis Leymerie and the bivalves, Inoceramus pictus
J. de C. Sowerby, Pycnodonte ztesicularis (Lamarck) and Neithea quin-
quecostata (J. Sowerby) are the commonest fossils. Other records include the
bivalves, Plagiostoma globosa (J. de C. Sowerby), Plicatula inflata J. de C.
Sowerby, Chlamys beaveri (J. Sowerby), Entolium orbiculare (J. Sowerby),
the brachiopods Terebratulina (dominantly T. striata Wahlenberg), Lingula,
Concinnithyris, and the worm tube ‘Serpula’ umbonata Mantell. No am-
monites are recorded from these beds in the coastal successions, although by
comparison with inland exposures (Kennedy, 1969), the sequence is regar-
ded as being Late Cenomanian Calycoceras guerangeri Zone (Figure 2). The
Late Cenomanian age is confirmed by the occurrence of Inoceramuspictus (cf.
Troger, 1981).
Microfossil Assemblages and the Oceanic Anoxic Event 15
_

3.2 Plenus Marl Formation

No detailed macrofaunal analysis of the Plenus Marl has been published


which is specific to the Dover succession, although Jefferies (1962, 1963a, b)
presented a model which integrated data from throughout the Anglo-Paris
Basin (including Dover). The macrofaunal records presented here, there-
fore, are generalised distributions which have yet to be substantiated in detail
at Dover. Nevertheless, these records enable a broad comparison to be made
with our detailed microfossil and calcareous nannofossil data.
The fauna of the basal Plenus Marl (Bed 1) is very similar to that of the
underlying formation. The main changes which occur at the base of Bed 1
(Jefferies, op. cit.) are that Plicatula injata disappears, Holaster treceks
occurs only rarely, and Orbirhynchia multicostata Pettitt becomes common.
In addition, Pycnodonte vesicularis and Neithea quinquecostata are larger in
Bed 1 than in the underlying chalks. A range of other fauna is also recorded
(Jefferies, op. cit.).
There is an abrupt and extensive change in the macrofauna at the base of
Bed 2. This is associated with, but certainly not limited to, the appearance of
aragonitic groups in the assemblage. Nineteen species disappear at this level
(Jefferies, 19636), including the bivalves Plagiostoma globosa and Neithea
quinquecostata, the brachiopods Lingula and Orbirhynchia multicostata, the
starfish Trachyaster sp. cf. T. rugosus (Spencer) and the echinoid Holaster
trecensis. Seventeen species of calcitic-shelled macrofossils appear or become
common in Bed 2, notably bivalves, Plicatula barroisi Peron, Entolium
membranaceum (Nilsson), echinoids Cidaris perornata Forbes, Yjilocidaris
hirudo (Sorignet), Discoidea minima Agassiz, starfish Calliderma smithiae
Forbes, and the brachiopod ‘Rhynchonella’ lineolata Phillips var. carteri
Davidson. In addition, a number of aragonitic-shelled groups appear,
including ammonites Metoicocerasgeslinianum (d’orbigny), Scaphites sp. cf.
S. equalis J. Sowerby, bivalves Nucula sp., ProtocardiahiIlana sp., Gramma-
todon sp. cf. G. cenomanezse (d’orbigny), gastropods Solaria sp., Cerithium
sp., Aporrhais sp., and the scaphopod Dentalium sp.
The upper Plenus Marl (Beds 3-8) contains faunas which are similar to,
but less abundant than, that of Bed 2. In addition, however, Beds 4-6 contain
common belemnites Actinocamax plenus (Blainville) and the bivalves
Oxytoma seminudum (Dames), Aequipecterr arlesiensis (Woods), while Bed 7
yields abundant Orbirhynchia wiesti (Quenstedt). Aragonitic fossils in parti-
cular, are generally less common in Beds 4-8 than in Beds 2-3.
Xmmonite records (Jefferies, 1962, 1963a, 6; Wright and Kennedy, 198 1)
include common Metoicoceras geslinianum which appears in oyster-mould
preservation in the upper part (A. S. Gale pers. comm., 1987) of Bed 1 and
ranges throughout the formation (Wright and Kennedy, 1981, Table 1).
Calycoceras naviculare Mantel1 appears in the Abbots Cliff Chalk and ranges
up to Bed 3 of the Plenus Marl. Beds 2--3 have also yielded Puzosia odiensis
Kossmat, Anapuzosia dibleyi (Spath), Vascoceras diartianum (d’orbigny)
and Scaphites sp. cf. S. equalis. Sciponoceras gracile (Schumard) appears m
Bed 4 and ranges up into the basal Dover Chalk, while Bed 5 has yielded
Pseudocalycoceras dentonense (Moreman). The uppermost Plenus Marl (Beds
7-8) contains Euomphaloceras septemseriatum (Cragin) and Hamites sp.
These ammonite assemblages together with Sumitomoceras cautisalbae
Ll’right and Kennedy (known from SE Devon) was used by Wright and
Kennedy (1981, p. 118) to detine theil- Late Cenomanian Metoicoceras
geslinianum Zone (Figure 2). ,4 I,ate Cenomanian age is confirmed by the
common occurrence of Znocevamus pictus which ranges throughout the
formation.

3.3 Doz’er Chalk Formation

The hardground sequence at the base of the Melbourn Rock Beds yields
Inocevamus pictus, Sciponocevas hohemicum ante&s Wright and Kennedy,
Tylocidaris hirudo spines, pycnodonteine oysters and hexactinellid sponges
(Woodroof, 198 1; Wright and Kennedy, 198 1). Inoceramus pictus ranges up
to just over 1 m above the top of the Plenus Marl. Immediately above this,
inoceramids \vith the shape of MJltiloides but possessing an external orna-
ment similar to that of I. pictus occur in association with true Mytiloides spp.
(C. J. Wood, pew. comm., 19X7), particularly AZ. mytifoides (Mantell).
A4Jltiloides labiatus (Schltitheim) appears in the upper part of the Shake-
speare Cliff Member, immediately above the Melbourn Rock (Woodroof,
1981).
‘I’he Melbourn Rock at Dover is best known for its abundant echinoid
fauna (Price, 1877; Rowe, 1900; Jukes-Browne and Hill, 1903), particularly
Discoidea dixoni Forbes and C’avdiaster pygmaeus Forbes. Salenia granulosa
Forbes, Tylocidaris hirudo and rare Glzrphocyphus radiatus Hoenig also occur,
and ~‘o~~ulussubvotwzdus Forbes and 6. ‘castanea’ (Brogniart) make their first
appearances towards the summit of these beds. This echinoid fauna occurs
throughout the Shakespeare Cliff Member, but is uncommon above the
Melbourn Rock.
‘I’he calcarenitic component of the Melbourn Rock primarily consists of
comminuted inoceramid shells with subordinate echinoid debris. However,
the brachiopods Orbirhynchia cuzievi (d’orbigny), Concinnithvris sp.,
Terebratuli?la sp. and hexactinellid sponges are common throughout the
Shakespeare Cliff Member. The kival\res PIicatula sp., Plagiostoma sp. and
Lima sp. are also recorded.
Ammonites are uncommon in SE England above the Plenus Marl,
although Sciporzoceras bohemicum antevius and S. gvacile occur in the basal
part (see above). Watinorevas sp. is recorded from Merstham (Wright and
Kennedy, 1981), while Mammites twdosoides (Schliiter) and Lezoesiceras
pevamplunz (Mantell) occur in the higher beds of the Melbourn Rock at
Do\,er, and rarely in the upper part of the Shakespeare Cliff Member.

3.4 DeJinition of the Cenomanian- Tzwonia~l boundav>

It is generally accepted (Birkelund et al., 1984) that ammonites and


inoceramid bivalves provide the best means of defining the Cenomanian-
‘I’uronian boundary. ‘I‘he base of the Pseudaspidocerasflexuosum Zone (top of
the Neocardioceras juddii Zone) has gained some acceptance with Tethyan
ammonite workers (Cobban, 1984), while the appearance of a Watinoceras
coloradoense Zone ammonite assemblage is generally used in northern
Europe (Wright and Kennedy, 1981; Kennedy et al., 1983; Robaszyn-
ski, 1980, 1982; Kennedy, 1984). Unfortunately, although Watinoceras is
hlicrofossil .&emblages and the Oceanic Anosic Event 17

common, the nominate subspecies W. coloradoense coloradoense (Hender-


son) is absent in Europe and, in addition, the base of the zone is drawn at a
lower level than the base of the zone of the same name in the U.S.A. (Cobban,
1984). This lower level probably corresponds approximately to the base of
the P. _jkxuosum Zone (Birkelund et al., 1984).
‘I’he base of the P.j?exuosum Zone of North America is characterised by the
appearance of common M~~tiloides opalensis senszt Kauffman non B&e [but
above the first occurrence of M. sp. aff. LV. submvtiloides (Seitz) sensu
Kauffman; Kauffman et al., 1977; Hancock, 19841. The widespread ap-
pearance of early Mytiloides spp. at this level is an event which can be traced
throughout the Tethyan and Boreal regions (Birkelund et al., 1984; Ernst
et al., 1984; Hancock, 1984).
Common Sciponoceras and the absence of Metoicocevas geslinianum in the
basal hardgrounds of the Melbourn Rock at Dover suggest that the
hardgrounds lie within the latest Cenomanian Neocardioceras juddii Zone
(I&Tright and Kennedy, 1981; Figure 2). Records of Watilzoceras from
IL’Ierstham, however, indicate that the higher beds of the Melbourn Rock are
earliest ‘I’uronian VV. coloradoense Zone. L\mmonite records are insufficient
to position the stage boundary more precisely.
I,ate Cenomanian Irzocrramus pictus are also common in the basal hard-
grounds, and ,tiloides spp. appear immediately above them. This major
change in the inoceramid biva1L.e fauna is used here to define the base of the
‘I’uronian in east Kent.

4. Foraminifera (M.B.H. & P.N.I,.)

Planktonic and benthonic foraminiferal assemblages were studied in detail


across the boundary sequence. A total of 18 planktonic and 48 benthonic
species and \.arieties were identified (Table l), with all samples yielding some
specimens identifiable to specific level. In addition, one further planktonic
species \vas recorded from immediately above the succession which was
studied in detail. Dramatic changes occur in the abundance, specific
composition and diversity of these assemblages immediately below the stage
boundary.

4.1 .Ilaterial and methods

Thirty-four 2 kg samples were collected from a 27 m thick sequence, for


foraminiferal, ostracod, calcareous nannofossil and stable-isotope analysis.
The uppermost Abbots Cliff Chalk, the Plenus Marl, and the basal Dover
Chalk were sampled (numbers ABC l--21) at Abbots Cliff path (National
Grid Reference TR268385; Figure 3), 5 km WSW of Dover (Figures 1, 3).
The remainder of the lower Dover Chalk was sampled (numbers AKSl---13)
at Xkers Steps (TR29739-1) at the western end of Shakespeare Cliff, 2 km
WSW of Dover (Figure 1). Sections exposed at these localities overlap
considerably (Robinson, 1986), but the Cenomanian-Turonian boundary
sequence is only easily accessible across the intervals sampled here.
Material was collected at approximately metre-intervals, with sample
density being increased to around 10 cm towards the top of the Plenus Marl
and base of the Dover Chalk (Figure 3). The 2 kg samples were split in the
18 I. Jarvis et nl.

Table 1. AlphabetIcal list of foraminiferal species and their stratigraphic distribution AC, Abbots Cliff
Chalk [upper Capel-le.Ferne Member]; PM, Plenus Marl Formation; DC, Dover Chalk Formation
[Shakespeare Cliff Member]).

Species Distribution

Ammodiscus cretaceus (Reuss) AC. PM, DC


Arenobulimina sp. A AC
A. advena (Cushman) AC, PM
A. bulleta Barnard and Banner AC
A. preslii (Reuss) DC
Astrocolus sp. A AC, PM
Ataxophragmium depressurn (Perner) AC, PM, DC
Dentalina sp. A AC
D. sp. B AC, PM
D. sp. C PM
Dicarinella algeriana (Caron) AC, PM, DC
D. hugni (Scheibnerova) AC, PM, DC
D. imbricato (Mornod) PM, DC
Dorothia gradata (Berth&n) AC
Eggerellina breois Marie AC. PM
E. mariae Ten Dam AC, PM
Frondicularia sp. A DC
F. cordai Reuss AC
F. sp. cf. F. gaultina Reuss AC, DC
Gaoelinella baltica Brown AC, PM
G. berthelini (Keller} (= G. tourainensis Butt) AC, PM, DC
G. cenomanica (Brown) .4C, PM
G. intermedia (Berth&n) AC, PM
G. reussi (Khan) ilc, PM
Guembifitra cenomana (Keller) XC
Guembilitriella sp. A AC
Gyroidinoides parva (Khan) AC
‘Hedbergella’ aprico Loeblich and Tappan AC, PM, DC
‘H.’ archaeocretacea (Pessagno) DC
‘H.’ brittonensis (Loeblich and Tappan) AC, PM, DC
H. delrioensis (Carsey) AC, PM, DC
H. planispira (Tappan) PM
H. simplex (Morrow) PM
Helvetoglobotruncana praehelvetica (Trujillo) DC
H. helvetica (Bolli) DC
Heierohelix globulosa (Ehrenburg) DC
H. moremani (Cushman) 4C. PM, DC
Lenticulina rotulata (Lamarck) var. A AC. PM, DC
L. rotulata var. B AC, PM, DC
L. mu/am var. c DC
I,it~gulogae~elinella aumalensis (Sigal). DC
L. globosa (Brotzen) AC, PM. DC
Marginotruncana marginata (Reuss) DC
Marssonella sp. A PM. DC
.U. trochus (d’orbigny) AC, PM, DC
M. trochus var. osycona (Reuss) AC, PM, DC
M. trochus var. turris (D’Orbigny) AC, PM, DC
hieojlabellina sp. A PM
Nodosaria sp. A AC, PM
Oolina sp. A AC
Pondaglandulina sp. A AC. PM
Plectina cenomana Carter and Hart XC’, PM
P. mariae (Franke) AC
Praeglobotruncana stephani (Gandolf) AC, PM
Pseudospiroplectinata plana (Gorbenko) PM
Ramulina aculeata (d’orbigny) AC
Rotalipora cushmani (Morrow) AC, PM
R. greenhornensis (Morrow) AC. PM
field into four subsets: approximately 1.5 kg for ostracods, 0.5 kg for
foraminifera, 5Og for stable-isotope analysis and 5 g for nannofossils.
Foraminifera were isolated following disaggregation of the sediment using
white spirit (Jarvis et al., 1987). A single sample (ABC15) produced no
identifiable foraminifera after white spirit processing, and required the
identification of specimens in thin section.

4.2 BiostratigraphJr

The foraminiferal assemblages of the late Cenomanian to early Turonian are


exceptionally well known, having been described by Jefferies (1962), Owen
(1970), Carter and Hart (1977), Robaszynski et al. (1980), Hart and Bigg
(1981) and Hart (1982). Zonally significant planktonic foraminifera (Robas-
zynski and Caron, 1979a, b) are abundant throughout the succession, except
where affected by the anoxic event (Hart and Ball, 1986). The distribution of
foraminifera (Figures 6-8) and their relative abundance (Figure 9) have been
used to define a precise biostratigraphy at Dover. Representative specimens
are illustrated in Figures 1 O--l 2.

4.2.1 Abbots Cliff Chalk Formation. Samples from the Abbots Cliff Chalk
(Figures 6, 7) consistently yielded abundant foraminiferal assemblages, with
over 50°, of the fauna being represented by planktonic taxa (Figure 9). The
planktonic fauna is dominated by Hedbergella spp., with rotaliporids [Rotali-
pora cushmarzi (Morrow), R. greenhornensis (Morrow)] and praeglobo-
truncanids/discarinellids each averaging only 25O, of the planktonic total.
It is important to note the appearance of the Dicarinella group, within
the uppermost 3 m of the formation (Figure 6), since this genus is norm-
ally regarded as a Turonian-Santonian indicator. The benthonic fauna
(Figures 6, 7) is of high diversity and is typical of the late Cenomanian (Zone
13 of Carter and Hart, 1977), being dominated by species of Arenobulimina
[e.g. Figure 10(a)], Gavelinella [Figures 1 O(l)-(n)], Lenticulina [Figure 10(j),
(k)], Plectina [Figure 10(g)] and Tritaxia [Figure 10(b), (c)l.

4.2.2 Plenus Marl Formation. The eight beds of the Plenus Marl (Jefferies,
1962, 1963a, 6) each contain a distinctive foraminiferal assemblage. Bed 1 is
characterised by a fauna which is very similar to that in the underlying
Abbots Cliff Chalk (Figures 6,7). The diversity of the benthonic assemblage
is also comparable, but there is an increased percentage of planktonic taxa
(Figure 9). Rotalipora cushmani [Figures 12(b), (c)] is considerably more
abundant (the genus constitutes > 50°10 o f the planktonic assemblage), and
t
cll

0”

cl
*n
0
00
0
Microfossil Assemblages and the Oceanic Anoxic Event 21

the population contains larger, gerontic individuals. These features are


considered to be indicative of a deeper-water, possibly nutrient-rich,
environment (Leary, 1987).
At the top of Bed 1 there is a dramatic change in the benthonic assemblage
(Figures 6,7), with the majority of the typical Cenomanian groups becoming
extinct. At the same time there is a shift in the GaveZineZZa reussi (Khan)-
G. berthelini (Keller) plexus towards later morphotypes. Finally, there is a
major decline in the abundance of Rotalipora (Figure 9), caused in part
by the disappearance of R. greenhornensis [Figure 12(d)]. This particular
combination of events can be correlated throughout southern England
(Carter and Hart, 1977; Hart, 1985; Leary, 1987).
The faunas of Beds 2 and 3 (Figures 6, 7) are almost identical, except that
the R. cushmani assemblage gradually declines in both size and numbers
(Figure 9). Within the same interval, however, the early dicarinellids
[Dicarinella algeriana (Caron), D. hagni (Sheibnerova)] increase in abun-
dance, compensating for the reduced number of Rotalipora, and as a result
the total percentage of planktonic individuals remains unchanged. The
benthonic fauna at this level has already acquired a ‘Turonian’ (i.e. low
diversity) aspect, with a much reduced fauna dominated by Lingulogavelin-
ells globosa (Brotzen) [Figures 1 l(m), (n)], Gavelinella berthelini (Keller)
[Figures 1 l(k), (I)] an d varieties of MarssoneMa trochus (d’orbigny) [Figures
10(e), 1 l(c)]. The appearance of Textularia chapmani Lalicker [Figure 1 l(b)]
occurs in Bed 2, and this species is restricted to the Plenus Marl.
Rotalipora cushmani disappears at the top of Bed 3 (Figure 6) and the
diversity of the benthonic assemblage continues to decline. In Bed 4 there is a
marked drop in the proportion of planktonic species (although there is no fall
in overall abundances; Figure 9), and benthonic taxa remain the dominant
proportion of the foraminiferal assemblages throughout the remainder of the
Plenus Marl. A major increase in the proportion of praeglobotruncanids and
early dicarinellids in the planktonic assemblage occurs at this level. This
diversification is believed to have been a response to the extinction of
Rotalipora, which enabled the former groups to diversity and expand into the
upper portions of the niches left vacant by that extinction. Significant drops
in abundance and diversity of both planktonic and benthonic taxa occur in
Bed 6 (Figure 9), and this trend continues within the overlying beds where
the planktonic assemblages consist almost entirely of hedbergellids. The ex-
tinction of EggereZZina spp. [e.g. Figure 10(f), 1 l(d)] in Bed 6 and the
disappearances of the last Cenomanian morphotypes of Gavelinella
[Figures 10(l)-(n)], mark the end of characteristic Cenomanian forms in the
benthonic assemblage. All specimens from Beds 4-8 show surface pitting
which is interpreted as indicating dissolution of their tests on the sea-
floor. The planktonic taxa are generally the worst affected by dissolution,
particularly in Beds 6-8.

4.2.3 Dover Chalk Formation. The cemented nodular chalks of the Mel-
bourn Rock are extremely difficult to process for foraminifera, but even in
more marly intervals (where the samples are less indurated, and larger
numbers of individuals were recovered; Figure 9), the fauna is characterised
by low numbers and low di\:ersities. The planktonic fauna is generally poorly
Planktonics

Stage

Formation

90
60

70

60

i
I I iI iI
0 0
I
I I I 0 0

Lithology

Figure 6. Foraminiferal distribution and lithostratigraphy across the Cenomanian-Turonian boundary


at Dover, I-Planktonic and some benthonic species. Additional benthonic species are shown in
Figure 7. Lithologies and symbols are explained in Figure 5. The discontinuous horizontal line
marks the top of Plenus Marl Bed 3 which is coincident with the top of the Rotalipora cushmani Total
..:.:, .$.>:.:.:::..:;: ::“:..:..

o-o-. ,::~~~~~~~
:i’:g.:yi::;I: .:: : .:::;, :i:::::
g
0
,,::, : ::::i .,.,:.,_
::. ::::,,:::....>..y.,
a
-
.- -m- - --y$&l!y.-
,.a_
,‘,.
:.. ::;

e:
.,. : ” /
.-
o-o- l -.-~

o-o-o-o-o

o-o-o-o-a-o:
l - l- . - . -

0-o-a’ -----a-
o-o-*
Cenomanian Turonian
ABBOTS CLIFF CHALK DOVER CHALK

Capel-le-Ferne Member Shakespeare Cliff Member

,.::fy;
‘:‘:~.,:i?J,~:,::~ .,.
D+--- o-o-o rritaxia mactedyenr

l- .---.-o-o r. pyramldata

o---*-o Gare,,“e,,a berthe,;“,

0. ceno*an,ce
o-o-o-o
0 GyrOidlnOideS paw.3

o-e- -- -o-o- Marrsonella troEh”S

_-_-----. *taiophr*Pmi”m depresr”m


Microfossil Assemblages and the Oceanic Anoxic Event 25

0 n i c s

preserved in the basal beds, but above this diversities gradually increase,
caused by the’ appearances of ‘Hedbergella’ archaeocretacea (Pessagno)
[Figure 12(h)] and Dicarinella spp. [including D. imbricata (Mornod)]. Also
appearing in the MeIbourn Rock Beds are low-spired hedbergellids that
develop faint marginal keels [Helvetoglobotruncana ‘praehelvetica’ (Truj-
illo)]. There is then a complete gradation between such early forms and true
(Boreal) Heln~etoglobotrzuzcanan helvetica (Bolli). The transition may not be
synchronous with such changes in ‘l’ethyan areas, as this evolution is
probably at least in part a function of available water depth (Hart and Bailey,
1979; Hart, 1982). Marginotvuncana marginata (Reuss) [Figures 12(i), (j)]
Turonian
DOVER CHALK

Shakespeare Cliff Member

r .yr:>.‘)
,;_.+ ~;,‘:;:’
0 ,
2 \’ ..J_’ . . ..
@ 8
E
l
9 . ;
-0-o
I

I
. - - - - ..- - - -
--m-v ----

-o-0

--me- --- ___----.

-0 0 ; o-o-.
,
I .-•-

I
XIicrofossil .%ssrmblages and the Oceanic Xnosic Event 27

Benthonics

I
I
I
I
0

I
0

i
I
I
0
I
I
I
1 0

::,
I
I I
I

3
‘.

7
I
.: I “.j I
I _,
: :

:
:..
’ I

3
1

_- I _. .- .- ; .,__
-I -
I
0 1 I
I
I
I
I
I
I
.I
I
I
I I
i 0
I I
I I I I
-

U~!LlOJfl~ ueluelUoua~

I I

,
: I
Microfossil Assemblages and the Oceanic Anosic Event 29

appears above the Melbourn Rock, in the higher beds of the Shakespeare
Cliff Member (Figure 8). The first occurrence of true H. helvetica (Figure
12e-g) was noted immediately above the top of the Shakespeare Cliff
Member, in the Warren Marls (lithostratigraphic terminology of Robinson,
1986), approximately 7 m above the top of our detailed sampling (Figure 8).
Planktonic assemblages in the upper beds of the Shakespeare Cliff Member
are characterised by increasing proportions of praeglobotruncanids and
dicarinellids, and declining percentages of hedbergellids (Figure 9). This
changing population structure is considered to be a response to increasingly
favourable, probably deeper water, conditions during the early Turonian.
The benthonic fauna is of low abundance and diversity throughout the
Shakespeare Cliff Member, although there is a gradual increase in diversity
towards the top of the unit (Figure 9). Particularly noteworthy are the
appearances of Arenobulimina preslii (Reuss) [Figure 12(l)], which occurs in
the Melbourn Rock, and Tritaxia tricarinata var. jongmansi Schijfsma
[Figure 12(k)] and LinguZogaveZin~ZZa aumalensis (Sigal) [Figure 12(m)],
which are first recorded in the higher beds of the member (Figure 8).

4.2.4 Discussion. The foraminiferal biostratigraphy of the Cenomaniani


Turonian boundary in England and France was discussed by Carter and
Hart (1977), who erected a series of parallel planktonic and benthonic
zones for the Cenomanian. These zones were based on type sections
at Dover (Shakespeare Cliff borehole) and Merstham (Plenus Marl sec-
tions at Greystone Limeworks, cf. Jefferies, 1962, 1963a, b). Overviews of
the foraminiferal changes which occur across the stage boundary were also
given by Hart and Bailey (1979), Robaszynski and Caron (1979a, b), Hart and
Bigg (1981), Hart (1982) and Hart and Swiecicki (1987).
Although the major changes that occur at the Cenomanian-Turonian
boundary are now well documented, detailed foraminiferal ranges have only
been published for a single English Plenus Marl section, that at Shillingstone
Quarry in Dorset (Carter and Hart, 1977; Hart, 1985). The biostratigraphy
of a few comparable sequences in different facies, however, has been
described for sections in Dorset and Devon (Carter and Hart, 1977; Hart,
1985; Jarvis et al., 1987, 1988), and in Humberside (Hart and Bigg, 1981).
The foraminiferal biostratigraphy of the Plenus Marl at Cap Blanc-Nez in
the Boulonnais has been described by Robaszynski (in Robaszynski et al.,
1980; Robaszynski and Amedro, 1986), and the Cenomanian-Turonian
boundary successions in the SW Anglo-Paris Basin were considered by
Robaszynski et al. (1982). Extensive new data on foraminiferal assemblages
from Cenomanian-Turonian boundary sequences throughout onshore and
offshore southern England and in the North Sea Basin were presented by
Leary (1987).
Detailed analysis of these published foraminiferal records from
Cenomanian-Turonian boundary successions at other localities is un-
necessary since, within the limits of the resolution provided by previous
(generally considerably less detailed) studies, no major differences exist
between our records from Dover, and those presented for similar sequences
elsewhere in the Anglo-Paris Basin.
From the above discussion, it is clear that we can identify the top of the
Microfossil Assemblages and the Oceanic Anoxic Event 31

R. cushmani Total Range Zone (TRZ), as used throughout the world


between palaeolatitudes 40-45”N and S. The appearance of Helvetoglobo-
truncana helvetica may be diachronous, but this taxon is clearly of im-
portance in the identification of the early to mid-Turonian. Between the top
of the R. cushmani TRZ and the base of the H. helvetica TRZ, there is an
interval of time represented by a very improverished planktonic fauna
(Figures 6, 8, 9). Hedbergellids and early ‘whiteinellids’ (‘Hedbergekz spp.)
characterise this interval, and while the zonal index only appears in the upper
half of the unit (Figure 6), that is probably sufficient to allow the use of
‘Hedbergella’ archaeocretacea as a means of defining this Partial Range Zone
(PRZ). Throughout the succession the benthonic fauna is very variable,
changing from bed to bed, and is highly impoverished particularly immedi-
ately below the stage boundary. It is considered unjustifiable, therefore, to
set up a parallel zonation based on this fauna.

5. Ostracoda (D.H. & A.R.)


Thirty-two ostracod species were recorded (Table 2) from the succession, all
samples yielding some specimens (defined as being a complete carapace, a
single valve or a significant fragment of a valve) identifiable to specific level.
Stratigraphic studies indicate that, as with the foraminifera, major changes
occur in the abundance, specific composition and diversity of these as-
semblages, through the sequence.

5.1 Material and methods


The 34 samples studied were 1.5 kg splits taken in the field from 2 kg bulk
samples collected for foraminiferal, ostracod, calcareous nannofossil and
stable-isotope analysis (see Section 4.1 for details). Ostracod samples were
broken by hand into 1 cm-sized pieces. More indurated material was
fragmented inside thick polythene bags using a geological hammer. Ham-
mering was avoided wherever possible to minimise damage to specimens.
Fragments were dried at 60°C overnight, and disaggregated while hot
using 15 o/ow/v hydrogen peroxide. Marly chalks disintegrated immediately;
other lithologies were simmered gently on a hotplate for 30 minutes or longer
(up to 10 hours), depending on the speed of breakdown. Larger undis-
aggregated pieces (a high proportion of the sediment in some cemented
chalks) were separated using a 3 mm sieve, and discarded. The mud-sized
component was removed by washing through a 63 ,um sieve. Breakdown at
this stage was aided by gently rubbing the residue against the mesh with
naked finger-tips.

Figure 10. Cenomanian-Turonian foraminifera from Dover I. SEM photomicrographs of typical late
Cenomanian benthonic species. 411 of these taxa are common throughout the upper Cenomanian
except Ataxophragnium depressurn, which is a rarity that occurs in flood abundance in Plenus Marl
Bed 1. (a) Arenobuliminn advena (Cushman), ABC7. x 75; (b) Tritaxia macfadyeni Cushman, ABC7,
x 10: (c) Tritczxia pyramid&a Reuss, ABC6, x 115; (d) Dorothiagradata (Berthelin), ABCl, x 115;
(e) Marssonella trochus var. turris (d’orbigny). ABCl, x 150; (f) Eggerellina breuis Marie, ABCl.
x 115; (g) Plertina ma&e (Franke), ABCl, x 150; (h) Nodosaria sp. A, ABC4, x 55; (i) Dentalina sp.
A. ilBC8, x 55; (j) Lenticulina rotulata var. A, .4BC2 x 115; (k) Lenticulina rotulata var. B, ABC 1,
x 115; (1) Gawlinella baltica Brotzen, ABCl, x 75; (m) Gawlinella cenomanica (Brotzen), ABCl,
x 115; (n) Gavelinelln intermedia (Berthelin), ABC3. x 1 SO: (o) Atu.xophragmium depressurn (Perner).
ABc’7, x 115.
Microfossil .&emblages and the Oceanic anoxic Event 33

I;~gurc 12. Cenomanian-Turonian foraminifera from Dover I I I. SEA4photomtcrographs of planktonic


specws and early Turonian benthonic additions. (a) Hedbergella delrioensis (Carsey), ABCl, x 75:
(b) Rot&porn cushmani (Morrow), ABC7, x 1 SO; (c) Rotalzpora cushmani (Morrow), ABC7, x 115;
(d) Rotalipova greenhornensis (Morrow), ABC%, x 115; (e) Hele,etoglobotruncana helvetica (Bolli),
AKS‘C’. x 150: (f) Helvetoglobotruncana helewtica (Bolli), AKS’C’, x 1SO; (g) Hel~letoglobotrzlnrancr
helaetira (Bolli), AKS‘C’! x 1SO; (h) ‘Hedbergella’ archneocretacea (Pessagno), AKS7. x 115;
(i) Marginotruncana margrnafa (Reuss), AKSlO, x 150; (j) Marginotruncana marginata (Reuss).
AKSlO, x 150; (k) Tritaxza tricarinata var. jongmansi Schljfsma, AKS12, x 55; (1) Arenobulimirzu
preslii (Reuss), AKS7, x 150; (m) I,ingulogavelinella anmalerrsis (Sigal), AKS13, x 150.
‘I’ahle 2. Alphabetical list of ostracod species and then stratlgraphic distribution (Xc‘, Abbots Cliff
C’halk Formation [upper Capel-le-I‘wnr Llcmber]; PM, PI cnus Rlarl Formation; IX:, I)ovcr Chalk
Formation [Shakespeare Cliff Member] ).

SpWieS IIistributron

Bairdoppilata sp. A
H. pseudoseptentrionalis Mertens, 1956
B. southerhamerlsis Weaver, 1982
B~thoceratirra herrigi \Veaver, 1982
B. umbonata (Williamson, 1817)
B. umbonatoides (I&ye. 1964)
C’urfsina senior Porkom$. 1967
(‘ythereis sp. A sense Weaver, lY82
c ‘. sp. cf. C’. lo,~~ur~~~z I’orkwny. 1063
C~\‘t/rr/d/us,,. cf. (‘. clzclt/lrl?namls \veaver, 19x2
C. . U,I,C~~1‘ULVVezi\t‘r. I OX?
C‘.sp. cf. (‘. ~ofrtrrzl‘tn van Yeen. 1932
C’. wata Koemer. 1 X41
(‘~therrlloidea bonnrmai 1Veaver. 1982
r. hindei &ye, 1964
(‘. kayei Weavrr, 1982
C’. obliquirugata (Jones and Hinde, 1890)
C‘. stricta (Jones and Hinde, 1890)
Herrigocvtherr donzei (W”eaver, 1982)
Imhotepia eugl_vpheo Weaver. 1982
Isoc~thereis elongatn \Veaver, 1982
l,:~~~x~~who~ blrtrhel1crrsi.s\VC;I\er. 1OX?
Mosarleberis sp. X
Owtliella alata Weaver, 1982
0. donzei Weaver, 1982
Pawac~thrreis subpawa (Porkomq, 1967)
Phodrucythere cuneifortnis Weaver, 1982
Pontoc~prella robusta Weaver, 1982
Psrudob~thoc_vthere colini \Veaver, 1982
Ptervgocvthere sp. cf. P. dinrinuta Weaver, 1982
/‘. S,‘. ct. I’. lutlcr-r.stata (hsquet, 1854)
I’. sp. cf. I’. rohrrsto (Jones aml Hinde. 1 XYO)

Residues were dried at 6O”C, stored in glass phials, and picked for
ostracods using standard micropalaeontological techniques.

5.2. Biostratigvnph>f

English Cenomanian ostracods have been described in detail recently by


Weaver (1981, 1982), but there has been no comparable work on the
Turonian. Consequently, our work includes many new records of Turonian
taxa for southern England. Stratigraphic variation in the composition of
ostracod assemblages has been illustrated by Figures 13 and 14, which show
the distribution of ostracod species versus our lithostratigraphic logs.
Representative specimens are shown in Figures 15--l 9.

5.2.1 Abbots Clzff C‘halk Formation. ‘The Abbots Cliff Chalk (Figure 13)
yielded species typical of the English Upper Cenomanian (Weaver, 1981,
1982). ‘f’he lowest sample (ABCl) contained more than 200 specimens
comprising 16 species, the most diverse fauna in the succession (Figure 13).
Platycopids dominate the assemblage, being represented by four species of
Cytherella and tvvo species of Cytherelloidea. Cytherella ouata (Roemer)
[Figure 1 S(a)] constitutes > 50”,, of the total population. ‘l‘he dominant
podocopids, each comprising - 4’,, of the total fauna are Pontocyprella
klicrofossil Assemblages and the Oceanic Anoslc Event 35

rohusta Weaver [Figure 16(c)], Bairdoppilata pseudoseptentrionalis Mertens


[Figure IS(j)], Bythoceratina umbonatoides Kaye [Figure 16(g)] and Lox-
oconcha? bluebellensis Weaver [Figure 19(i), (j)]. Other samples contained
comparable assemblages to ABC1 (Figure 7) but yielded fewer specimens
(< 100).

5.2.2 Plenus Marl Formation. Ostracod assemblages from the basal Plenus
Marl (Bed la of Jefferies, 1962, 1963a, b) are similar to those in the
underlying Abbots Cliff Chalk. The lowest sample, however, yielded more
than 300 individuals, with C. oejata again dominating (- 709b) the fauna.
Isocytheresis elongata Weaver [Figure 19(d)] first appears at this level (Figure
13), comprising - 5 Y!bof the assemblage. This species is common through-
out the Plenus Marl at Dover but was not recovered from any other level,
although elsewhere in southern England it is known to be common through-
out the Middle and Upper Cenomanian (Weaver, 1982). Imhotepia euglyphae
Weaver [Figure 19(c)] and Pseudobythocythere colini Weaver [Figure 19(h)]
were only identified in the lower part of Bed 1, although these species are also
known to have more extensive ranges in the Cenomanian (Weaver, 1982).
Phodeucythere cuneiformis Weaver [Figure 16(d)] occurs throughout Bed I,
and is apparently entirely restricted to it (Weaver, 1982). A single adult
carapace recorded from the top of the Abbots Cliff Chalk (Figure 13) was
probably introduced from the basal Plenus Marl as part of a burrow fill. The
upper part of Bed 1 (lb of Jefferies, 1962, 19630, 6) yielded a similar
assemblage to the basal portion except for the absence of I. euglyphae and P.
colini and the occurrences of Cytherelloidea bonnemai Weaver [Figure 1 S(e)]
and Cythereis sp. A [Figure 18(a)-(d)], both of which are first recorded in the
underlying Abbots Cliff Chalk (Figure 13). A significant fauna1 change
occurs at the top of Bed 1, with the disappearances of four species,
C_vtherelloidea bonnemai, Oertliella alata Weaver, L.? bluebellensis and P.
cuneiformis.
Assemblages from Beds 2-3 contain fewer individuals (- 100) than those
from Bed 1 (Figure 13), but are of comparable diversity. Bythoceratina
umbonata (Williamson) [Figure 16(f)] is first recorded in Bed 3 and ranges
into the basal Dover Chalk, although it is known to occur throughout the
Cenomanian in southern England (Weaver, 1982). A major decline in
diversity begins in Bed 4 which yielded the last occurrences of P. robusta, B.
pseudoseptentrionalis and Herrigocythere donzei Weaver [Figure 19(a), (b)],
and the first appearances of Cytherelloidea hindei Kaye [Figure 1 S(f)] which
occurs sporadically throughout the remainder of the sampled interval, and
Pterygocythere sp. cf. P. robusta (Jones and Hinde) [Figure 16(j)]. The latter
species was recorded only from Beds 4-7 (Figure 13), although elsewhere it
occurs as a rarity lower in the Cenomanian (Weaver, 1982). A high
proportion of Cytherella specimens in Bed 4 and the higher parts of the
Plenus Marl (Beds 5-S) show presumed dissolution effects, with valves
displaying rough, etched surfaces, in contrast to their normal smooth
appearance.
Diversity continues to fall in Beds 6 and 7 (Figure 13), which yielded
assemblages containing only eight species. These assemblages are dominated
by platycopids, the only podocopids being Pterygocythere sp. cf. P. robusta,
I. elongata and Cythereis sp. A (the last species being represented only by
Cenomanian Turonian
ABBOTS CLIFF CHALK PLENUS MARL DOVER CHALK

Cape+le-Ferne Member Shakespeare Cliff Member

-
0-O -0-o-o-a B -0-o
*
. .

--
-a -_
0 o-o-
7 5
39 Turonian
2 u, E; DOVER CHALK 0”
E? .-
c”.g
3 c- Shakespeare Cliff Member .s
iVicrofossi1 .ksemblages and thv Oceanic ;\nosic Event 39

‘igurr 15. Cenomanian-Turonian ostracods from Dover I. SEM photomicrographs showing represen-
tative specimens of Cytherella spp., Cvtherelloidea spp. and Rairdoppilata pseudoseptentrionn1i.s; all
external lateral views. Th e f o 11owmg .. abbreviations are used in the description of Figures 15 IY:
>I = male; F = female; RV = right valve; I>\: = left valve; car. = carapace; ext. lat. =external lateral
view; int. lat. =internal lateral view; dors.=dorsal xieu. All specimens illustrated have hren
deposited in the British Museum (Natural History); catalogue numbers (OS 1234 etc.) refer to thl
collections of the BM(NH). (a) (‘y~herrllrz ovnt~r Rocmer, AHC‘I, F, RV, oSl312Y, > cr.<;
(1~) C’Jtherrlla sp. cf. c’. chathumrnsis Weaver. XBCI, ~u\enilr, K\‘, oSl 3132 x 0.i; (c) ~‘~thrll~i
corrrar’a Weaver, ABCI, F. R\‘, oSi310. x h5; (d) ~~rlrerrlla sp. cf. ( ‘. contracta \ an Vren, AHC’I
1;.RV, oS1 3 131, x 65; (e) Cythrrelloidea bot~nrmai IYra\ er. AIK‘l, F, I,\‘. oS13092, x X0; (f) CI\,tlwr
tlloidea hindei Kaye, ABC1 3, F, RI’, oS13 142 x 75;(g) (‘_~~thrrr//oidertr ohliqzrirugata (Jones 8r Iiinde).
.AKSll, F:, RV, oS13144, x 00; (h) ~‘~hwllmkn krrwi M’WVCT-, ARCl. F, R\‘, oS1 i 13-I.
x 90; (i) C’ytherelloidea strrcto (Jones & Hlnde) XH(‘Z. I:. 1.1.. 051 WW. xX0: (j) Raivdopp~lcrtcr
ptCl~dOS~~t(‘ntrlOIlNllS \\lel-tens. -\RC0. I,\‘, &13Olft j -!.5.
ure IO. Cenomanian-‘l’urr,nidn ostracods from Dover II. SEM photomicrographs shox+Ing re
sentatl\e spccimcns of Auirdoppilntrr spp., Porrtocyprello robustu, Phodeucytherr cuneijormis, ‘BZ’-
thocrratinu spp. and Pterygorythe “pp.; all extcmal lateral viexks. Abbreviations arc describe zd ‘in
Figure 15.(a) Bairdopilnta southeuhanwrsis Weavrr, ABCI, l.‘i:, oS13139. x 50; (b) Bairdopilat asp.
A, AKSI3, LV. oS13154, x 60; (c) Pmtoc~prella robusta Weaver. XBC3. I>\‘, 0513141, x 7.0;.
(d) Phodeuc_vtkerr cme~jo~mzs W’ea\w ABC7, RV, oS13 193, x 125; (c) Rythoceratina h<v-n@
Weavrr, ARCl, F, I,)‘, r&13089. x 75; (f) f?ythocwstitz~umbortataWilliamson, ABC19, LV.
oS13 117. x 35; (g) Bythorrrntinu umbonutoides &ye. ABCl, l,V, 0513088, x 75; (h) Pfel ‘yggo-
cythrre sp. cf. P. dintinuta Weaver, ABCl9. LV, us131 18. x 70; (i) Pterygocythere sp. cf. P. lati-
rristata (Bosquet), .4X9, RV, oS13 101, x 60; (j) Ptrrygrxythere sp. cf. P. rohustn(Jones and
Hinde), ABC13, I,V. oS131 17, x 60.
Microfossil Assemblages and the Oceanic Anoxic Event 41

Fig1 are 17. Cenomanian-Turonian ostracods from Dover III. SEM photomicrographs show ,ing
representative specimens of Mosaeleberis sp. A. Abbreviations are described in Figure 15. (a) AK S3,
M, LV, ext. lat., oS13124, x 75; (b) Detail of dorsomedian area of (a), x 200; (c) AKS2, F, LV, (ext.
lat.,oS13123, x75;(d)AKS12,F,RV, ext. lat.,oS13148, x 75;(e)AKS12, F, LV,ext. lat.,oS131 49,
x 75; (f) Detail of dorsomedian area of (e), x 200; (g) AKS13, F, LV, ext. lat., oS13126, x 75;
(h) AKS13, F, LV, int. lat., oS13127, x 75; (i) AKS13, F, car., dors., oS13152, x 75; (j) AKS13 . F.
RV, int. lat.. oS13150, x 75; (k) Detail of anterior hinge element of (jj. x 350.
42 1. ~;11.\.is rl a/.

Figure 18. Cenclmanian-‘l‘ur(,nlan ostracods from Dover IV. SE41 photomicrographa shwwng repre-
sentative specimens of C’vthereis spp. and C’urfsina senior.Abbreviations are described in Figure 15.
(a) Cythereissp. A., ARCS, F, R\‘. ext. lat., oS13119, x 65; (b) (‘ythereis sp. A., ARCI, M, RV, ext.
lat.. oSl3085, x 65; (c) C’yther-ris sp, A-\..AR~‘X,~u~enile, RY. est. lat.. oS13120, x 65; (d) C~derris
sp. A.. ABCl, 141. RV, int. lat., oS13086. x 65; (e) C’ythereis sp. cf. C’. longaeua Porkom$, AKSl, LV,
ext. lat., oS13146, x 60; (f) Detail of posteroventral part of(e), x 150; (g) Curjsina senior PorkornG,
AKS, ?F. LV, ext. lat., oS13147. x 75; (h) Detail ofposterodorsal areaof( x 350; (i) Curfsinasenior
Porkom~, AKSIO, ?X’l. car., right wt. lat.. oSl3128. x 75; (j) &tail of dorsomedian area of(i),
x 3.50.
Xlicrofossil Assemblages and the Oceanic Xnosic Event

F;IyuI.c IO. (‘cnomanian-‘l’urcjnian ostracods from Dover 1.. SEM photomicrographs showing rrl >rc-
sentative specimens species. all external lateral views except (j). Abbreviatiom
of several arr
described in Figure 15. (a) Hevrigoc\jthere ciorzzcG (\!‘eaver), ABC9, R\‘, 0S13107. x 90:
(b) Herrigocythere donzri (Weaver), ARCI, LV, oS13091, x 90; (c) Imhotepia euglyphea W’ra ver-.
ARC7, F, car., left side, oS13098, x 100; (d) Isocytherrrs elongate Weaver. ABC7, car., left s ide.
nSl3102, x95; (e) Parewcythereis suhpawa (Porkom+), AKSll, R\‘, oS13153, x 100; (f) Oe, vtli-
r//n domei Weaver, .4BCl. LV, oS13135, x 100; (g) Oertliefla akatcz Wezaver, ABC%, LV, oS131 121,
x 90; (h) Pseudobvthoc~lthera colini Weaver, .4BC7, F, RV. oS13100. x 150; (i) Lo.tocont ./lU?
hlrtebel/en.sis Weaver. ARCI, F. I,V. oS13138, x 105: (j) I,oxoconcha? bluebellmsis Weaver, ABC’1 , 1:.
KY, int. lat., oS13137, x 10.5,
in sample ABC1 5, and fragments of Bythoceratina hevrigi and B. umbonata
in sample ABC1 7.
Lowest Turonian sediments taken from immediately above the basal
hardground sequence (samples ABC1 9, 20), however, yielded more abund-
ant ( - 150 specimens) and diverse (6-9 species) assemblages. These are again
dominated by five platycopid species (Figure 13), with C. ozlata forming
> 700, of the total fauna. Podocopids are limited to four species, which
include the highest occurrences of B. herrigi (sample ABC19) and
B. umbonata (sample ABC20). Pterygocythere sp. cf. P. diminuta Weaver
[Figure 16(h)] was recorded only from this level. Sample ABC21 yielded
only a few Cytherella spp.
Higher material from the lower ‘l’uronian (samples AKSl-13) contained
assemblages which continue to be dominated by platycopids, particularly
ci. ozlata (Figures 13, 14). Diversity gradually increases through the
Shakespeare Cliff Member, however, as species which are important in
the higher Turonian at Dover (on the evidence of our own limited studies)
appear for the first time. First appearances in the Melbourn Rock are
Cytheveis sp. cf. C. Zongaeva Porkorny [sample ASKl; Figures 18(e), (f)]
and Curfsina senior Porkorny [Figures 18(g)-(j)] with Mosaeleberis sp. A
(sample AKS2; Figure 17). The new species Parvacythereis subparva
(Porkorny) [Figure 19(e)] and Cytherelloidea obliquirugata (Jones and Hinde)
[Figure 15(g)] first appear in the upper beds of the Shakespeare Cliff
Member, at the level of the Round Down Marl (Figure 14). Complete adult
valves of Bairdoppilata sp. A [Figure 16(b)], also common higher in the
Turonian, were only found in our top sample (AKS13), although juvenile
or fragmentary specimens from several lower horizons are tentatively
assigned to this species (Figures 13, 14).
The diversity of the ostracod assemblages at the top of the sampled
interval, however, is still only half that of assemblages recovered from the
upper Abbots Cliff Chalk and basal Plenus Marl. Preliminary studies of more
widely-spaced samples taken throughout the low and mid-Turonian at Akers
Steps, however, indicate that the trend to higher diversity is continued in the
overlying beds of the Dover Chalk.

5.3 Taxonomic notes

We intend to give full systematic treatment to our Cenomanian-Turonian


ostracods at a later date, when we have completed work on other sections as
well as Dover. However, it seems useful at this stage to provide short
taxonomic notes on a few species.

Cytherelloidea obliquirugata (Jones and Hinde, 1890) [Figure 1 S(g)]


This species was originally described on the basis of a juvenile specimen.
Neale (1978) noted the possibility that C. auricularis (Bosquet) was a senior
synonym; the form which Clarke (1982) illustrated as C. obliquirugata
appears to be quite distinct from our specimens, which are identical with the
form she assigned to C. auricularis; she did not, however, discuss any
relationship between the two.
Microfossil Assemblages and the Oceanic Anoxic Event 45

Pterygocythere sp. cf. P. diminuta Weaver, 1982 [Figure 16(h)]


Our single specimen differs in details of outline, particularly the posterior
margin, from Weaver’s (1982) illustrations of the type of P. diminuta; it also
closely resembles the form illustrated by Neale (1978) under the name of
Alatacythere robusta (Jones and Hinde) (see below).

Pterygocythere sp. cf. P. laticristata (Bosquet, 1854) [Figure 16(i)]


Our specimens correspond to those illustrated by Weaver (1982) and
are more elongate that those figured by Neale (1978) under the name of
P. laticristata.

Pterygocythere sp. cf. P. robusta (Jones and Hinde, 1890) [Figure 16(j)]
Our specimens, like those of Weaver (1982), p assess a short longitudinal rib
just below the posterior half of the dorsal margin; in this respect they differ
from those illustrated by Neale (1978) ( as Alatacythere robusta) which
resemble P. diminuta Weavrer.

~Wosaeleberis sp. A [Figures 17(a)-(k)]


Specimens from the lower part of the range (e.g. AKS2, 3) of this species
possess distinctive primary and secondary intercostal reticulation [e.g.
Figures 17(a)-(c)], while higher up (e.g. AKS13) we found forms, identical
in every other respect, in which the intercostal areas are only faintly reticulate
or smooth [Figures 17(d)-(g)]. ‘l‘h e reticulate forms seem to be identical with
iVIosaeZeberis interruptoidea (Van Veen) serzsu Porkorny. Porkorny’s material
was Turonian, but Van Veen’s was Maastrichtian; Porkorny stated that he
examined a Maastrichtian specimen and could see little difference. However,
Clarke (1983) illustrated a Maastrichtian specimen from Van Veen’s collec-
tion, cited as Imhotepia interruptoidea (Van Veen), which differs in several
respects from the Turonian form: it has more prominent dorsal (continuous),
median and ventral ribs, and possesses an anterior marginal ridge, and there
appear to be differences in shape and reticulation. Porkorny’s Turonian form
is therefore probably a new species, but studies of type material are required
to sort out this problem once and for all. -4 further complication is that the
smooth forms from the British Turonian seem very similar to Kasteneis
l Prostezzeis) nodifera (Kafla) as illustrated by Porkorny (1963a)-and this
form seems to be closely related to Mosaeleberis macrophtha~ma (Bosquet),
which occurs in the Santonian and Campanian of Britain (Neale, 1978).

Cythereis sp. cf. Zongaeva Porkorny, 1963 [Figures 18(e)-(f)]


We have found this distinctively reticulate species in several samples from
higher up the Turonian succession at Dover. It is very similar to two
subspecies described by Porkorny (1963b), C. longaeva longaeva and
C. longaeva lysicensis, except that it lacks spines. Three further subspecies
of this form, described by Clarke (1983), are also similar: C. longaeva prior
(Upper Coniacian) is more elongate than our specimens, while in C. Zongaeva
triaculeata (Santonian-Campanian) the median and dorsal ridges are linked
posteriorly. Comparisons must also be made with Cytheresis zygopleura
Porkorny (Turonian-Coniacian), Cythereis luzicensis Porkorny (Coniacian),
46 I. JX\ li f’/ t/l.
~._.._ .-

and a Campanian species from France, Mauritsina agedincumemis


(Damotte), as illustrated by Babinot et al. (1985). Further studies are
needed to determine to which, if any, of these species and subspecies our
form belongs.

5.4 Discussion
Assemblages throughout the section are dominated by platycopids, parti-
cularly Cytherella ozrata. One species, C_vtherelloidea bonnemai disappears at
the top of Bed 1 of the Plenus Marl, to be replaced by C. hindei which first
appears in Bed 4 (Figure 13). Podocopids, on the other hand, show a dramatic
decrease in diversity through Beds 1 -8, particularly at the tops of Beds 1 and
4. lsocythereis elongata is the only podocopid to occur in every Plenus Marl
sample and only three podocopids appear to have struggled through into the
overlying Dover Chalk. One of these, Cythereis sp. L4 is represented only by a
single fragment at the base of the Melbourn Rock, while the others, both
species of BJtthoceratina, disappear in the basal few decimetres of the
‘I’uronian.
The refilling of the niches left vacant by these disappearances was ver)
gradual. l‘he lower numbers of ostracods obtained from samples above the
Plenus Marl can undoubtedly be attributed, at least partly, to difficulties in
processing more indurated chalks. ‘l‘he evidence from the easily disag-
gregated samples from the Plenus Marl suggests, however, that they do, to
some extent, reflect a real reduction in abundance as well as diversity.
Work on Cenomanian ostracods in southern England by Weaver (1982)
included bed-by-bed sampling of the Plenus Marl at Shillingstone (Dorset)
and Pitstone (Hertfordshire), which demonstrated that major changes in the
ostracod fauna occur through these sequences. The changes recorded by
Weaver (1982) are generally similar to those demonstrated by us at Dover,
but they differ in detail. At Dover the major disappearances occur at the
tops of Beds 1 and 4, but at Pitstone the tops of Beds 1 and 2 are the sig-
nificant levels. At Shillingstone only the top of Bed 2 shows a major effect,
although several species disappear at the top of the underlying ‘Grey Chalk’
(= Abbots Cliff Chalk).
The sequence of changes which occur in the ostracod assemblages appears
to be virtually the same at all localities. Losoconcha? bluebeltensis and
Imhotepia eugl_vphae (the latter \vas not recorded at Shillingstone; Weaver,
1982) are, for instance, the first species to disappear. HervigocJltheve donzei
and Pontocypvella robusta go next, at the top of Bed 2 at Shillmgstone and
Pitstone, but at the top of Bed 4 at Dover. Isocytheveis elorzgata survives as far
as Bed 8 at all localities. Cythevelloidea hindei first appears in Bed 4 at Dover,
in Bed 3 at Pitstone, but at Shillingstone is occurs throughout the Plenus
Marl as well as in the underlying ‘Grey Chalk’, and at Southerham (Sussex)
Weaver (1982) recorded a few specimens in the Middle Cenomanian. It
would be premature to speculate on the significance of these geographical
differences in the lithostratigraphic distribution of the different species. It
does seem, however, that appearances and disappearances are stratigraphi-
tally lower to the west and south of Dover. ‘l’he extension of our current
investigations to other localities, both in England and on the Continent,
should be informative in this respect.
Microfossil Assemblages and the Oceanic Anosic Event 47

Weaver (1982) also examined a few samples from above the Plenus Marl,
noting difficulty in processing the harder chalks, but those species which he
did record are essentially the same as those found at Dover in the Melbourn
Rock: Bythoceratina herrigi, B. umbonata, Cytherelloidea hindei and numer-
ous Cytherella spp.
Detailed ostracod records are not available for other areas of southern
England or northern France. Studies of ostracods in the Cenomanian-
Turonian boundary sequences of the Saumur (Maine-et-Loire) and Givray-
de-Touraine (Indre-et-Loire) areas (Damotte in Robaszynski et al., 1982),
SW Anglo-Paris Basin, have revealed only low diversity faunas. These have
few species in common with those in southern England and a direct
comparison is not possible.

6. Calcareous nannofossils (M.K.E.C.)

Seventy-three species of calcareous nannofossil were identified from the


Cenomanian-Turonian boundary sequence at Dover (Table 3). All samples
yielded adequate material for analysis, though abundances were lowest in the
basal beds of the Melbourn Rock. Most species range throughout the interval
studied but, nevertheless, significant changes in specific composition and
diversity were recorded.

6.1 Material and methods

Samples were 5 g splits taken in the field from 2 kg samples collected for
microfossil and stable-isotope analysis. Thirty-two samples were analysed,
sample numbers AKS9 and 13 not being taken for calcareous nannofossil or
stable-isotope analysis.
Calcareous nannofossils were extracted after first soaking the samples in
Calgon overnight. Disaggregated fines were suspended by shaking, an
aliquot being pipetted-off and used to produce a smear slide. The remaining
suspension was poured-off and centrifuged until a clear solution was
obtained. A second concentrated smear-slide was produced from the cen-
trifuged residue. Nannofossils were identified using light microscopy on
both the original and preconcentrated smear slides.

6.2 Biostratigraph)

The biostratigraphy of Upper Cretaceous calcareous nannofossils in south-


ern England has been described recently by Crux (1982), who proposed a
zonal scheme for the area. Our study, however, is the first detailed analysis of
the Cenomanian-Turonian boundary sequence. The stratigraphic range of
each species was recorded (Table 3) along with the specific diversity in each
sample (see Section 11 .l), stratigraphic range charts being produced for the
most important species (Figures 20, 21). N annofossil species selected for the
range charts are those discussed in detail in the text, plus those species which
are typical of the stratigraphic interval studied.

6.2.1 Abbots Cliff Chalk Formation. The Abbots Cliff Chalk yielded 58
nannofossil species (Table 3), with diversities ranging from 34-45 species in
individual samples. The ranges of 21 species are shown in Figure 20. The
‘l’ahlc 3. Alphabetical list of nannofossil species and their stratigraphIc distribution (.4(‘, .4bhots (‘lift
Chalk [upper Capel-lr-Frrnr 1lcmbrr]; PM, PI enus Marl Formation; DC, Dover Chalk Formation
[Shakespeare (‘liff LIemhrr]).

Species Distribution

Ahmucllerella octoradiata (Gorka, 1957) Reinhardt, 1964 Lx


Axopodorhabdus albianus (Black, 1967) Wise and Wind, 1973 AC, PM
L-2.dietomannri (Rheinhardt. 1965) \Vmd and \Vise, lY,X3 A(‘, PA1
Biscutum blackii Gartner, 1968 AC. PM
B. dubium (Noel, 1965) Griin, Prins and Zweili, 1974 AC, Pxl
B. ellipticum (Gbrka, 1957) Griin and Allemann. 1974 AC, PM, DC
Braarudosphaera afvicana Stradner, 1961 AC, PM, DC
B. bigelowii (Grun and Braarud, 1935) Deflandre. 1947 DC
Brionsnnia enormis (Shumenito. 196X) .Manivlt, 1951 .1C, PM, DC
B. signata (Noel, 1969) Noel. 1970 AC, I’M, DC
Chiastozygus litteravius (G6rka. 1957) Manivit, 1971 AC, PM, DC
C. platyrhethum Hill, 1976 AC, PM. DC
Corollithion achylosum (Stovrr, 1966) Thierstein, 1971 AC. PM, DC
C. exiguum Stradner, 1961 DC
c‘. kennedyi Crux, 198 1 AC
C’retarhabdus conicus Bramlette and Martini, 1964 XC, PM. DC
c’. striatus (Stradner, 1963) Black, 1973 AC’, PLI
Cribrosphaerella primitiva Thierstein 1974 DC
C’yclagelosphaera sp. cf. C. margevelii Noel, 1965 PM
CJlindrafithus coronatus Bukry, 1969 DC
C. lafittei (No&l, 1957) Black, 1973 AC, PM, DC
Discorhabdus ignotus (Gbrka, 1957) Perch-Nielsen, 1968 AC, PM. DC
Dodekapodorhabdus noelae Perch-Nielsen, 1968 AC, PM
Eiffellithus sp. cf. E. eximius (Stover, 1966) Perch-Nielsen, 1968 PM, DC
E. turriseiffelii (Deflandre. 1954) Reinhardt, 1965 AC, PM, DC
Ellipsagelosphaera keftalremptii Grtin and Allemann, 1975 AC, PM, DC
Flabellites oblonga (Bukry, 1969) Crux, 1982 AC, PM, DC
Gartnerago obliquum (Stradner, 1963) Noel, 1970 AC, PM, DC
Gvantarhabdus covonadwntis (Reinhardt, 1966a) Grun and Allemann, 1975 AC, PM, DC
Haqius circumradiatus (Stover, 1966) Roth, 1978 XC, PM, DC
Helicolithus trabeculatus (G6rka. 1957) Verbeek, 1977b AC. PM, DC
H. sp. .4 XC. PM, DC
Lithastrinusflotalis Stradner. 1962 ,4C:, PM. DC
L. moratus Stover. 1966 DC
1,. sp. A DC
Lithraphidites acutum Verbeek and Manivit, 1977 AC. Pn,I
L. ala&s Roth and Thierstein, 1972 AC
L. carniolensis Deflandre, 1963 AC. PM. DC
Mantzitella pemmatoidae (Manivit, 1965) Thierstein, 1971 AC, PM, DC
Microrhabdulus decovatus Deflandre 1959 4C, PM, DC
Microstaurus chiastius (Worsley, 1971) Grlin and Allemann, 1975 .4C, PM, DC
Nannoconus dauelillieri Deflandre and Deflandre, 1959 DC
N. elongatus Bronnimann, 1955 DC
N. multicadus Deflandre, 1959 DC
N. regularis Deres and Acheriteguy. 1980 DC
N. truth Briinniman, 1955 ilc, PM, rx
N. indet PM, DC
Parhabdolithus embergeri (N&l, 1958) Stradner, 1963 AC, PM, DC
Pwdiscosphaera columnata (Stover, 1966) Perch-Nielsen, 1984a .\C. PM, DC
P. cretacea (Arkhangelsky, 1912) Gartner, 1968 AC, PM, DC
P. ponticula (Bukry, 1969) Perch-Neilsen, 1984a .4C, PM, DC
P. spinosa (Bramlette and Martini, 1964) Gartner, 1968 AC, PM, DC
Quadrum gartneri Prim and Perch-Nielsen, 1977 DC
Retecapsa angustijorata Black, 1971 AC, PM, DC
R. crenuiata (Bramlette and Martini, 1964) Grun and Allemann, 1975 AC. PM, DC
Rhagodiscus achlyostaurion Hill, 1976; emend Crux, 1981 AC, PM, DC
R. angustus (Stradner, 1963) Reinhardt, 1971 AC, PM, DC
R. asper (Stradne;, 1963) Reinhardt, 1967 AC, PM
Srompanella indet. AC. PM, DC
Microfossil Assemblages and the Oceanic Anoxic Event 49

Scapholithusfossilis Deflandre and Fert, 1954 AC


Sollasites horticus (Stradner, Admaiker and Maresch, 1966) Black, 196X AC, PXl, DC
Tetrapodorhabdus sp. cf. T. derorus (Deflandre in Deflandre and Fert. 1954)
Wind and Wise in Wise and Wind. 1977 AC, PXI
Tranolithus gab&s Stover, 1966 AC, PM. DC
T. orionatus (Reinhardt, 19660) Perch-Neilsen, 196X AC, PM, DC
’ I’erkshinella’ dibrachiata Gartner 1968 AC, PM
Watznaueriabanzesae (Black, 1959) Perch-Nielsen. 1968 AC, PM, DC
U’. biporta Bukry, 1969 AC, PM, DC
Zygodiscus diplogrammus (Deflandre in Deflandre and Fert, 1954) Gartner.
I Y)hX
Z. elegans Gartnrr, 1968
Z. erectus (Deflandre and Fert 1054) Lezaud. 1968
Z. sp. cf. Z. minimus Bukry, 1969
Z. minimus Bukry, 1969
Z. pseudonnthophorus Bramlett and Martini, 1961

appearances of four species, Cretarhabdus striatus (Stradner) Black, Lithra-


phidites acutum Verbeek and Manivit, Nannoconnus truittii Bronnimann,
Flabellites oblonga (Bukry) Crux, which occur above the base of the section,
are not considered significant, since all are known to range lower in the
Cenomanian elsewhere (Crux, 1982; Perch-Neilsen, 1985). Species recorded
only from the Abbots Cliff Chalk (Table 3) are: Corollithion kennedyi Crux
and Lithraphidites alatus Roth and Thierstein, which are known to disappear
within the Cenomanian (Crux, 1982), and Scapholithusfossilis Deflandre and
Fert, which has a more extensive stratigraphic range than that recorded here.
The appearance of Helicolithus sp. A at the very top of the Abbots Cliff
Chalk (and, therefore, possibly in a burrow-fill originating from overlying
Plenus Marl) may be more important, since this species has only been
described recently (Mortimer, 1987), and its range is not yet fully
documented.

6.2.2 Plenus Marl Formation. Fifty-eight species of nannofossil were re-


corded from the Plenus Marl (Table 3), 22 of which are plotted in Figure 20.
Assemblages and specific diversities in the lower beds (l-4) are similar
(36-44 species) to those in the Abbots Cliff Chalk, but above that, nan-
nofossil diversity falls rapidly, declining to only 24 species in Bed 8.
Two nannofossil species, Cyclagelosphaera sp. cf. C. margerelii Noel,
Eiffeellithus sp. cf. E. eximius (Stover) Perch-Nielsen appear in the Plenus
Marl, while indeterminate Nannoconus spp. are also seen for the first time.
Eiffellithus sp. cf. E. eximius is similar to E. eximius in having a central cross
close to the axes of the ellipse of the coccolith. In the specimens found here,
however, there is a slight twist of the cross arms away from the ellipse axes
(Figure 20), where they enter the rim. Eiffeellithus sp. cf. E. eximius appears
just above the base of the Plenus Marl (Bed lb) and ranges throughout the
remainder of the sampled interval.
Several species were not recorded above the Plenus Marl, Axopodorhabdus
albianus (Black) Wise and Wind, A. dietzmannii (Reinhardt) Wind and Wise,
Riscutum blackii Gartner, B. dubium (Noel) Grun, Prins and Zweili,
Cretarhabdus striatus (Stradner) Black, Cyclagelosphaera sp. cf. C. margerelii
Noel, Dodekapodorhabdus noelae Perch-Nielsen, Lithraphidites acutum
Verbeek and Manivit, Rhagodiscus asper (Stradner) Reinhardt, Tetra-
Cenomanian Turonian
ABBOTS CLIFF CHALK DOVER CHALK

Cape+le-Ferne Member Shakespeare Cliff Member

I MRB

*-•- o-0-0

o-o-o-o-

o-o-o-.-

o-o-
& ..j:,
-
..::.
0 - w+<
;i
o-o- l -o-o--_~*,~

l-•-•-•-.-•_+
Microfossil Assemblages and the ( keanic .Anos~c Event 51

i i i

I I I
I 7 i
l 0 i

I I I I
l !, l 0
52 1. JanIs rt LZl

podorhabdus sp. cf. T. decorus (Deflandre in Deflandre and Fert) Wind and
Wise in Wise and Wind, ” Vekshinella” dibrachiata Gartnkr and Zygodiscus
elegans Gartner. Of these, only the disappearances of A. albianus, C. striatus,
L. acutum, and R. asper were believed to represent extinction levels, which all
occur towards the top (Beds 7 and 8) of the formation (Figure 20). Other
species are known to have more extensive stratigraphic ranges elsewhere.

6.2.3 Dover Chalk Formation. Sixty calcareous nannofossil species were


recorded from the Dover Chalk (Table 3), 26 of which are plotted in Figures
20, 21. Specific diversity is lowest near the base of the formation where a
minimum of 10 species is recorded (sample ABC 18), but recovers to 30-40
species throughout most of the Shakespeare Cliff Member. A number of
stratigraphically important appearances and disappearances occur in this
part of the sequence.
Several nannofossil species appear immediately above the base of the
Turonian, within the lower part of the Melbourn Rock (Figure 20).
Particularly important is the first record of Quadrum gartneri Prins and
Perch-Nielsen (a basal Turonian index species; Birkelund et al., 1984).
Lithastrinus sp. A, L. moratus Stover and Nannoconus multicadus (Deflandre)
also appear around this level. Lithastrinus sp. A is an unnamed variant which
has eight rays, and from our data appears to be restricted to the basal
Turonian (Figures 20, 21). Other species of Nannoconus (N. elongatus
Bronnimann, N. regularis Deres and Acheriteguy) are also represented for
the first time in this part of the section. Only two species appear higher in the
succession, both just above the top of the Melbourn Rock (Figure 21). These
are Cylindralithus coronatus Bukry (a low Turonian index species; Crux,
1982) and Ahmuellerella octoradiata (Gorka) Reinhardt. Less informative
species which were only recorded from the Dover Chalk (Table 3) are,
Braarudosphaera bigelowii (Grim and Braarud) Deflandre, Corollithion
exiguum Stradner, Cribrosphaerella primitiva Thierstein, Nannoconus
dauvillieri Deflandre and Deflandre, Zygodiscus sp. cf. Z. minimus
Burkry and Z. minimus Bukry.
Several disappearances occur in the Dover Chalk. The biostratigraphi-
tally important species Microstaurus chiastus (Worsley) Grun disappears
within the basal part of the Melbourn Rock (Figure 20) at the Cenomanian-
Turonian boundary. Corollithion achylosum (Stover) Thierstein, on the other
hand, disappears just above the top of the Melbourn Rock (Figure 21). This
species was regarded previously as becoming extinct within the Cenomanian
(e.g. Perch-Nielsen, 1985), our records necessitating an extension of its range
into the basal Turonian. Microrhabdulus decoratus Deflandre also disappears
in the basal part of the Melbourn Rock, just above the base of the Turonian.
This species, however, is known to range throughout the Late Cretaceous
(Perch-Nielsen, 1979, 1985). Disappearances of possible significance higher
in the succession are Helicolithus sp. A and Nannoconus regularis which range
up to around the top of the Melbourn Rock (although both species are also
known to have more extensive stratigraphic ranges elsewhere), and Lith-
astrinus sp. A which also disappears at this level, and is probably restricted to
the earliest Turonian.
Microfossil Assemblages and the Oceanic Anoxic Event 53

6.3 Discussion

Our records allow the delimitation of Crux’s (1982) calcareous nannofossil


zones in the sequence. The summit of the Abbots Cliff Chalk, the Plenus
Marl and the basal beds of the Dover Chalk (samples ABCl-20) can be
assigned to the Microrhabdulus decoratus Interval Subzone, the uppermost
subzone of the Cenomanian Eiffeellithus turriseiffeelii Interval Zone. This
subzone is defined by the first occurrence of M. decoratus up to the first
occurrence of Quadrum gartneri. Microrhabdulus decoratus is found through-
out the Abbots Cliff samples (ABCl-21) and in the lowest sample from Akers
Steps (AKSl; Figure 20). Quadrum gartneri, however, appears immediately
above the base of the Turonian (sample ABC21), as defined by the
macrofaunal data.
The remainder of the section studied (i.e. the bulk of the Shakespeare Cliff
Member) is assigned to the Cylindralithus coronatus Partial Range Subzone,
the basal subdivision of the early Turonian Q. gartneri Interval Zone (Crux,
1982). The base of the C. coronatus Subzone is defined (Crux, 1982) by the
first appearance of Q. gartneri, its top being identified by the first occurrence
of Lucianorhabdus quadrijidus Forcheimer, a species which was not found
during the present study. Crux (1982) commented that C. coronatus is
common in its subzone, although we were only able to identify the species in
three samples (Figure 21; AKS8, 10, 11) from the Shakespeare Cliff
Member.
It is noteworthy that the appearance of Q. gartnerz’ and the base of the
Turonian defined by the appearance of mytiloid inoceramids coincide so
closely at Dover, both occurring within less than 1 m of each other in the
succession. These data support the use of Q. gartneri as an important index
for defining the base of the Turonian (cf. Birkelund et al., 1984).
5c pecimens of the genus Nannoconus are very rare throughout the
Cenomanian part of the sequence. A major change occurs in the middle of the
Melbourn Rock (samples AKS33.5), however, where the abundance and
diversity of Nannoconus undergo marked increases, and the genus becomes
an important part of the nannofossil assemblage. It has been suggested
(Roth, 1981; Roth and Bowdler, 1981) that Nannoconus was more abundant
in higher energy ‘shallow-water’ environments (continental margins, shal-
low plateaux, epicontinental seas). Work in progress (M. K. E. Cooper,
unpublished data) on Lower Cretaceous Nannoconus species in northern
Tunisia supports this idea. The influx of Nannoconus within the Melbourn
Rock, therefore, conforms with the higher energy facies represented by these
beds.
The nannoplankton stratigraphy of the Chalk at Cap Blanc-Nez (Pas-de-
Calais) near Calais, northern France was studied by Manivit (in Robaszynski
et al., 1980), who analysed 12.5 samples from the Albian-Santonian. The
number and location of samples taken across the Cenomanian-Turonian
boundary is, however, unclear.
Manivit (in Robaszynski et al., 1980) recognised the Microrhabdulus
decoratus ‘Zone’ in the Cap Blanc-Nez sequence, noting the range of the
index species as Middle Cenomanian to Santonian. No temporary disap-
pearance of M. decoratus in the low Turonian (cf. Dover) was noted.
Turonian
DOVER CHALK

Shakespeare Cliff Member

MRB I
1’Iicrofossil A4ssemblages and the Oceanic Anoxic Event 55

I
0 0

_ - _ .

I .
i I I
Quadrum gartneri appears (marking the base of the Q. gartneri Zone) within
the low Turonian at Cap Blanc-Nez (Manivit, op. cit.), some distance abovre
the base of the stage, and possibly, therefore, stratigraphically higher than at
Dover. The disappearances of Axopodorhabdus albianzts close to the
Cenomanian-Turonian boundary, and of Prediscosphaera columnata in the
lower Turonian, occur at similar levels at both localities. Lithraphidites
acutum, which disappears in the Upper Cenomanian at Dover, is recorded as
ranging into the lower Turonian at Cap Blanc-Nez (Manivit, op. cit.).
The first appearance of Nannoconus multicadus around the level of the first
Q. gartneri has been noted at both Dover and Cap Blanc-1Vez, although
N. truittii is recorded higher at the former locality. It can be concluded,
therefore, that there is good general agreement between nannofossil
records from Dover and Cap Blanc-Nez. Minor discrepancies are probably
due in part to the (presumably) larger number of samples analysed from the
Cenomanian-Turonian boundary sequence at Dover.
Comparative calcareous nannofossil data have also been published for the
Turonian stratotype region of Touraine in the SW Anglo-Paris Basin.
Working in the area around Saumur (Maine-et-Loire), Manivit and Zeigh-
ampour (in Robaszynski et al., 1982) recognised both the M. decoratus ‘Zone’
and the Q. gartneri Zone in the succession. Similar data were presented
(Manivit in Robaszynski et al., 1982) for the Civray-de-Touraine (Indre-et-
Loire) borehole.
Several species are common to both the Dover and Touraine sequences,
but there are some differences in detail. Quadrum gartneri appears some
metres above the base of the Turonian in Touraine (Manivit, op. cit.).
Axopodorhabdus albianus occurs only in the low to mid-Cenomaman ot
Touraine, and does not extend up into the M. decoratus ‘Zone’, while
Ahmuellerella octoradiata appears earlier than at Dover, within the low
Cenomanian. Prediscosphaera columnata disappears in the lowest Turonian
in western France (lower than at Dover), while Lithraphidites acutzdm ranges
higher, disappearing only at the summit of the lower Turonian. Manivit
(op. cit.) noted that species of Nannoconus are generally rarer in the
marginal calcarenitic and detritus-rich sediments of Touraine than they are
in coeval chalks, although at Dover we have noted that Nannororzus first
becomes abundant in the calcarenitic chalks of the Melbourn Rock.
The variations in the ranges of nannofossil species between Dover and
Touraine is probably the result of provincialism, but more data are required
from sections elsewhere in the Anglo-Paris Basin to identify the palaeo-
geographical factors which might control the stratigraphic distribution of
Cenomanian-Turonian calcareous nannofossils.

7. Dinoflagellate cysts (B.A.T.)

Dinoflagellate cyst assemblages across the Cenomanian-Turonian boundary


yielded 37 species and subspecies (Table 4). Most samples contained some
identifiable cysts, although uppermost Cenomanian and lowest Turonian
sediments proved to be barren. Substantial differences in the abundance,
specific composition and diversity of the dinoflagellate cyst assemblages
occur through the succession.
Microfossil Assemblages and the Oceanic Anosic Event 57

Table 4. Alphabetical list of dinotlagellate cyst species and their stratigraphic distribution (AC, Abbots
Cliff Chalk [upper Capel-le-Ferne Member]; PM, Plenus XIarl Formation; DC, Dover Chalk
Formation [Shakespeare Cliff Member]).

Species Distribution

.dldo$ia d@andrei (Clarke and Verdirr, 1967) Stover and l<\,ltt, 197X xc
~n~~~rcasplzaern euteiches (Davey. 1969) Davey. 1979 AC, P>l, DC
C’anrtingia collic~eri Cookson and Eisenack, 1960 AC, PM, DC
Cleistosphneridiun~ ? aciculare Davry, 196Y AC, PM
C. mmntum(Deflandre, 1937) Davry. 1969 AC
C. ~lnvuluw~ (Davey, 1969) Below, 19X2 AC. PM. IX’
(‘rihroperidinium ewilicristatum (Davey. lY69) Stover and I:vitt. lY78 ilc
fJ_~c/onephelium distincturn Deflandre and Cookson. 1955 .4C. PM. DC
(‘. mmbroniphortm Cookson and Eisenack, 1962 AC. PM. DC’
E.w&osphaeridium bi’dztnz (Clarke and Verdier, 1967) Clarke et ni. lY68 A<‘
E. pbragmites Davey et al.. 1966 PM
E‘romae granulosa (Cookson and Elsenack, 1974) Stover and Evitt, 1978 AC
Hete~osphoeridilrm ? heterrrranthum (Defiandre and Cookson, 1055)Eisrnack
and Kjellstrom, 1971 AC. PM, DC
H~ctrirllosphaeridizrm dijjicile Manum and Cookson, 1964 IX
H. palmatum (White. 1842 ex Bronn. 1848) D ownie and Sarjeant, 1965 AC. PM, DC
Kallnsphnrridium sp. B Tochcr and Jarvis, 19X7 DC
K. rlrrgnesiorum (.22anum and Cookson, iY64) Tocher and Jarvis, lYX7 1’11, DC
Kiokansium pol~$as (Cookson and Eisenack. 1962) B&W. 1YX2 AC‘, Phi
Lebwidocysta chlnm_ydattr (Cookson and Eisenack, lY62) Stover and Evitt,
197X AC
1, d&ccata (Davey and Verdier, 1973) Stover and Evitt, I Y78 .4C. DC
Odrmtochitina mstata Albcrti, 196 1; emend Clarke and Verdier. 1967 AC, I’M, DC’
0. nprrculata (0. Wetzel. 1933) Dellandre and Cookson, lY55 XC‘, PXI, DC‘
O/iec)sphaeriditcNI albertense (Pocock, 1962) Davey and \Villiams. 1 Y60 XC
0. rwnples (N’hite. 1812) Dave! and Williams. 19h6 A<‘, I’>l. 1x
0. p~~c~ulun~ Jain, 1977 AC‘, DC
0. prolixispinosum Davcy and $Villiams, 1966 AC’, PM. DC
Prolr.~osphaeridiiunr conulun~ Davry, lY6Y .A\(‘, PZI. DC
Ptrrodinium cingularmr vrtiruiutzcnt (Davch- and Williams. I ‘MI Lentin and
\Vllliams, 1981 AC
Rh>whodiniopsiv sp, cf. R. aptiona Drfandre, 1935 AC
SanoGzsphaera mirroreticrtlata Rrideaus and McIntyre. I’)75 DC
,S. wtundata Clarke and Verdier. 1967 rx-
SI,,,tllsldirrirrrr,~1’. .A ‘I’ocher and Jat-1-x. IW7 1’.\1.DC:
S sp. 13Tocher and Jarvis, 1987 Lx:
S. sp. C Tocher and Jarvis, 1987 P1I. DC
Slc~rrllosphueridium? longifrtrcatum (Flrtion, 1952) Dave? et (II.. 1966 .4C
?;rn?,osphaeridiunr variecalanrus Dave? and Williams, 1966 P1\1
,Gwccwr.s cerntioides (Dellandre. lY37) Lentin and Williams, lY73 XC

7.1 Material and methods

Thirty-nine 20 g samples were collected for dinoflagellate cyst analysis. This


material was obtained from the Abbots Cliff (samples ABC/Dl-21) and
Akers Steps (samples AKS/Dl-18) sections, but was collected independ-
antly of that used for the other parts of the study (Section 4.1) Sample
positions (Figures 22, 23) do not all correspond, therefore, with those used
for other microfossil analyses, and consequently a separate sample-
numbering scheme, (using a D suffix; see above) has been employed.
Samples were prepared by BP Research, Sunbury, using standard palyn-
oi::Gcal methods
Cenomanian Turonian 0 0 l 0
ABBOTS CLIFF CHALK PLENUS MARL DOVER CHALK
N : “: 2
I II, II ; I 2
Capel-le-Ferne Member - ww PYJjVo) Shakespeare Cliff Member s o b
I III I, 0”

r
3
0
0
2

.-.-
R;licrofossil Assemblages and the Oceanic Xnosi,, Event

.. 0
I
0

Note cysts are not drawn lo a constant scale


60 I. JX\lS rt ni.

Number of specimens

0 single 0 51-100

. 2-10 l 101-200

0 11-20 l 201-500

8 !I-50 501-1000
0

. . .

RDM

.
. 0 .
. .
I .
.
I
.
I .
.
I I .
I I
0 1 . ;. 0
. .

.
I
I .I .
1

I
I
.

Lithology
Figure 23. Dinofagellatr cyst distribution and lithostratigraphy in the higher beds of the lower
‘ruronian at Dover. The succession is a continuation of that shown in Figure 22. Lithologies and
symbols are explained in Figures 5. 6 and 13. The horizontal dashed line indicates the top of the
Melbourn Rock.
Microfossil Assemblages and the Oceanic Anoxic Event 61

0
.
.

Note: cysts are not drawn to a constant scale


62

7.2 Biostratigraphy

Previous palynological studies of the Upper Cenomanian and lower


Turonian of southern England include those of Clarke and Verdier (1967) for
the Isle of Wight, Davey (1969, 1970) for Sussex, Surrey and Hampshire,
and Tocher and Jarvis (1987), J arvis and Tocher (1983), and Jarvis et al.
(1987,1988) for SE Devon. No previous work has been published on the east
Kent succession. The distributions of dinofagellate cysts and their abun-
dances at Dover are plotted against lithostratigraphy in Figures 22,23, which
enable the definition of a precise biostratigraphy for these sections. Repre-
sentative specimens are illustrated in Figures 24, 25.

7.2.1 Abbots Cliff ChaZk Formation. Samples from the Abbots Cliff Chalk
yielded dinoflagellate cyst assemblages typically consisting of several
hundred to one thousand individuals (Figure 22). Twenty-seven species and
subspecies were recorded (Figure 22; Table 3), eleven of which are restricted
to this formation. Assemblages invariably contained Oligosphaeridium com-
plex (White) Davey and Williams as the most abundant cyst, accompanied
by common Batiacasphaera euteiches (Davey) Davey [Figure 24(c)],
Cyclonephelium distinctum Deflandre and Cookson [Figure 25(g)],
C. membraniphorum Cookson and Eisenack, Hystrichosphaeridium palmatum
(White ex Bronn) Downie and Sarjeant, Odontochitina costata Alberti;
emend. Clarke and Verdier [Figure 24(j)] and 0. operculata (0. Wetzel)
Deflandre [Figure 24(i)]. All of these species continue up into the overlying
Plenus Marl (Figure 22). Other common constituents were Cleisto-
sphaeridium armatum (Deflandre) Davey, C. clavulum (Davey) Below,
Prolixosphaeridium conulum Davey [Figure 25(h)], Pterodinium cingulatum
reticulaturn (Davey and Williams) Lentin and Williams [Figure 24(b)] and
Rhynchodiniopsis sp. cf. R. aptiana Deflandre [Figure 24(g), (h), (k)].
Eleven species were restricted to the Abbots Cliff Chalk (Figure 22) and of
these, seven [Aldorfia deflandrei (Clarke and Verdier) Stover and Evitt,
Cleistosphaeridium armatum, Exochosphaeridium bifidum (Clarke and Ver-
dier) Clarke et al., Leberidocysta chlamydata (Cookson and Eisenack) Stover
and Evitt, Pterodinium cingulatum reticulatum, Surculosphaeridium? longifur-
catum (Firtion) Davey et al. and Xenascus ceratioides (Deflandre) Lentin and
Williams] have been recorded from post-Cenomanian sediments elsewhere
in the Anglo-Paris Basin (Clarke and Verdier, 1967; Foucher, 1979, 1981;
Foucher in Robaszynski et al., 1980, 1982; Jarvis and Tocher, 1983; Tocher
and Jarvis, 1987; Jarvis et al., 1987, 1988).
Cribroperidinium exilicristatum (Davey) Stover and Evitt and Fromea
granulosa (Cookson and Eisenack) Stover and Evitt [Figure 25(k)] were also
restricted to the Abbots Cliff Chalk, and have not been recorded from
sediments younger than Cenomanian elsewhere in southern England
(Davey, 1969; Jarvis et al., 1988). Oligosphaeridium albertense (Pocock)
Davey and Williams [Figure 25(i)], a third restricted species, is recorded for
the first time from the Anglo-Paris Basin but elsewhere it ranges throughout
the mid and upper (Barremian to Maastrichtian) Cretaceous (Yun, 1981;
Below, 1982).
A cyst referred to Rhynchodiniopsis sp. cf. R. aptiana occurred only in the
lllicrofossil Assemblages and the Oceanic Anosic Event 63

(i)
* ”

1~1g:ure 25. Crnoman~an-‘l’u~~)~llan dmofla~ellate cysts from Dmw 11. Light phot~)mlc~oKraphs of wmc
representative sprcnnens. (a) Oligospharridium poculum Jain. AKS/D 19, x 300;(h) Oligosphaeridirrm
poculum Jain, XKS/DlO, x 300; (c) Kallosphaeridium ringnesiowm (Illanum and Cookson) Tocher
and Jarvis, ABC/DlS, x 300; (d) Kallosphaevidium ringnesiovum (1lanum and Cookson) Tocher and
Jar\is, ABC;D15,, x 300; (e) Oligosphaeridiumpvolixispinosum Davey and Williams. ABC/D-b, x 300;
(f) Oligosphaeridzum pwlixispinosum Dabey and Williams, ABC/D+, x 300; (g) (‘_wlone~helirrm
distincturn Deflandre and Cookson, ABC/Dl2. x 300; (h) Pvolisosphaeridium conulum Davey,
ABC/D6, x 300; (i) Oligosphaeridium albertense (Pocock) Davey and Williams, ABC;D4, x 300; (j)
Leberidocysta dej7occata (Davey and Verdier) Stover and Evitt, ABC:I>Z, x 300; (k) Fromeagmnulosu
(Cookson and Eisenack) Stover and Evitt. ABC/D6, x 300; (1) Knllospkar~idium vingnesiorum
(\\lanum and Cookson) Tocher and Jarvis, ABCiDl-l, x 300.
Microfossil Assemblages and the Oceanic Anoxic Event 65

Abbots Cliff Chalk [Figures 22, 24(g), (h), (k)]. Rhynchodiniopsis uptiana has
previously been recorded from Aptian and older sediments (Deflandre, 1935;
Below, 1981). It is unclear whether our material from Dover is referable to
that species or represents a closely related form. If the former can be
demonstrated, then our records indicate that either the range of the species
should be extended into the Late Cenomanian, or that reworking of Aptian
sediments occurred during the Late Cenomanian in the Dover area.

7.2.2 Plenus Marl Formation.. The Plenus Marl yielded nineteen species and
subspecies of dinoflagellate cysts (Figure 22; Table 3). Assemblages in the
lower beds (l-3) are dominated by the same seven species which are
commonest in the Abbots Cliff Chalk (Figure 22), although 0. complex does
not invariably remain the most abundant cyst. The numbers of cysts
recovered from the lower beds of the Plenus Marl are generally lower (several
horizons yielded < 100 individuals per sample) than in the underlying
formation, and specific diversity is also lower (Figure 22). Cleistosphae-
rid&m? aciculare Davey [Figure 24(e)] and Kiokansium polypes (Cookson
and Eisenack) Below, which occur rarely in the Abbots Cliff Chalk, both
disappear within the lower beds of the Plenus Marl (Figure 22). The former
species has not been recorded in the Anglo-Paris Basin before, although it
has been found in Albian-Cenomanian sediments elsewhere (Davey, 1969),
whereas the latter is known from post-Cenomanian sediments within the
basin (e.g. Tocher and Jarvis, 1987). Heteros@zaeridium? heteracanthunz
(Deflandre and Cookson) Eisenack and Kjellstrom, Oligosphaeridium pro-
lixispinosum Davey and Williams [Figure 25(e), (f)] and Prolixosphaeridiunz
conulum [Figure 25(h)] also disappear in Beds l-3, but all reappear in the
lower Turonian.
.4 pronounced drop in cyst abundance (samples typically contained 50 or
fewer individuals per sample), and diversity occur in Bed 4, followed by the
progressive disappearance of most cysts in Beds 5 and 6 (Figure 22). This
decline is, however, initially accompanied by increased proportions of
Kallosphaeridium ringnesiorum (Manum and Cookson) Tocher and Jarvis
[Figure 25(c), (d)]. This species is as a minor component of the assemblage in
Bed 1 (Figure 22) but constitutes 60°0 of the cysts in Bed 5 and 96”, in Bed 6.
The uppermost beds of the Plenus Marl contain few dinoflagellate cysts, Bed
7 yielding only two specimens of K. ringnesiorum and Bed 8 a single example
of Odontochitina operculata. The appearance of K. ringnesiorum in the upper
part of Bed 1 at Dover, is the earliest record of this species in the Anglo-Paris
Basin.
The distribution of dinoflagellate cysts through the Plenus Marl suggests
that the environmental changes, which in Bed 4 caused the decline of the
dominant cyst assemblage, initially favoured K. ringnesiorium, allowing it to
reach an acme immediately prior to the disappearance of all cysts at the
bottom of the Dover Chalk.
It is noteworthy that the decline in dinoflagellate cysts coincides with a
marked increase in the proportion of calcispheres (Section 2.2). At least some
calcispheres (particularly Pithonella spp., a form which is common in the
Plenus Marl) may be calcareous-walled dinoflagellate cysts (Keupp, 1979).
The inverse relationship between the two groups, therefore, suggests that the
decline in organic-walled dinofagellate cysts during the latest Cenomanian
and earliest Turonian was balanced by an increase in calcareous-walled
groups.

7.2.3 Dover Chalk Formation. Twenty-two species and subspecies of dino-


flagellate cysts were rcorded from the Dover Chalk (Table 4; Figures 22,23),
although the basal beds, including those spanning the Cenomanian-
Turonian boundary, proved barren. Rare cysts (represented initially by a
single example of Oligosphaeridium complex) reappear approximately 4m
above the base of the formation, but remain uncommon (generally < 10
individuals per sample) throughout the Melbourn Rock Beds (Figures 22,
23). There is a progressive increase in the abundance and diversity of cysts
towards the top of the sampled interval, with assemblages from the top of the
succession typically consisting of SO-100 individuals.
The first appearance of Hystrichosphaeridium dificile Manum and Cook-
son, a low Turonian indicator (Tocher and Jarvis, 1987), occurs towards
the top of the Melbourn Rock (Figure 23), approximately 7.5 m above the
base of the Dover Chalk. Senoniasphaera rotundata Clarke and Verdier
[Figure 24(a)] is first recorded from immediately above the top of the
Melbourn Rock, confirming a low Turonian age for the appearance of this
species (Foucher, in Robaszynski et al. 1982; Tocher and Jarvis, 1987).
Senoniasphaera microreticulata Brideaux and McIntyre, a species which also
appears in the low Turonian elsewhere (Tocher and Jarvis, op cit.), is as a
major component of the assemblages from the uppermost part of the
succession, appearing approximately 1 m above the Round Down Marl
(Figure 23).

7.3 Discussion

Data presented in previous palynological studies of Cenomanian-Turonian


boundary sequences in southern England (and elsewhere in the Anglo-Paris
Basin e.g. Foucher in Robaszynski et al., 1980,1982; Pas-de-Calais, Saumur,
Touraine), indicate that major differences exist between the dinoflagellate
cyst assemblages recorded by us from Dover and those which exist in coeval
sediments from other parts of the basin.
Dinoflagellate cyst assemblages from the Abbots Cliff Chalk at Dover are
dominated by Oligosphaeridium complex, and to a lesser extent by species
such as Cyclonephelium distinctum, Odontochitina costata and 0. operculata,
species which are thought to be relatively tolerant, cosmopolitan forms
(Tocher and Jarvis, 1987). These assemblages compare favourably, there-
fore, with that described from an equivalent horizon near Saumur (Foucher,
in Robaszynski et al., 1982). Equivalent samples from Pas-de-Calais
(Foucher in Robaszynski et al., 1980), the Isle of Wight (Clarke and Verdier,
1967) and Devon (Jarvis et al., 1988), however, contain assemblages which
are more or less dominated by Palaeohystrichophora infusorioides D&an&e,
a species which is absent at Dover. Furthermore, various forms of Achomos-
phaera and Spiniferites, which are regarded as typical shelf-dwelling forms
(Toucher and Jarvis, 1987), are found in the assemblages from all of these
areas, but not at Dover.
Dinoflagellate cyst assemblages in the lower beds (1-4) of the Plenus Marl
iLlicrofossil Assemblages and the Oceanic Anosic Event 67

at Dover are characterised by the same species which are abundant in the
underlying Abbots Cliff Chalk. The upper beds (S-7), however, contain
mainly Kallosphaeridium ringnesiorum. While forms such as C. distinctum and
0. complex are relatively common throughout the basin (e.g. Foucher in
Robaszynski et al., 1980, 1982), B. euteiches and K. ringnesiorum are rarer.
The latter two species are thick-walled proximate cysts [Figures 24(c), 25(c),
(d)], and have not been recorded from the Plenus Marl or its equivalents at
any other localities.
Dinoflagellate cyst assemblages from the Plenus Marl in the Pas-de-Calais
(Foucher, in Robaszynski et al., 1980) and Isle of Wight (Clarke and Verdier,
1967), and the Plenus Marl equivalent in SE Devon (Jarvis et al., 1988)
continue to be dominated by Palaeohystrichophora infusorioides and Spin~fe-
rites ramosus, as in the underlying chalks, although a local abundance of
Microdinium irregulare Clarke and Verdier occurs in the Isle of Wight
succession. The assemblage described from the lateral equivalent of the
Plenus Marl in Saumur (Foucher, in Robaszynski et al., 1982) was relatively
diverse but abundances were low, and no particular form was noted as being
dominant. Nevertheless, this assemblage also contains P. infusorioides and
S. ramosus. ‘fhe anomalous composition of the dinoflagellate cyst assem-
blages noted in the Abbots Cliff Chalk, therefore, continues through the
remainder of the Cenomanian at Dover, prior to the disappearance of all
cysts near the top of the stage.
Samples from the basal 4m of the Turonian Dover Chalk Formation
at Dover proved barren, but above this Oligosphaeridium complex and
Cyclonephelium distinctum are again the dominant cysts. There are also
localised abundances of Batiacasphaera euteiches, Kallosphaeridium
ringnesiorum, Senoniasphaera microreticulata and the low ‘l’uronian indices
Hystrichosphaeridium dificile and Senoniasphaera rotundata.
By contrast, an assemblage recorded from a comparable interval on the Isle
of Wight (Clarke and Verdier 1967) is dominated by Spiniferites ramosus and
Odontochitina costata, and to a lesser extent by Cyclonephelium membrani-
phorum and Canningia colliveri. In SE Devon (Jarvis and Tocher, 1983;
Tocher and Jarvis, 1987; Jarvis et al., 1987, 1988) the most abundant
forms are Palaeohystrichophora infusorioides, S. ramosus and Cyclonephelium
distinctum. Samples from the ‘Craie marneuse’ of Saumur (Foucher, in
Robaszynski et al., 1982) are dominated by 0. operrulata and Oligosphae-
ridium complex but contain common Spiniferites and P. infusorioides,
while at Civray-de-Touraine the assemblages are also dominated by the
S. ramosus group and P. infusorioides.
Clearly, therefore, there is good evidence for significant geographical
variation in Cenomanian and Turonian dinoflagellate cyst assemblages, with
the Dover samples in particular proving atypical.

8. Stable-isotope geochemistry (G.A.C. & I.J.)

The use of stable isotopes, particularly carbon isotopes (Si3C), as a means of


correlating Cretaceous sequences has been demonstrated previously (Scholle
and Arthur, 1980; Letolle and Pomerol, 1980; Pomerol, 1983; Jenkyns, 1985;
Hilbrecht, 1986; Hilbrecht and Hoefs, 1986; Schlanger et al., 1987; Jarvis
et al., 1988). The most distinctive feature of the Upper Cretaceous carbon-
isotope curve, is the major positive 6t3C ’ excursion’ or ‘spike’ which occurs
worldwide, close to the Cenomanian-Turonian boundary (Scholle and
Arthur, 1980, Figure 2; Schlanger et al., 1987, Figure 2).
It is argued that the high 6t3C values in Cenomanian-Turonian limestones
indicate an enrichment in the global ocean ’ 3C (Jenkyns, 1980,1985; Scholle
and Arthur, 1980; Jenkyns and Clayton, 1986; Graciansky et al., 1986;
Arthur et al., 1987; Schlanger et al., 1987). Increased 613C values resulted
from the preferential extraction of “C from sea water by marine plankton,
the organic matter of which was not recycled back into the oceanic reservoir
because of widespread burial of organic carbon in the ocean basins during the
OAE.
We have determined the 613C and 6t80 contents of 32 samples taken
across the Cenomanian-Turonian boundary in the Abbots Cliff and Akers
Steps sections (Table 5). These data have been plotted against lithology
(Figures 26, 27) to indicate stratigraphic trends, and to illustrate more easily
relationships between stable-isotope variation, fossil assemblages, and the
position of the stage boundary.

8.1 Material and methods


Small 5Og subsamples were taken in the field from 2 kg bulk-sediment
samples collected for foraminiferal and ostracod analysis (Section 4.1). Care
was taken to extract only unweathered material, all superficial sediment
being discarded. Individual samples were broken into - 0.5 cm fragments in
the laboratory. To avoid any residual contamination from weathered
material, the samples were then sieved and the <0.5 mm fraction was
discarded. The resulting chips were hand-picked to isolate any remaining
weathered or discoloured material, and then dried overnight at 105°C.
Samples of approximately 30g were finely ground in an agate mortar and
pestle, and the powders stored in glass phials.
Liberation of CO, for analysis was achieved by reacting 2 ml of anhydrous
100” ,) phosphoric acid with - 3 mg of sample at 25”C, using a method
similar to that described by McCrea (1950). The gas was analysed on a VG
Isogas SIRA 12 Mass Spectrometer at the Department of Earth Sciences,
Liverpool University. Standard correction procedures were employed and
the results expressed as per mil (“/,,) deviations from the PDB international
standard (Craig, 1957). The sample reproducibility [quoted as the mean
standard deviation (on) for replicates of individual powers] for both chalks
marl was 0.044’$, for carbon and O.O67Y&, for oxygen (n = 9).

8.2 Stratigraphic trends


Carbon isotope values show clearly defined trends through the sequence
(Figures 26, 27). The Abbots Cliff Chalk (Table 5) contains values ranging
from 2.3x, 613C at the base of the sample interval, to 2.7%” at the top of the
formation. The basal sample of the Plenus Marl (Bed la of Jefferies, 1962,
1963a, b) has a similar carbon isotope composition to the underlying bed.
Immediately above this, however, there is a sharp increase in 613C values
(Figure 26), which reach a peak of -4.7x, in Beds 6-8 of the Plenus Marl.
Microfossil Assemblages and the Oceanic Anoxic Event 69

Table 5. Carbon and oxygen stable-isotope data for uppermost Cenomanian to basal Turonian chalks,
Dover. Sample numbers correspond to those in Figure 3. All data are bulk sediment determinations
(ND, not determined; Bed numbers after Jefferies [1962, 19630, b]; MRB, Melbourn Rock Beds).

ABC 1 Abbots Cliff Chalk 2.246 -2.614


ABC 2 Abbots Cliff Chalk 2.375 -2.603
ABC 3 Abbots Cliff Chalk 2.376 -2.720
ABC 4 Abbots Cliff Chalk 2.462 -2.549
ABC 5 Abbots Cliff Chalk 2.504 -2.802
ABC 6 Abbots Cliff Chalk 2.677 -2.921
ABC 7 Plenus Marl (Bed 1a) 2.592 -3.242
ABC 8 Plenus Marl (Bed 1b) 3.009 -3.128
ABC 9 Plenus Mari (Bed 2) 3.929 -2.655
ABC 10 Plenus Marl (Bed 3) 3.826 -2.785
ABC 11 Plenus Marl (Bed 4) 4.259 -2.188
ABC 12 Plenus Marl (Bed 6) 4.677 - 1.988
ABC 13 Plenus Marl (Bed 7) 4.378 -2.895
ABC 14 Plenus Marl (Bed 8) 4.745 -2.413
ABC 15 Dover Chalk (MRB) 4.357 -3.165
ABC 16 Dover Chalk (MRB) 4.263 -3.193
ABC 17 Dover Chalk (MRB) 4.154 -3.235
ABC 18 Dover Chalk (MRB) 4.261 -3.093
ABC 19 Dover Chalk (MRB) 4.285 -3.007
ABC 20 Dover Chalk (MRB) 3.883 -3.179
ABC 21 Dover Chalk (MRB) 3.409 -3.471
AKS 1 Dover Chalk (MRB) 3.011 -3.825
AKS 2 Dover Chalk (MRB) 2.805 -3.346
AKS 3 Dover Chalk (MRB) 3.070 -2.758
AKS 4 Dover Chalk (MRB) 2.796 -3.055
AKS 5 Dover Chalk (MRB) 2.851 -2.936
AKS 6 Dover Chalk (MRB) 2.625 -3.127
AKS 7 Dover Chalk (ZIRB) 2.766 -2.824
AKS 8 Dover Chalk 2.810 -2.519
AKS 9 Dover Chalk SD ND
AKS 10 Dover Chalk 2.818 -2.544
AKS 11 Dover Chalk 2.918 -2.645
AKS 12 Dover Chalk 2.695 -2.657
ARS 13 Dover chalk ND ND

Carbon isotope values decline steadily in the basal beds of the overlying
Dover Chalk (Figure 26), reaching values of 3.0x, approximately 2 m above
the base of the formation. Remaining samples in the lower Dover Chalk
exhibit values of 2.6-3.1x, 613C (Table 5). Th e carbon stable-isotopes thus
display a major ‘excursion’ occurring in the Plenus Marl and basal Dover
Chalk (Melbourn Rock Beds). The peak of the anomaly occurs at the top of
the Plenus Marl, immediately below the formational boundary and, there-
fore, also below the base of the Turonian as defined by appearance of
Mytiloides inoceramid bivalves (Section 3.4).
Comparable carbon stable-isotope data through the Plenus Marl have
recently been presented by Schlanger et al. (1987, figure 11) for shore
sections at the eastern end of Shakespeare Cliff (Figure 1). These authors
additionally sampled the uppermost 1.5 m of the Abbots Cliff Chalk and the
basal 1 m of the Dover Chalk (Melbourn Rock Beds). Schlanger et aI.‘s
(1987) data differ from ours in two respects. Firstly, their 613C values reach a
70 I. Jarvis rt t/l

6 13c %o [PDB) 6 180 %o (PDBI

3c

H. archaeocretacea

.
I

2 1c

02 oc

1 9c

-1 8C

:
;I

i -1 2:
1 1C
DC

,30
13C

70

t50

R. cushmani TRZ

:
:
:

Figure
!

26. Carbon
Cenomanian/Turonian
and oxygen
boundary
c I

stable-isotope
, I

distribution
at Dover. Major lithologies
I
I ’

and iithostratigraphy
and symbols are explained
I ’ I

across the
in Figures 5
and 6.
Microfossil Xssemblages and the Oceanic Anosic Event 71

6 13c %o (PDBI 6 I80 %o (PDBI


4
7
L -f
I,
I
<

_ 1:

1;

- 1.

l( 0

H. archaeocretacea
PRZ
c I

I 0

---___ I-
0
_- -

, 0

\
.
0
I
I
0

I I
gurt’ 27. Carbon and oxygen stable-lsotopedistrlbuti(ln and lithostratygaphy III the higher beds of the
lower Turonian at Dover. The succession is a continuation of that shwn In Figure 26. Lithologies
and sgnhol< at’r explained in P~gurr 5. 6 and 8.
maximum within the Melbourn Rock Beds (although still within the
Cenomanian), and not within the upper Plenus Marl; their values do not
show any decline within their sampled interval (despite their uppermost
sample being from - 2 m above the base of the Melbourn Rock). Secondly,
the isotopic variation in the section is greater in our samples (2.3-4.8x, 6ljC)
than that illustrated by Schlanger et al. (1987). Their data display constant
values of - 3x, 6l 3C f rom 1.5 m below the base of the Plenus Marl up to Bed
4, and then an increase to a maximum of 4.2x, in the basal Melbourn Rock
Beds. The reasons for these differences in detail are unclear, but it is
noteworthy that our carbon stable-isotope data correspond more closely with
trends displayed by associated microfossil assemblages (see below).
The oxygen stable-isotope data (Figures 26,27, Table 5) may cautiously be
interpreted as having a ‘background’ value of approximately -2.6x, 6180.
Within the Plenus Marl, the oxygen isotope values show considerable
variation, with a slight positive excursion around Beds 6-8. Within these
beds, there is a positive correlation between 613C and 6lsO values. The most
negative value is - 3 .S%, which occurs approximately 2 m above the base of
the Dover Chalk. Oxygen values return to ‘background’ levels immediately
above the Melbourn Rock Beds (Figure 27). The positive correlation of 613C
with 6180 in the upper beds of the Plenus Marl is contrary to the negative
correlation observed in German sections by Hilbrecht and Hoefs (1986).
These authors interpreted the oxygen data as indicating lower sea-water
temperatures during the latest Cenomanian. Our results correspond more
closely with those of Scholle and Arthur (1980) who noted a drop in al80
values at, or close to, the Cenomanian-Turonian boundary. At this stage,
therefore, we prefer to use the oxygen stable-isotope data solely as a means of
isolating diagenetically altered data points (cf. Jarvis et al., 1988), and to
indicate any obvious diagenetic trends. We do not as yet place any sig-
nificance on the stratigraphic variation in the 6180 curve.

8.3 Possible diagenetic factors

Diagenesis will only alter bulk-sediment isotope values if there has been
significant interaction between the sediment and the surrounding pore-fluid
(cementation or recrystallisation). Such interaction commonly occurs during
early diagenesis in sediments composed of metastable mineralogies (arag-
onite or Mg-calcite). However, the Chalk was deposited as predominantly
low Mg-calcite and, with the possible exception of hardground horizons,
is unlikely to have undergone significant early diagenetic modification.
Furthermore, the carbon stable-isotopes of any metastable mineralogies
which were present would probably not have altered extensively since,
compared with oxygen stable-isotopes, carbon isotopes are relatively
immune to diagenetic modification. This is because the volume of carbon
in carbonate sediments vastly exceeds that in solution within their coexisting
pore waters, and the fractionation of carbon isotopes between calcium
carbonate and dissolved bicarbonate is relatively small at near surface
temperatures (Emrich et al., 1970).
Nevertheless, diagenetic modification of bulk-rock carbon stable-isotope
ratios in pelagic limestones can occur. Precipitation of carbonate cements in
Microfossil Assemblages and the Oceanic Anoxic Event 73

association with the post-depositional oxidation of organic matter may


lighten 613C values (Scholle and Arthur, 1980), although these processes
would reduce a carbon ‘excursion’ rather than emphasise it. One possible
cause for a diagenetic increase in 613C would be the precipitation of
carbonate by bacterial methanogenesis (Hudson, 1977; Irwin et al., 1977).
This process occurs in organic-rich sediments below the sulphate reduction
z&e, but is unlikely to have occurred to any extent in the Chalk, where lam
sedimentation rates (estimated to be 2-4 cm ka- ’ by Hancock, 1975) would
have ensured the destruction of reactive organic matter in shallower (oxic
and post-oxic) environments.
Oxygen isotope ratios are far more readily altered than S’ 3C values during
diagenesis, partly because oxygen isotopes show large temperature-related
fractionation. Primary carbonate aI80 values will largely reflect ocean-water
temperatures and salinities (Savin, 1977; Anderson and Arthur, 1983), but
they will be modified considerably by the addition of cements during burial
diagenesis. Finally, recrystallisation after uplift may involve the precipita-
tion of new isotopically light carbonate cements during meteoric diagenesis.
Untectonised onshore European chalks have average al80 values of - 2.9%”
(Scholle, 1974, 1977), generally ranging between - 2 and -4x,. Burial and
later diagenesis, however, leads to lighter oxygen isotope compositions,
producing values of -ST&, or less (Jlzrrgensen, 1987). The oxygen stable-
isotope values reported here (Table 5) are closely comparable to those of
average chalks, and are only marginally lighter than expected values for
carbonate produced by late Cretaceous coccoliths (estimated to be between
-0.5 and --2.00x, 6180 by Jorgensen, 1987). Our 613C and 6’*0 values do
not covary (Figure 28), and there is, therefore, no evidence for the major
diagenetic modification of any individual sample.
It is apparent, however, that most marls exhibit higher al80 values (and
possibly a1 3C values) than their adjacent chalks (Figures 26, 27). This could
be explained by the carbonate in the marls being less diagenetically altered
than that in the chalks. However, marls generally display more extensive
evidence of compaction and pressure solution than adjacent chalks (Garrison
and Kennedy, 1977; Scholle et al., 1983). It has been argued (Scholle, 1977)
that dissolution of carbonate occurs preferentially in carbonate-poor beds
(marls) and that this carbonate commonly precipitates in adjacent carbonate-
rich horizons (chalks) as an overgrowth cement. This results in an accentu-
ation of primary sedimentary variation and bedding rhythmicity.
Redistribution of small amounts of carbonate from marls into chalks at
elevated temperatures (i.e. during burial) would explain why the 6180 values
shift to more negative values in the more carbonate-rich intervals of the
Plenus Marl (Figure 26). Limited cementation might also partly explain (cf.
Scholle, 1974, figure 23, 1977; Hudson, 1977) the apparently lighter 613C
values of these chalks, although it would be expected that such small shifts in
6 “0 would be accompanied by only negligible changes in 6’ 3C values (see
above). Petrographic studies demonstrate, however, that coarse-grained
spars are absent in both chalks and marls, cementation being limited to
micrite and microspars even within strongly indurated horizons.
It is concluded that although oxygen-stable isotope values demonstrate
that limited diagenetic modification has occurred in the sequence, these
74

5
0 chalk

1 * marl

4-
.

d3C %o
(PDBI l

-4 -3 -2

d80 960 (PDB)

Figure 2X. Carbon versus oxygen stable-isotope values for the Cenomanian-Turonian boundary
succession at Dover. Values are listed in Table S. Note the absence of any correlation between the two
sets of data.

changes have had a minimal effect on 6i3C values. Maximum changes may be
up to -0.4x, 613C ( as estimated from adjacent chalk-marl values; Figure
26), but these lie within a positive carbon isotope ‘excursion’ of - 2.5x,.

9. Stratigraphic position and correlation of the OAE (I. J. et al.)

Previous papers on the Cenomanian-Turonian OAE have suffered from the


use of vague and loosely defined biostratigraphic frameworks. The position
of the carbon stable-isotope excursion has generally been referred to the
Actinocamax plenus Zone (e.g. Pomerol, 1983) or the ‘Hedbergella’
(= Whiteinella) archaeocretacea PRZ (e.g. Schlanger et al., 1987). A more
rigorous approach was adopted recently by Hilbrecht (1986; Hilbrecht and
Hoefs, 1986), who presented stable-isotope and key macrofossil and plank-
tonic foraminiferal records from boundary sequences in Germany.
The precise stratigraphic position and probable synchroneity of the
Cenomanian-Turonian OAE can only be demonstrated if detailed macro-
fossil and microfossil biostratigraphies are integrated with carbon stable-
isotope data. The extact correspondence between stable-isotope and bio-
stratigraphic information is also required if the effects of the OAE on the
marine biota are to be assessed accurately. These data, in turn, provide a
necessary foundation for modelling possible palaeoceanographic causes and
consequences of the event.
blicrofossil Assemblages and the Oceanic =\noslc Event 75

9.1 Integration of stable isotope and biostratigraphic data from Dover

Our stable-isotope data demonstrate that there is a major increase in 613C


vralues within Plenus Marl Bed 1 at Dover. This marks the onset of the OAR,
the effects of which are rapidly seen in the benthonic fauna, with the
extinctions of many common Cenomanian macrofossil, foraminiferal and
ostracod species at the top of Bed 1 (Figure 29). The disappearance of the
planktonic foraminifera Rotalipora greenhornensis also occurs at this level.
This major extinction event is, however, immediately preceded by the
appearances of several new benthonic and planktonic species (Figure 29),
including the ammonite Metoicoceras geslinianum, the ostracods Phodeucy-
there cuneiformis (which is restricted to Bed 1) and Isocvthereis elongata,
the calcareous nannofossil Helicolithus sp. .A, and the dinoflagellate cyst
Kallosphaeridium ringnesiorum.
A second major biotic event occurs in association with a further marked
increase in 613C values at the top of Plenus Marl Bed 3, where the ammonite
Calycoceras naviculare and the planktonic foraminifera Rotalipora cushmani
both become extinct (Figure 29). The disappearance of the latter marks the
top of the internationally recognised R. cushmani TRZ. These extinctions are
again associated, however, with appearances (Figure 29) of new species, the
belemnite ,4ctinocamaxplenus, the bysally attached pterioid bivalve Oxytoma
seminudum, and the ammonite Sciponoceras gracile. A diversification of the
dicarinellid planktonic foraminifera (including the appearance of Dicarinella
imbricata) also takes place at this level.
Carbon stable-isotope values reach a maximum of approximatelv 4.700,,
fii3C in Plenus Marl Beds 6-8. Disappearances within this interval include
macrofauna (A. plenus, M. geslinianum, 0. seminudum) and microfauna
(Textularia chapmani, Isocythereis elongata), but the greatest changes occur
in the microflora (phytoplankton) with the temporary disappearance of all
dinoflagellate cysts, the extinctions of four species of calcareous nannofossil
(Figure 29), and a proliferation of calcispheres. Unlike the earlier events, no
new forms appear at this horizon.
The Cenomanian-Turonian boundary is defined by the change in the
inoceramid bivalve fauna, with the replacement of Inoceramus pictus by
Mytiloides spp. This occurs at a point where carbon stable-isotope values
begin a marked decline, returning to background levels only 1 m above the
stage boundary. The ammonite Sciponoceras gracile and the calcareous
nannofossil Microstaurus chiastius both disappear close to the stage bound-
ary. The end of the carbon stable-isotope excursion within the lowest
Turonian is accompanied by the appearances of the planktonic foraminifera
‘Hedbergella’ archaeocretacea and the calcareous nannofossil Quadrum gart-
neri. There follow the progressive appearances of new species in all fossil
groups through the lower Turonian, and the reappearances of most dino-
flagellate cyst species. It is noteworthy that no marked extinction or
evolutionary events occur through the remainder of the sampled interval.
To summarise, the onset of the carbon stable-isotope excursion is
accompanied by a marked decline and turnover in the benthonic fauna, with
the extinctions of many typical Cenomanian forms (Plenus Marl Bed 1
event). Comparable changes in the planktonic fauna occur much later
Macrofauna Benthonic Ostracods
foraminifera

PM Plenus Marl

ACC Abbots Cliff


Chalk

1:1gure 29. Biostratigraphlc summary chart for the Cenomaman-Turonian boundary succession at
I>o\-rr. The stratigraphIc ranges of representative macrofossil and microfossil species are indicated
bv the vertical lines. Dashed vertical lines indicate intermittent occurrences; the discontinuous line
w:ithin the Gacelinelln veussi G. berthelini plexus indicates a shift from dominantly G. veussi to
dominantly G. berth&G morphotypes. Lithostratigraphic units are drawn to scale. The horizontal
stippled area indicates the stratigraphic extent of the carbon stable-isotope excursion. Dtscontinuous
horizontal lines mark lithological boundaries (base of the Plenus Marl, base of the Melbourn Rock,
top of the Melbourn Rock). The continuous horizontal line indicates the base of the Turonian.
defined by the appearance of Mytifoidrs spp. inoceramid bivalves.
I

I Cenomanlan I Turontan I
(Plenus Marl Bed 3 event). Major extinctions and disappearances only occur
in the microflora (phytoplankton) at the peak if the isotope excursion in
Plenus Marl Beds 6-8. The Cenomanian-Turonian boundary occurs at a
level where 613C values are declining, typical ‘I’uronian fossil species
appearing gradually throughout the Early Turonian. It is clear, therefore,
that there is a close correspondence between the 6’ 3C curve and a sequence of
floral and fauna1 events which accompany the Cenomanian-Turonian
boundary succession.

9.2 Correlation with the NE Atlantic continental margin

Recent work on sequences sampled by the Deep Sea Drilling Project


(DSDP) from the NE Atlantic (Graciansky et al., 1986, 1987; Herbin et al.,
1986; Kuhnt et al., 1986; Stein, 1986a, b; Stein et al., 1986) provides an
extensive data base for that region. The succession on the Goban Spur at
DSDP Site 551 is of particular significance since it is located near the edge of
the NW European continental slope, only 1100 km WSW of Dover. Here,
the boundary sequence consists of 4 m of barren, black, organic-rich (111,
Co_,) laminated mudstones, which contain abundant radiolarians and have
bulk-sediment carbonate S1 3C values of -4x, (Cunningham and Kroop-
nick, 1985). Rotalipora cushmani and R. greenhornensis both disappear
immediately below the base of these mudstones, and Helvetoglobotruncana
praehelvetica appears immediately above them (Hart, 1985). These data
indicate that the mudstones are an exact, but highly attenuated, equivalent of
the Plenus Marl and basal Melbourn Rock at Dover. A stratigraphic gap at
the base of the mudstones is suggested by the synchronous disappearances of
both species of Rotalipora, since at Dover R. cushmani ranges higher than
R. greenhornensis.

9.3 Correlation with SE Devon

The most western onshore Cenomanian-Turonian boundary succession in


southern England is located in SE Devon, where the sequence is also highly
attenuated (Jarvis and Woodroof, 1984), consisting of ‘marginal’ detritus-
rich limestones and calcarenitric chalks. The carbon stable-isotope strati-
graphy, and foraminiferal and dinoflagellate cyst biostratigraphies of a
representative section at Hooken Cliffs, near Beer, have been described
recently by Jarvis et al. (I 988). H ere, the base of the carbon stable-isotope
excursion occurs immediately above the surface of a local hardground in the
upper part (Pinnacles Member) of the Cenomanian Beer Head Limestone.
In contrast to the gradual increase in carbon stable-isotope values seen at
Dover, at Hooken Cliffs there is a sudden shift of + 2x, 613C (Jarvis et al.,
1988), producing values of >4.4%, for the white chalk matrix in the bed.
Intraclasts from above the hardground surface, however, yielded carbon
stable-isotope values which lay midway between those of the hardground
(and the underlying limestones), and their white chalk matrix. These, and
other data, were interpreted by Jarvis et al. (1988) as indicating that a
significant hiatus (equivalent to the uppermost Abbots Cliff Chalk and
Plenus Marl Beds I-3) occurs at the surface of the hardground.
hIicrofossi1 Assemblages and the Oceanic Anoxic Event 79

Rotalipora disappears suddenly at the level of the main carbon stable-


isotope shift in SE Devon, below which it constitutes a major proportion
of the foraminiferal assemblage (Jarvis et al., 1988; Figure 5). This obser-
vation provides further evidence for a stratigraphic gap, since at Dover
there is a gradual decline in the proportion of Rotalipora which cor-
relates with increasing 6i3C values. In both areas Actinocamax plenus
appears immediately above the extinction of Rotalipora. An assem-
blage of reworked dinoflagellate cysts is also associated with the sudden
stable-isotope shift in the SE Devon succession.
The Cenomanian-Turonian boundary is precisely defined in SE Devon by
the appearance of abundant Mytiloides spp. bivalves and occasional Watino-
ceras spp. ammonites, immediately above a younger hardground surface
(Haven Cliff Hardground), which marks the base of the Seaton Chalk
Formation. Carbon stable-isotope values at this level are ~3.9%~ 6i3C,
which compare favourably to values obtained from beds close to the stage
boundary at Dover (Section 8.2). The appearances of low Turonian biostrati-
graphic index microfossil species such as Helvetoglobotruncana praehelvetica
and the dinoflagellate cyst Senoniasphaera votundata occur < 1 m above the
stage boundary in SE Devon, indicating that this part of the sequence is also
severely condensed compared with that at Dover. It is concluded that there is
good general agreement between the carbon stable-isotope and biostrati-
graphic data from the two areas, and that the OAE was, therefore, probably a
synchronous event throughout southern England.

9.4 Correlation with Humberside and the North Sea

Comparable carbon stable-isotope (Schlanger et al., 1987, p. 375), macro-


fossil (Jefferies, 1963~) and microfossil (Hart and Bigg, 1981) data are avail-
able for the Cenomanian-Turonian boundary succession in Humberside, NE
England. This sequence includes a thin (< 1 m) unit of dark-coloured
organic-rich marls at the base of the Welton Chalk Formation (Wood and
Smith, 1978), known collectively as the ‘Black Band’. A prominent omission
surface occurs at the base of the sequence, which is approximately 25 cm
below the main bed of black marl (the Black Band sense strict0 cf. Jefferies,
1963~).
As in SE Devon, the carbon stable-isotope curve (Schlanger et al., 198’7;
Figure 5) displays a sudden shift, from < 30/,, 6i3C below the omission
surface, to -4.3x,, immediately above it. This suggests that a stratigraphic
gap equivalent to the lower Plenus Marl (Beds l-3) at Dover, occurs in the
Humberside succession. The presence of a hiatus is supported by biostrati-
graphic data, particularly the last occurrence of Rotalipora cushmani immedi-
ately below the omission surface (Hart and Bigg, 1981), and the first records
of Actinocamax plenus from immediately above it (Jefferies, 1963~).
The precise position of the Cenomanian-Turonian boundary is uncertain
due to the lack of detailed inoceramid bivalve or ammonite records. The key
foraminiferal species Helvetoglobotruncana praehelvetica and H. helvetica
appear a short distance above the main black marl, indicating that this part of
the succession is also severely attenuated (cf. SE Devon) compared with the
equivalent interval at Dover. Carbon stable-isotope values fall rapidly within
the black marl (Schlanger et a/., 1987), reaching values of -3j’;,, 8°C
immediately above it, and remaining at this level for at least the next 8 m of
sediment. By comparison with Dover and SE Devon, therefore, the isotope
values suggest that the stage boundary falls within the main black marl of the
‘Black Band’, as indicated previously by Schlanger rt al. (1987).
Despite variations in detail, the Plenus Marl constitutes a relatively
uniform facies throughout most of southern England (Jefferies, 1962,
1963a, b). The Plenus Marl of the North Sea, however, is lithologically
distinct from its onshore equivalent, consisting of up to several metres of
black to dark brown organic-rich calcareous mudstones (Burnhill and
Ramsay, 1981). These mudstones yield radiolaria and planktonic foramini-
fera, but contain only a sparse benthonic fauna. Rotalipora disappears at the
top of the immediately underlying Hidra Formation, but no stable-isotope
data are currently available from the succession. Lithological similarities to
the ‘Black Band’ of Humberside, however, suggest that the North Sea Plenus
Marl is also equivalent to the Plenus Marl and basal Melbourn Rock of
southern England.

9.5 Correlation with Germany

Hilbrecht & Hoefs (1986) concluded from their study of German


Cenomanian-Turonian boundary successions (which included a review of
Schlanger et aZ.‘s 1987 data from Germany), that the base of the carbon
stable-isotope excursion coincided with the extinction of Rotalipora cush-
maw’ and the appearance of Actinocamax plenus. They placed the upper limit
of the excursion immediately above the appearances of Mytiloides spp. and
Helvetoglobotruncana ‘helvetica’. These conclusions were, however, based on
samples collected at intervals of several metres (Hannover-Misburg, NW
Germany; Hilbrecht and Hoefs, 1986; Figure 2), or on data from successions
containing a demonstrable stratigraphic gap (Muhlberg, SE Germany;
Hilbrecht and Hoefs, 1986; Figures lb, 3). Their work does not, therefore,
possess the stratigraphic resolution of that presented here for Dover.
Nevertheless, strong similarities exist between the English and German
data. In both regions the boundary succession is marked by a facies change
from limestones to marls (see also Schlanger et al., 1987), the latter
containing abundant evidence of reworking, and containing prominent
omission and erosion surfaces. In both areas the disappearance of Rotalipora
greenhornensis occurs above the facies change, and below the extinction level
of R. cushmani (and the appearance of Actinocamax plenus). In both areas the
disappearance of R. cushmani occurs at a point on the carbon stable-isotope
curve where 613C values reach approximately 4x, (Figure 26; Hilbrecht and
Hoefs, 1986; Figures 2, 3). This exact correspondence between the two sets
of isotopic and biostratigraphic data provides strong additional support for
the stratigraphic use of carbon-stable isotopes, and suggests that the OAE
was synchronous in England and Germany. Our more detailed isotopic and
biostratigraphic data from Dover indicate, however, that the onset of the
OAE was before and not coincident with the extinction of Rotalipora. In both
areas the highest S13C values occur above this level.
Possible differences between the positions of the top of the carbon stable-
Microfossil Assemblages and the Oceanic Anoxic Event 81

isotope excursion are more difficult to assess. In both England and Germany
the top of the excursion lies immediately above the appearance of Mytiloides
spp. (i.e. base of the Turonian). Hilbrecht and Hoefs (1986) suggested that
this is coincident with the appearance of H. ‘helvetica’. This conclusion was
based, however, on data from attenuated successions in SE Germany, the
species being absent to the NW. In addition, the identification of H. helvetica
relies strongly on the species concepts of the individual micropalaeontolo-
gist, since it forms part of the Hedbergella delrioensis-Helvetoglobotruncana
praehelvetica-Hi. helvetica plexus. We record the appearance of H. praehel-
vetica several metres above the top of the carbon stable-isotope excursion at
Dover, while H. helvetica appears more than 20m above the base of the
Turonian, above the interval studied in detail here.

9.6 Stratigraphic conclusions and lateral zlariation


It is concluded that, within the limits of precision for available data, there is a
close correspondence between carbon stable-isotope trends and the ap-
pearances and disappearances of key fossil taxa in the Cenomanian-Turonian
boundary successions of the NE Atlantic margin, southern England, NE
England, the North Sea and Germany. These observations provide strong
evidence that the carbon stable-isotope excursion (and therefore the OAE)
was isochronous, at least throughout NW Europe.
The good agreement of absolute carbon stable-isotope values at key
biostratigraphic levels in both England and Germany, suggests that the
isotope stratigraphy may be used as a powerful independent tool for the
correlation of Upper Cretaceous pelagic sequences. Sudden shifts in carbon
stable-isotope values, like those in SE Devon and Humberside, are readily
explained by the presence of hiatuses. These hiatuses are confirmed by
biostratigraphic evidence, demonstrating that stable isotopes may addition-
ally be used as a means to identify and assess the extent of gaps in the
stratigraphic record.
Data are available, therefore, from a wide range of depositional environ-
ments, ranging from ocean margins (DSDP Site 551) representing original
water depths of l-2 km (Waples and Cunningham, 1985), through epiconti-
nental settings (Dover, Humberside, Germany and the North Sea) where
water depths were probably a few hundred metres (Hancock and Kauffman,
1979), to shallower areas (SE Devon) having water depths in the order of
100 m, or less (Jarvis and Woodroof, 1984). In most cases the Cenomanian-
Turonian boundary succession is lithologically, faunally and isotopically
distinct, consisting of carbonate-poor, locally organic-rich facies intercalated
within dominantly limestone successions. Possible palaeoenvironmental
causes of these anomalies will now be considered.

10. Palaeoceanography and the OAE (I. J. et al.)

The relationships between palaeoceanography and oceanic anoxic events are


the subject of an extensive literature. Data from Cenomanian-Turonian
boundary successions throughout NW Europe and its continental margins
(Section 9). however, provide us with lithological, fauna1 and geochemical
82 I. _lar\Is rt tri

data, which impose major constraints on constructing models of the


development of the 0.4E in the area.

10.1 Evidence for a Cenomanian- Turonian upwelling event

The key to understanding the palaeoceanography of the Cenomanian-


Turonian boundary event is the nature of the organic-rich mudstones which
were deposited widely in the ocean basins (e.g. Site 551) and in deeper-water
epicontinental areas (e.g. North Sea) at that time. These mudstones owe their
colour to a high concentration of dominantly marine (e.g. Stein, 1986a)
organic-matter. Radiolarians are also a major component in this facies. Both
of these characteristics indicate that the mudstones must have been deposited
under areas of abnormally high surface productivity. It is postulated,
therefore, that a major change in oceanic circulation during the latest
Cenomanian brought nutrient-rich Atlantic deep-m:ater towards the ocean
surface (cf. Summerhayes, 1987), causing a massive productivity event. Such
an upwelling event (Figure 30) is the only reasonable way to explain such a
massive change in ocean productivity.
High surface productivity invariably leads to the development of a marked
oxygen-minimum zone in the underlying water column (Figure 30), because
oxygen is utilised during the breakdown of organic matter as it rains
downwards from the surface waters. The exact relationship between the
development of strong oxygen-minimum zones and the deposition of
organic-rich sediments is the subject of some controversy. It is generally
agreed, however, that the deposition of organic-rich sediments is not directly
related to seawater oxygenation levels, but is controlled by rates of organic
matter production, water depth and bulk sediment accumulation rates (Seuss
et al., 1987; Calvert, 1987). The absence of organic-rich sediments, there-
fore, does not imply a lack of productivity or the absence of a strong oxygen-
minimum zone, since many modern upwelling areas (e.g. Gulf of California;
Calvert, 1987) are not areas of organic-rich sediment accumulation.
The presence of laminated mudstones of Cenomanian-Turonian age on
the NW European continental margin and in the North Sea indicates the
presence of anaerobic, possibly anoxic (sensu Rhoads and Morse, 1971; see
Section 11_2), bottom waters in these areas (Figure 30). The abundance of
radiolarians in the same sediments provides strong evidence that this anoxia
was related to the development of an intense osygen-minimum zone caused
by enhanced ocean-surface productivity (Figure 30). These conclusions are
supported by palaeoceanographic modelling (Barron in Arthur et al., 1987),
which predicts the occurrence of upwelling conditions off the NW European
continental margin at that time.
The occurrence of burrows throughout the Plenus Marl, demonstrates
that in shallower areas bottom-waters retained at least small amounts of
oxygen (i.e. remained oxic). This is consistent with the absence of radio-
larians from Plenus Marl facies, which suggests that productivity was also
probably lower (Figure 30). (Th e absence of radiolarians must be treated
with some caution, however, since evidence for widespread dissolution of
biogenic silica in the Chalk [Hancock, 1975; Clayton, 19861, means that the
lack of siliceous microplankton could equally be a result of diagenetic
Microfossil Assemblages and the Oceanic ;\nosic Event 83

NE Atlantic Anglo-Paris Basin North Sea


continental margin Basin
SITE 55, DEVON DOVER

I I I

SITE 55, DEVON DOVER

I I I

mid Cenomanian

basemen,

Figure 30. Schematic representation of the palaeoceanographic changes which may have caused the
Cenomanian-Turonian Oceanic Anoxic Event (OAE). The mid to early late Cenomanian was
characterized by low rates of oceanic turnover, and therefore only low to moderate surface
productivity. An oxygen-minimum zone was probably poorly developed in the water column, being
represented by relatively well-oxygenated water. Widespread deposition of nannofossil oozes took
place everywhere above the carbonate compensation depth (CCD). A rapid increase in the rate of
oceanic turnover during the latest Cenomanian (possibly associated with a major transgressive pulse)
caused widespread upwelling of nutrient-rich deep-ocean water, particularly along the NW
European continental margin. A consequent increase in surface productivity caused the formation of
an expanded and mtensrhed oxygen-minimum zone m the water column, and the wdesprcad
deposition of radiolarian-bearing organic-rich muds in deeper-water areas. In shelf settmgs
productivity was lower and marly nannofossil oozes (Plenus Marl facies) were deposited. At the
height of the upwelling event, the oxygen-depleted (dysaerobic) upper portion of the oxygen-
minimum zone impinged onto the Chalk Shelf, causing a widespread decline in benthonic diversity.
A fall in the rate of oceanic turnover at the end of the Cenomanian produced a waning of the
productivity event, and during the early Turonian there was a gradual return to a mid-Cenomanian-
like palaeoceanographic regime.
processes.) Neither the presence of bioturbation, nor the absence of organic-
rich sediments prove that an oxygen-minimum zone was not developed in
the water column. On the contrary, we will argue that the decline of
microfossil abundances and diversities in close association with rising 13C
values in the Plenus Marl (see Section 11.1) strongly suggests that bottom
waters became significantly depleted in oxygen (cf. Hart and Bigg, 1981;
Hart, 1985; Hart and Ball, 1986), even if not truly anoxic (Section 11.2).

10.2 Causes and consequences of the upwelling event

The combination of circumstances which caused the Cenomanian-Turonian


upwelling event remains poorly understood. An influx of nutrient-rich
oxygen-deficient bottom water from the northern South Atlantic, triggered
by the separation of Africa from South America, has been suggested by
Summerhayes (1987) as a possible cause for the event. Arthur et al. (1987)
favour a different model. It is widely held that the earliest Turonian was a
period of peak transgression, caused by a world-wide high stand of sea level
(Hancock, 1975; Hancock and Kauffman, 1979; Hart, 1980a; Haq et al.,
1987). Arthur et al. (1987) have argued that the large increase in shelf-sea
area caused by the transgression led to enhanced production of warm saline
water, which sank to form bottom-water masses. This caused an increase in
the rates of oceanic turnover because in the Cretaceous oceanic circulation
was salinity driven (cf. Brass et al., 1982). Increased circulation caused
enhanced upwelling, which triggered the OAE.
A rapid rise in sea level during the latest Cenomanian is indicated by
foraminiferal assemblage data (Hart, 1980a), and could also be at least partly
responsible for the anomalous clay-rich, locally calcarenitic, sediments
(e.g. Plenus Marl) which characterise this interval. Data from other Upper
Cretaceous successions (Jarvis and Gale, 1984) suggest that rapid rises and
falls in sea level can both produce attenuated sequences in pelagic shelf-
carbonates. Decreased rates of deposition during the latest Cenomanian are
indicated by the presence of reworked sediment, and the occurrences of
omission and erosion surfaces in the Plenus Marl and other facies. The
anomalously high clay-content of many Cenomanian-Turonian intervals
may also relate to the rapid sea level rise, being caused partly by an increased
flux of detrital non-carbonate (derived from recently inundated land areas),
and partly by the preferential removal of fine-grained carbonate during
phases of increased current activity.
The palaeoceanographic model proposed by Arthur et al. (1987) predicts,
therefore, that the Cenomanian-Turonian boundary event involved the
upwelling of warm saline bottom-water off the European continental shelf.
This conflicts, however, with available palaeobiogeographic interpretations.
Data from Tethyan macrofossil and microfossil assemblages (Kuhnt et al.,
1986) suggest that there was a temporary spread southwards of cooler water
Boreal groups associated with the OAE. In southern England, the occur-
rences of Actinocamax plenus and Oxytoma seminudum in the upper Plenus
Marl indicate the spread of cool-water north Boreal species into the area
(Jefferies, 1962, 1963a, b). These fauna1 observations, therefore, support the
development of a cold-water upwelling environment, and not the warm-
Microfossil Assemblages and the Oceanic Anoxic Event 85

water model proposed by Arthur et al. (1987). The disappearance of chorate


dinoflagellate cysts, and a temporary proliferation of proximate cysts such as
Kallosphaeridium ringnesiorum, may also relate to cooler water conditions,
since chorate cysts are generally regarded as being characteristic of warm
water-masses (Davey, 1970; Davey and Rogers, 1975).
Clearly, the dynamics of the Cenomanian-Turonian upwelling event
remain poorly understood, and no current model adequately explains all of
the characteristics of the sequences which were deposited at that time.
Nevertheless, it remains generally accepted that the event did lead to
widespread anoxia in the oceans. The impact of these palaeoceanographic
and environmental changes on the marine biota will now be considered.

11. Effects of the OAE on the Chalk Sea biota (I. J. et al.)

The proven temporal correspondence between the Cenomanian-Turonian


upwelling event (which caused the OAE) and the carbon stable-isotope
excursion, implies that the OAE was the immediate cause of increased 6i3C
values in shelf limestones. By comparing the relationship between carbon
stable-isotope curves and microfossil abundance and diversity data, there-
fore, it is possible to gauge the effects on the Chalk Sea biota of the
palaeoceanographic changes associated with the Cenomanian-Turonian
OAE.

11,l Microfossil abundance and diversity

When the specific diversity of each microfossil group is plotted stratigraphi-


tally (Figure 31), it becomes clear that there is a major decline in all groups
within and immediately above the Plenus Marl. Diversity trends vary
inversely with the carbon stable-isotope curve, diversities reaching minima
at or immediately above the peak in 613C values. Microfossil abundances are
also lowest at this level (Figure 32), although through most of the sequence
diversity is not directly controlled by variations in abundance.
A significant increase in ostracod abundance and diversity occurs at the
base of the Plenus Marl (Figures 31, 32), which is coincident with a fall in
dinoflagellate abundance and diversity. Increased foraminiferal abundances
at the same level, however, are not accompanied by changes in the diversity of
either benthonic or planktonic taxa (Figures 31, 32). The isotopic shift at
the facies boundary is minimal, and it is not considered likely that the OAE
had a significant effect at that level. Changes in abundance and diversity are
partly explained by the sedimentological characteristics of the basal Plenus
Marl, which displays strong evidence of reworking and winnowing of the
sediment (Section 2.4). Increased abundances of ostracods and foraminifera,
therefore, partly reflect the higher proportion of sand-sized bioclasts in the
sediment. Equally, winnowing may have reduced the numbers of dinoflagel-
late cysts deposited, because of their smaller size and lower density. The
specific composition and biometrics of the planktonic foraminiferal popul-
ation, however, indicate that the basal Plenus Marl was probably deposited
in deeper-water than the underlying Abbots Cliff Chalk (Hart, 1.980a; Leary,
1987).
86

I I I I
I 1

I ’
1 ,

6 13c

..,.,. .,.,.:
::::::..I:
....
,.,.
:.,.:.
.,. . ‘. .:.
., ,. .,
i; .,::,j.: Melbourn Rock
:..:.......
Beds

: ,: .::.:.:.:.:,
:..::li”iiijliliiiij’& ___
-k
,,.:: :. ../._
:.S.”
.::. .‘*.x__._
%, (PDBl ,, (-jstracods ?._-,._

‘jll
2 4 5
I
0
I
10
1
20
I
0 10
I 1
20

Isotope stratigraphy Benthonic diversity

Number of I Number of

Melbourn Rock

I’ Dmoflagellate cysts 1 Foraminifera ’

602’0d 10

Planktonic diversity
Microfossil Assemblages and the Oceanic Anoxic Event 87

The biostratigraphic changes which occur in association with the major


rise in carbon stable-isotope values at the top of Plenus Marl Bed 1 (Section
9.1), are manifest by a massive drop in benthonic foraminiferal diversity
(Figure 31). Ostracods also display a small drop in diversity at this level.
These changes are caused by the extinctions of many typical Cenomanian
species. Planktonic diversities, on the other hand, do not vary significantly
within the initial phase of the isotope excursion.
The microfossil turnover at the top of the Plenus Marl Bed 3 (typified by
the extinction of Rotalipora), which coincides with further increases in 613C
values, includes a further major fall in benthonic diversity (foraminifera and
ostracods; Figure 31), despite significant increases in abundance (Figure 32).
A sudden decline in dinoflagellate cyst abundance and diversity occurs at the
same level. The peak of the carbon stable-isotope excursion in Plenus Marl
Beds 6-8 is accompanied by a massive drop in the abundance and diversitv of
all microfossil groups (including the calcareous nannofossils but excluding
calcispheres; Figures 3 1, 32), which reach minima in the basal beds of the
Melbourn Rock.
Carbon stable-isotope values decrease significantly immediately above the
Cenomanian-Turonian boundary, and this decrease is accompanied by a
rapid recovery in specific diversity of planktonic foraminifera and calcareous
nannofossils (Figure 31), and to a lesser extent of benthonic foraminifera and
ostracods. Dinoflagellate cysts, on the other hand, are absent from the basal
Dover Chalk, only reappearing in the middle of the Melbourn Rock Beds.

11.2 Oxygenation levels and the biota

We have demonstrated that within the Cenomanian-Turonian boundary


succession at Dover there is a strong inverse correlation between 6r3C values
and microfossil diversity. Stratigraphic data (Section 10.1) indicate that the
carbon stable-isotope excursion coincides with widespread anoxia in the
ocean basins, caused by the Cenomanian-Turonian upwelling event (Section
10.2). The fall in microfossil diversity at Dover can be explained, therefore,
by declining seawater oxygenation levels, not only at the margins of the
NW European shelf and in the North Sea, but also within the central part
of the Chalk Sea.
Work on the effects of oxygenation levels on benthonic diversity (Rhoads
and Morse, 1971; Rosenberg, 1977, 1980; Andersin et al., 1978) indicate that
normal seawater oxygen levels of 2-8 ml 1 - ’ must fall to < 1 ml 1 -r before
producing a significant decline in the macrobenthos. A sharp drop in overall
benthonic diversity occurs at oxygen levels of 0.7-0.3 mll-‘, although
diverse agglutinating benthonic foraminiferal populations remain in modern
os!.gen-minimum zones with oxygen concentrations of 0.2-0.3 ml 1-r

Figure 3 1. Carbon stable-isotope stratigraphy and microfossil species diversity within the Cenomanian-
Turonian boundary succession at Dover. Discontinuous horizontal lines are lithological boundaries
(base of the Plenus Marl, base of the Melbourn Rock, top of the Melbourn Rock), the continuous
horizontal line marks the base of the Turonian. Note the decline in diversity of all groups evithin the
Plenus Marl in close association with rising 6’% values. Recoveries in diversity within the lower
Turonian arc’ more rapid in planktonic than benthonic groups.
I I 1

t Number of
613c specimens

,:::::::. :::.
.::,:: ::,,:::,
:. .A..~.~.~.
:. ::. :::. ::
::
:+,.,T...r..._ __ - - - - - - - - - __-_-______
,.,. .. ..

Melbourn Rock
Beds

I,....
:...
::.
,.:.:.
.F...
TUWllW
-------.Cenomanian

%o (PDBI .:YOstracods
I
I I
2- 4 5 - 400 0 500 1000

Isotope stratigraphy Microfossil abundance

Figure 32. Carbon stable-isotope stratigraphy and microfossil abundance within the Cenomanian-Turonian boundary succession at Dover. Symbols its in Figure 31.
Microfossil abundances are lowest immediately following the peak of the carbon stable-isotope excursion. Abundances and diversities are other\vise not clearI)
interrelated (compare Figure 31).
Microfossil Assemblages and the Oceanic ilnoxic Event 89

(Vercoutere et al., 1987). Bioturbation ceases at <0.2 ml l- ‘, and macro-


fauna is essentially eliminated at oxygen concentrations of <O.l ml 1-l.
Benthonic macrofossils occur throughout the Plenus Marl and basal
Melbourn Rock, and the entire sequence contains recognisable burrows,
although the prominance of Chondrites at some levels suggests that the
sediment itself was poorly oxygenated (cf. Bromley and Ekdale, 1984).
Clearly then, seawater oxygen levels remained about 0.2 ml l-l throughout
the OAE at Dover. Declining benthonic diversity in the upper beds of the
Plenus Marl, however, suggests that values of 0.3-0.7 ml 1-l 0, were
probably attained. It can be concluded, that neither true ‘anoxic’
(0 ml l-l 0,; Rhoads and Morse, 1971) nor ‘anaerobic’ (suboxic;
O-O.1 ml l-l 0, ,) conditions ever existed on the seafloor at Dover. The
bottom water must, therefore, have remained oxic (but probably dysaerobic;
0.1. 1 ml l-l 0, ,) throughout the OAE.
Declining benthonic diversity during the OAE is readily explained by
bottom-water oxygen levels falling below the minimum respiratory require-
ments of individual species, causing their disappearance or extinction. Our
data demonstrate that benthonic foraminifera, particularly species of Aren-
obulimina, Gavelinella, Plectina and Tritaxia, were most susceptible to
declining oxygenation, disappearing at the onset of the OAE (Plenus Marl
Bed 1 event; Figure 29). Some benthonic foraminifera, such as species of
Eggerellina, were more tolerant of oxygen depletion, disappearing only at the
height of the carbon stable-isotope excursion. Other forms, such as
Gavelinella berthelini, L zngu 1ogavelinella globosa, Text&aria chapmani and
smooth Lenticulina spp. were probably facultative anaerobes, becoming
commoner during the OAE as they proliferated due to the absence of
competition.
The ostracods display a similar decline to the benthonic foraminifera in
response to the fall in oxygen concentrations (cf. Weaver, 1981), with the
progressive annihilation of most podocopid species during the OAE (Figure
29). Modern ostracods are intolerant of oxygen levels of < 1.5 ml 1-i
(Peypouquet, 1983), and are consequently very rare in marked oxygen-
minimum zones. It is unclear, however, why the podocopids but not the
platycopids were so adversely affected by the OAE. Both groups are today,
and presumably were in the Cretaceous, benthonic (crawlers and bur-
rowers), and would therefore have experienced the same environmental
changes. We can speculate that platycopids are more tolerant of low oxygen
levels than podocopids, but we are unaware of any published data concerning
their modern representatives which might support this hypothesis. The most
obvious difference between the two groups is evinced by studies of the
appendages of their living representatives. Platycopids are filter-feeders,
while podocopids are scavengers, herbivores, or occasional carnivores
(Canon, 1933). It is not readily apparent, however, how these differences
could account for their varying responses to dysaerobic bottom conditions.
Another contrast concerns reproductive strategies; modern Cytherella and
CJ’therelloidea retain their eggs, the young passing through one or more
moults before leaving their mother’s carapace. Therefore, they are protected
from external conditions while at the most vulnerable stage in their life cycle.
Brood spaces are evident in female carapaces of Cretaceous species of
Cytherella and Cytherelloidea, and although some podocopids hav,e brood-
care, the genera found in the present study are not amongst them. It seems
possible, therefore, that podocopid species died out because they were unable
to reproduce successfully in a ‘difficult’ environment, while the platycopids,
although reduced in numbers, were able to maintain viable populations.
The changing structure and decline in diversity of planktonic forami-
niferal assemblages can also be related to seawater oxygen levels. ‘l’he fall
in diversity of planktonic species post-dates the major change in the
benthonic assemblages, indicating that surface waters remained oxygenated
longer than bottom waters. Modern planktonic foraminifera are loosely
stratified in the water column (Be, 1977), displaying groupings into
three faunas, broadly defined as ‘shallow-water’ (adult tests < SOm),
‘intermediate-water’ (mature tests 50-loom), and ‘deep-water’ (adult tests
> 1OOm) types. Individual species migrate diurnally within their preferred
depth range, as their feeding,cycles respond to changing light conditions. All
juveniles are found in surface waters shallower than 50 m, and progressively
migrate downwards to their preferred water depth during ontogeny, gameto-
genesis occurring only at the bottom of their depth range. The test
morphologies of these three faunas are distinct (Be, 1977), ranging from thin-
shelled non-keeled globular ‘globigerinid’ forms, which inhabit the shallow-
est water, to heavily calcified keeled less-inflated species which dominate
deeper water environments.
It has been suggested (Hart and Bailey, 1979; Hart, 1980b, 1985; Wonders,
1980; Caron and Homewood, 1983; Hart and Bail, 1986) that comparable
Cretaceous morphotypes probably also exhibited the same relative depth
relationships, although not necessarily in the same absolute depths. Non-
keeled genera (Hedbergella, Heterohelix), therefore, inhabited ‘shallow-
water’, incipiently keeled forms (Praeglobotruncana, Dicarinella, Helveto-
globotruncana) lived in ‘intermediate-water’, and keeled taxa (Rotalipora,
Marginotruncana) were restricted to ‘deep-water’ areas (Figure 33). The_
decline in diversity of planktonic foraminifera in response to the OAE,
therefore, can clearly be related to an expansion and intensification of the
oxygen-minimum zone in the water column. This initially caused the decline
and extinction only of ‘deep-water’ species (Rotalipora greenhornensis,
followed by R. cushmani) (Plenus Marl Bed 3 event; Figure 34A, B) which
were unable to reproduce as deeper waters became oxygen deficient.
‘Intermediate-water’ species, particularly the dicarinellids, temporarily
diversified to fill niches left vacant by the extinction of Rotalipora (Figure
34C), but soon afterwards they too disappeared as the oxygen-minimum
zone rose further in the water column. Finally, at the height of the OAE
(Plenus Marl Beds 6-8 and the basal Melbourn Rock; Figure 34D), the
planktonic assemblage was almost entirely restricted to ‘shallow-water’
hedbergellids.
The fall in diversity of the microflora associated with the OAE is less easily
understood. Dinoflagellate cysts were the first component of the microflora
to be effected by the OAE, displaying a rapid decline in diversity at and
immediately following the Plenus Marl Bed 3 event, and disappearing
entirely within the basal Melbourn Rock. Calcispheres, on the other hand,
reached an acme at the peak of the isotope excursion. Comparable observ-
Microfossil Assemblages and the Oceanic Anoxic Event 91

sea level

hedbergellids praeglobotruncanids rotaliporids

thin-walled, variable
trochospire, small
number randomly
placed pustules

\
periphery
\ ; \ / 4
some
I forms evolve \ thick-walled, biconvex
I with weakly to plano-convex,
I developed .I supplementary apertures,
two keels I thick well-developed
, ‘early dicarinellids’ 1 keels which thin
towards last-formed
\ / chambers
\ /
. -_ /

depth

Figure 33. Relatiohships between test morphology and water depths inhabited by Cenomanian-
Turonian planktonic foraminifera. Migration within the water column during ontogeny is shown
schematically. Juveniles of all groups inhabited shallow water. Individuals migrated into deeper
water during ontogeny, reproducing only at the lower limit of their depth range. Deeper water forms
were, therefore, more susceptible to oxygen-depletion within deep-water masses.
92 I. Jar\ is rr rd.

Plenus Marl Bed 2

Sea level

Upper Abbots Cliff Chalk B Plenus Marl Bed


Figure 34. Microfaunal diversity and oxygenation in the Chalk Sea. (A) Chalk microfaunas reached an
acme during the late Cenomanian, as increasing water depths led to increasingly favourable
conditions for both benthonic and planktonic groups. (B) A rapid development of dysaerobic bottom
waters (associated with a late Cenomanian productivity event in continental margin areas) occurred
immediately following the deposition of Plenus Marl Bed 1. Oxygen depletion caused the
disappearance of many typical Cenomanian benthonic taxa, and the extinction of Rotalipora
greenhornensis. (C) Continuing intensification and expansion of the oxygen-minimum zcme in the
Microfossil Assemblages and the Oceanic Anoxic Event 93

Sea level
J-Jr

‘. ” :, .’ ‘.”

_ C,‘thsre,,s s,.,,,. - -‘*d - - - - - - - - -- - - - - - Mara.&elf, s,J,,- - - - -


_-__-____--------- ___ _ _ - - --- --
_______________-----------------

_-__-__-_---------------------

____-_____---- ____~__----------

_______--------------

___-.___-__ ----- ~-~-------


__________.___---- ---- ------

Plenus Marl Beds 6-8 8 basal Melbourn Rock

Plenus Marl Bed 4


shelf water-column caused further declines in benthonic diversity and, by the time Plenus Marl Bed 4
was deposited, the extinction of Rotalipora cushmoni. (D) At the height of the anoxic event, the Chalk
Sea microfauna was restricted to a few species of low-oxygen tolerant benthonic foraminifera and
ostracods, and impoverished planktonic foraminiferal populations dominated by hedbergellids.
Contraction and weakening of the oxygen-minimum zone at the end of the Cenomanian allowed
the development of new ‘Turonian’ groups which gradually evolved to refill the niches left vacant by
the Cenomanian extinctions.
ations, including a rarity of dinoflagellate cysts in black shale facies (Tyson,
1987) and a proliferation of calcispheres during the OAE (Kuhnt et al., 1986;
Arthur et al., 1987) have both been noted in Cenomanian-Turonian suc-
cessions elsewhere.
It appears, therefore, that dinoflagellate cyst abundance is not related
simply to productivity (cf. Davey, 1971; Davey and Rodgers, 1975). Motile
dinoflagellates live in stably stratified water columns, and only produce cysts
during phases of hydrodynamic instability (Landry, 1977). The rarity of
dinoflagellate cysts in black shales, therefore, may have been caused by the
establishment of a stable hydrodynamic regime leading to the proliferation of
dinoflagellates, but mediating against cyst formation (cf. Tyson, 1987).
Calcispheres, on the other hand, appear to have been an opportunistic group,
their abundance probably reflecting an increased nutrient supply and a
decline in the zooplankton.
The calcareous nannofossil diversity curve is comparable to, but slightly
offset from, that for planktonic foraminifera, starting to fall in Plenus Marl
Beds 6-8 but reaching a minimum above the base of the Melbourn Rock. In
the absence of nannofossil abundance data, we cannot demonstrate whether
or not this apparent decline relates to falling abundances. A drop in
nannofossil abundance during the OAE might be expected, since nannofossil
deposition rates are dependent on zooplankton (chiefly copepod) faecal pellet
production (e.g. Degens and Ittekkot, 1987). The decline of zooplankton
during the OAE might, therefore, have caused decreased rates of nannofossil
deposition. Alternatively, falling nannofossil abundance and diversity dur-
ing the later phases of the OAE may reflect a decline in the supply of
nutrients, caused by extensive black shale deposition which permanently
removed nutrients from the oceanic reservoir. A fall in nutrients following
the peak of the OAE would explain why nannofossils declined later than the
other microfossil groups.

11.3. Expansion of the oxygen-minimum zone into the Chalk Sea

It is postulated (cf. Arthur et al., 1987) that the oxygen-minimum zone


occupied the water column from a minimum depth of 100-200 m to a
maximum depth of 1.552.5 km at the height of the OAE (Figure 30). We
interpret the decline in benthonic diversity of the Plenus Marl at Dover as
indicating falling bottom-water oxygenation levels, suggesting that the
upper limits of the oxygen-minimum zone (which by analogy to modern
systems was probably characterised by oxygenation levels of < 0.5 ml l- ’ e.g.
Vercoutere et al., 1987) impinged on the seafloor at Dover during or
immediately after the deposition of Plenus Marl Bed 1 (Figure 34B), causing
the extinctions which occur at that level. The continuing decline of the
microbenthos through the remainder of the Plenus Marl and basal Dover
Chalk, suggests that oxygen levels fell further throughout the latest
Cenomanian. At the same time, the oxygen-minimum zone expanded
through the water column (Figure 34) causing the progressive extinctions of
‘deep water’ and then the disappearances of ‘intermediate water’ species of
planktonic foraminifera.
Based on these data, the oxygen-minimum zone was at its most intense and
Microfossil Assemblages and the Oceanic Anoxic Event 95

reached its shallowest depth at the peak of the isotope excursion (Figure
34D), immediately before the start of the Turonian, during the deposition of
Plenus Marl Beds 6-8 and the basal metre of the Melbourn Rock. Calcareous
microfossil tests display evidence of seafloor dissolution within this interval,
a feature which is also characteristic of tests collected from modern oxygen-
minimum zones (Penrose and Kennett, 1979). The occurrence of only
‘shallow-water’ planktonic foraminifera at this time suggests that the
oxygen-minimum zone may have risen to depths as shallow as 50 m (cf. Bk,
1977). The recovery of microfossil diversity associated with the stage
boundary indicates the waning and contraction of the oxygen-minimum
zone during the earliest Turonian.

11.4 Extinction, evolution and the OAE

The recognition of extinctions as opposed to disappearances within the


succession requires a geographically and stratigraphically more extensive
database than that presented here. We consider, however, that many of the
disappearances within the Cenomanian-Turonian at Dover were caused by
regional and probably world-wide palaeoceanographic changes. If this is the
case, then the disappearances recorded at Dover are also probably extinction
levels for many species. This is confirmed by previous work (Hart, 1980b,
1985; Hart and Bigg, 1981; Hart and Ball, 1986; Hilbrecht, 1986; Kuhnt et
al., 1986) for some planktonic foraminiferal species such as Rotalipora
greenhornensis and R. cushmani. Even if it is assumed that only a small
number of the disappearances at Dover are true extinction levels, the OAE
must still have caused a dramatic change in the population structure of many
fossil groups.
The association of the Cenomanian-Turonian OAE with a stage boundary
is not coincidental since, we would argue, it was the OAE which caused the
major biotic turnover that defines that boundary. The OAE was responsible,
therefore, for the progressive and rapid extinction of many Cenomanian
species, which disappeared as their habitats became uninhabitable. The base
of the Turonian is defined by the evolution of new macrofossil species, but
many new microfossils also appear around the same level. Diversity data,
however, indicate that new species (particularly of benthonic forms) evolved
only gradually following the OAE, and diversities failed to recover to
Cenomanian numbers during the early Turonian.
The OAE model provides, therefore, a mechanism for marine extinction
e\.ents, without the need to invoke extraterrestrial causes, so 0,4Es may have
exerted an influence on rates of extinction and evolution throughout the
Phanerozoic.

12. Conclusions (I.J. et al.)

The onset of the Cenomanian-Turonian Oceanic Anoxic Event (OAE)


coincided with a major eustatic transgressive pulse during the late
Cenomanian. The OAE was initiated by an upwelling event, which caused a
widespread increase in ocean-surface productivity. Increased productivity
led to an expanded and intensified oxygen-minimum zone in the water
column, particularly in ocean-margin settings. Deposition of laminated
organic-rich muds became widespread (including parts of the NE Atlantic
continental margin and the North Sea), particularly in areas where
anaerobic-anoxic bottom waters were in contact with the seafloor. Progress-
ive burial of increasing amounts of marine organic matter caused enhanced
6l 3C concentrations in seawater and in coeval carbonate sediments. Upwel-
ling declined at the end of the Cenomanian, causing a gradual contraction of
the oxygen-minimum zone, and a return to normal marine deposition by the
early Turonian.
In epicontinental seas, continual deposition of fossiliferous bioturbated
organic-lean sediments during the OAE, indicates that in these areas bottom
waters never became completely anoxic. Major declines in benthonic and
planktonic diversity (which are directly associated with the OAE) suggest,
however, that penetration of poorly oxygenated oceanic water from the
North Atlantic, gradually led to dysaerobic conditions in the Chalk Sea.
Late Cenomanian sediments throughout southern England contain di-
verse benthonic and planktonic fossil assemblages which had evolved
gradually in response to rising sea-levels throughout the Cenomanian. A
further rapid increase in water depth (associated with increased current
activity) occurred immediately prior to, and during, the deposition of Plenus
Marl Bed 1, which contains a typical late Cenomanian biota.
The onset of dysaerobic bottom conditions took place immediately
following the deposition of Plenus Marl Bed 1, causing a major fauna1
turnover at that time. Microbenthonic diversity decreased suddenly during
the Bed 1 event, which led to the extinctions of many typically Cenomanian
benthonic taxa, including echinoids (Holaster trecensis), foraminifera
(species of Arenobulimina, Gavelinella, Plectina, Tritaxia), and ostracods
(several species of podocopid). The extension of low oxygen concentrations
above the seafloor is indicated by the synchronous extinction of the deep-
water planktonic foraminifera Rotalipora greenhornensis.
Bottom water oxygenation levels continued to decline during the depo-
sition of Plenus Marl Beds 223, causing a further fall in microbenthonic
(particularly foraminiferal) diversity. Anomalously diverse macrofaunal
assemblages associated with this interval, are a product of abnormal
diagenetic conditions, which favoured the preservation of aragonitic fauna.
Following the deposition of Bed 3, oxygen depletion in the bottom waters
reached a critical threshold, causing an accelerated decline in the benthonic
diversity for all groups, but particularly the podocopid ostracods (e.g.
Bairdoppilata spp., Herrigocythere donzei). The extension of dysaerobic
conditions into shallower parts of the water column is indicated by the
extinctions of the ammonite Calycoceras naviculare and the deep-water
planktonic foraminifera Rotalipora cushmani, although the presence of well-
oxygenated intermediate waters is indicated by a temporary proliferation of
early dicarinellid planktonic foraminifera. At the same time, a change in
water-mass distribution is indicated by the temporary appearance of north
Boreal macrofauna, including Actinocamax plenus and Oxytoma seminudum,
which are restricted to the upper part of the Plenus Marl (Beds 4-6).
Oxygen levels continued to fall during the deposition of upper Plenus Marl
Beds 668, probably reaching a minimum towards the end of this interval. By
Microfossil Assemblages and the Oceanic Anovic Event 97

this time most typical Cenomanian species had disappeared, and the
microbenthos was of low diversity, restricted to a few species of foraminifera
(Gavelinella berthelini, Lingulogavelinella globosa, Textularia chapmani,
smooth Lenticulina spp.) and platycopid ostracods (mainly Cytherella ovata).
Oxygen-depleted waters had now risen to their shallowest depth (possibly to
within 50m of the surface), leaving only reduced populations of shallow-
water hedbergellid planktonic foraminifera. All calcareous microfossils
display surface etching of their tests in these beds, indicating seafloor
dissolution of carbonate, comparable to that seen in modern oxygen-
minimum zones. Dinoflagellate cysts also display a major decline in diversity
at this time, and disappear entirely immediately following the deposition of
the Plenus Marl. Their decline is initially accompanied by a sudden
abundance of Kallosphaeridium ringnesiorum, and this is followed by a
proliferation of calcispheres, which constitute the main component of the
sediment at the height of the OAE. These changes may relate to the
establishment of abnormally stable hydrodynamic conditions at this time.
Oxygenation levels remained low during the deposition of basal beds of the
Melbourn Rock which contain faunas that are as impoverished as those in the
uppermost Plenus Marl. A decline in calcareous nannofossil diversity at the
end of the Cenomanian, however, may indicate falling nutrient levels
associated with the waning of the OAE. An initial phase of better oxygenation
immediately preceding the start of the Turonian (marked by the appearance
of Mytiloides spp. inoceramid bivalves), is evidenced by increases in both
benthonic and planktonic microfossil diversity.
The base of the Turonian is defined by the appearance of new species of
inoceramid bivalves and ammonites which were associated with increasing
oxygen levels in the Chalk Sea. New species appeared in all groups during the
early Turonian, as an increased number of habitats became reinhabitable,
and evolution filled the niches which had been left vacant by the late
CenoTanian extinctions. Appearances included new species of Areno-
bulimirza, Lingulogavelinella, and Tritaxia in the benthonic foraminifera,
and new species of podocopid ostracods (e.g. C)lthereis sp. cf. C. longaeeja,
Mosaeleberis sp. A, Curfsina senior). Planktonic foraminifera evolved new
intermediateiwater (Helvetoglobotruncana) and then deep-water (Margino-
truncana) genera, and several new species of calcareous nannofossil and
dinoflagellate cyst eventually appeared (including the biostratigraphically
important forms Q.uadrum gartneri and Senoniasphaera rotundata). Plank-
tonic groups generally recovered more rapidly than benthonic taxa, although
dinoflagellate cysts remained absent for a considerable interval of time.
Diversities never regained Cenomanian levels during the early Turonian,
despite evidence of continued increases in regional water depth. This
demonstrates the lasting impact of the OAE on the marine ecosystem.
We believe, therefore, that the Cenomanian-Turonian OAE caused a
sequence of extinctions during the latest Cenomanian. The recovery of the
biota following the event led to the evolution of the new species which define
the Turonian, enabling it to be distinguished from the under1yir.g
Cenomanian. Oceanic anoxic events provide, therefore, a mechanism ior
extinction and evolution in the marine realm which was almost certainly
operative throughout the Phanerozoic.
Acknowledgments

All authors wish to acknowledge the support of their respective institutions.


Fieldwork funding was partly provided by a National Advisory Body (NAB)
research grant to I. Jar\%. ‘I’he work was undertaken while G. A. Carson and
P. N. Leary were in receipt of Natural Environment Research Council
(XERC) research studentships. Stable isotope laboratory facilities in the
Department of Earth Sciences, Liverpool University were funded by NERC
Grant GR3/5339 (to Dr. J. D. Marshall) and Liverpool University. Process-
ing of dinoflagellate cyst samples was undertaken by British Petroleum
(Sunbury). Technical assistance was provided by Hilary Attenbourgh, Rosie
Gent, Brian McIlroy, Graham Mott and Linda Parry. D. Horne wishes to
thank I’. C. Mitchell of the Surrey Police, in whose safe custody the ostracod
faunas from this study were briefly held follo\ving a motor accident on the A3
near Guildford. Drs. Jim Marshall, Chris Paul and two anonymous referees
critically reviewed the manuscript, and suggested many notable improve-
ments which were incorporated into the final paper. Paula ‘l’horne-Jones
expertly manipulated the manuscript through its various forms on the kvord-
processor.

References

.-lndersin, 4.-B., Lassig, I,., Parkkonen, L. & Sandler, H. 197X. The decline of macrofauna in the deeper
parts of the Baltic proper. Kieler meeresforschungen Sonderheft, 4, 23 52.
Anderson, T. F. 8r Arthur, Xl. A. 1983. Stable isotopes of oxygen and carbon and their application
to sedimentologic and paleoenvironmental problems. In Stable Isotopes in Sedimentary Geology
(M. A. Arthur, T. F. Anderson, I. R. Kaplan, J. Veizer & L,. S. Land). Society of Emnomic Palaeon-
tologists and Mineralogists Short Course 10, l-1 -1- 15 1.
Arthur, Xi. .4., Schlangcr, S. 0. & Jrnkyns, 11. C. 1987. ‘l’hr Crnl)manian-‘I’uronlian Ocranlc Anox~c
E\rnt, II: palaeoceanographic controls on organic matter production and preservation. In fii’arine
I’etvolrwn Source Rocks (Eds J. B rooks 8r A. Fleet) Geological Society’ Special Publiration 26,401 -!20.
Bahinot, J. F., Cohn, J. I’. & Damotte, R. 1985. Ostracodes du Cretacb sup&rieure. In Atlas des
Ostracodrs de France (Ed. I-I. J. Oertli). &t//et& des Criltrrs de Recherchrs d’E.xp/orrrtion Production
E!f-Aquitnine, 9, 2 11 215.
Barrois, C. 1876. R&cherches sur le terrain C&ace suphrieur de I’Angleterre et de I’lrelande. ,V&zoires de
In SociPte pPologique ht AY’nrd, 1, 232 pp.
BG. A. W;. H. 1977.4n ecological. zoogeographic and taxonomic review of recent planktonic foraminlfera.
In Oceanic Micropalaeontology (Ed. A. T. S. R amsay) London: Academic Press. 1 100.
Below, R. 1981. Dinoflagellaten zysten aus den oberen Hauterive bis I:nteren Cenoman, Stid-West-
Marokkos. Palaeontographica, Abt. B. 176, 1 -1JS.
Below, R. 1982. Scolochorate zysten der Gonyaulacaccan (Dinophyceae) aus drr Untrrkwide Zlarokkos.
Palaeontographica Abt. B 182, lL51.
Birkelund, T., Hancock. J, M.. Hart, ,\l. B.. Ra\vson, P. F.. Remane, J. Robaszynski, I’., Schmid, F. &
Surlyk, F. 1984. Cretaceous stage boundaries: proposals. Bulletin of the Geological Societ_v of
Denmark, 33, 3-20.
Brass, C. W,:., Southam, J. R. & Peterson. W. 11. 1982. Warm saline bottom water m the ancient ocean.
.Vature. 296, 620-623.
Bromley. R. G. & Ekdale, A. A 1984. Chondrites: a trace fossil indicator of anoxia in sediments. Science.
224,872-X74.
Burnhill, T. J. & Ramsay, M. V. 1981. Mid-Cretaceous palaeontolqg and stratigraphy. central North
Sea. In Petroleum Geology of the Continental Shelf of iZTorth West Europe (Eds L. \‘. llling &
G. D. Hobson) London: Institute of Petr&um, 245-254.
Cal\wt, S. E. 1987. Oceanographic controls on the accumulation of organic matter in marine sediments.
In Marine Petroleum Source Rocks (Eds J. Brooks Sr A. J Fleet) Geological Society Special Publication,
26, 137-151.
Canon. H. G. 1933. On the feeding mechanism of certain marine ostracods. Trawsnrtions of the Royal
Society of Edinburgh, 51, 739 761.
Microfossil Assemblages and the Oceanic Anoxic Event 99

Caron, M. & Homewood, P. 1983. Evolution of early planktonic foraminifers. Marine Micropaleontology
I, 453462.
Carter, D. J. & Hart, M. B. 1977. Aspects of mid-Cretaceous stratigraphical micropalaeontology. Bulletin
of the British Museum of Natural History (Geology) 29, 1-135.
Clarke, B. 1982. Die Gattung Cq'ltherelloidea Alexander, 1929 (Ostracoda) in Schreibekreide: Richtprofil
van Laperdorf-kronsmoor-Hemmoor (NW-Deurschland). Geologisches fahrbuch A61,35-71
Clarke, B. 1983. Die Cytheracea (Ostracoda) im Schreibekreide-Richtprofil van LHgerdorf-
Kronsmoor-Hemmoor (Conaic bis Maastricht; Norddeutschland). Mitteilungen der Geologischen
und Paliiontologischen Institut der Universitat Hamburg. 54, 65--168.
Clarke, R. F. A. & Verdier, J.-P. 1967. An investigation of microplankton assemblages from the Chalk of
the Isle of Wight. Verhandelingen der konninklijke Nederlandse Akademie ran Wetenschapen, Ajdeeling
Natururkunde, Eerste Reeks, 24, l-96.
Clayton, C. J. 1986. The chemical environment of flint formation in Upper Cretaceous chalks. In The
Srientijic Study of Flint and Chert (Eds G. de G. Sieveking & NI. B. Hart). Cambridge: Cambridge
University Press, 43-54.
Cobban, W. A. 1984. Mid-Cretaceous ammonite zones, Western Interior, United States. Bulletin of the
Geological Society of Denmark, 33, 71-89.
Craig, H. 1957. Isotopic standards for carbon and oxygen and correction factors for mass-spectrometric
analysis of carbon-dioxide. Geochimica et Cosmochimica Acta, 12, 133-149.
Crux, J. A. 1982. Upper Cretaceous (Cenomanian to Campanian) calcareous nannofossils. In A
Stratigraphical Index of Calcareous Nannofossils (Ed. A. R. Lord) Ellis Horwood for the British
Micropalaeontological Society, 81-135.
Cunningham, R. & Kroopnick, P. iX’I. 1985. Inorganic and isotopic geochemistry of sediments from Sites
549 to 551, northeastern Atlantic. Initial Reports of the Deep Sea Drilling Project, 80, 1073-1079.
Davey, R. J. 1969. Non-calcareous microplankton from the Cenomanian of England, northern France
and North America. Part 1.Bulletin of the BritishMuseum &Natural History (Geology i, 17,103~180.
Davey. R. J. 1970. Non-calcareous microplankton from the Cenomanian of England, northern France
and North America, Part II. Bulletin of the British Museum of Natural History lGeologyi, 18,
333-397.
Davey, R. J. 1971. Palynological and palaeoenvironmental studies, with special reference to the
continental shelf of South Africa. In Proceedings of theSecolld Planktonic Conference, Rome 1970 (Ed.
A. Farinacci) 1, Edizioni Teconoscienza Roma, 331-347.
Dave?;, R. J. & Rogers, J. 1975. Palynomorph distribution m recent offshore sediments along two
traverses off S.W. Africa. Marine Geology, 18, 213-225.
Deegan, C. E. & Scull, B. J. (Eds.) 1977. A standard lithostratigraphic nomenclature for the Central and
Northern North Sea. Report of the Institute of Geological Sciences, 77125, 36 pp.
Deflandre, G. 1935. ConsidCrations biologique sur les microorganismes d’origine planctonique conserv&s
dans les silex de la Craie. Bulletin biologique de la France et de la Belgique, 69, 213-244.
Degens, E. T. & Ittekkot, V. 1987. The carbon cycle-tracking the path of organic particles from sea to
sediment. In Marine Petroleum Source Rocks (Eds J. Brooks &A. J. Fleet). Geological Society Special
Publication, 26, 121-135.
Dowker, G. 1870. On the Chalk of Thanet, Kent, and its connection wsith the Chalk of east Kent.
Geological Magazine, 7, 466-472.
Emrich, K., Enhalt, D. H. & Vogel, J. C. 1970. Carbon isotope fractionation during the precipitation of
calcium carbonate. Earth and Planetary Science Letters, 8, 363-371.
Ernst, G., Wood, C. J. & Hilbrecht, H. 1984. The Cenomanian-Turonian boundary problem in NW
Germany with comments on the north-south correlation to the Regensburg area. Bulletin of rhe
Geological Society of Denmark, 33, 103-l 13.
Foucher, J.-C. 1979. Distribution stratigraphique des kystes de dinoflagellks et des acritarches dans le
C&a& suptrieure du Bassin de Paris et de 1’Europe septentrionale. Palaeontographica Abt. B 169,
78-105.
Foucher, J.-C. 1981. Kystes de dinoflagell& du C&a& moyen europeen: proposition d’une echelle
biostratigraphique pour le domain nord-occidental. Cretaceous Research, 2, 331-338.
Carwon, R. E. & Kennedy, W. J. 1977. Origin of solution seams and flaser structures in Upper
Cretaceous chalks of southern England. Sedimentary Geolog_y, 19, 107- 137.
Graciansky, P. C. de, Deroo, G., Herbin, J. P., Jacquin, T., Magni, F., Montadert, L. & Miiller, C. 1986.
Ocean-wide stagnation episodes in the Late Cretaceous. Geologische Rundschau, 75, 1741.
Graclansky, P. C. de, Brosse, E., Demo, G., Herbin, J.-P., Montadert, L., Mtiller, C., Sigal, J. &
Schaaf, A. 1987. Organic-rich sediments and palaeoenvironmental reconstructions of the Cretaceous
North Atlantic. In Marine Petroleum Source Rocks (Eds J. B rooks & A. J. Fleet). Geological Societ?,
Special Publication 26, 3 17-344.
Hancock, J. M. 1975. The petrology of the Chalk. Proceedings of the Geologists’ Association, 86,499-535.
Hancock, J. M. 1984. Some possible boundary-stratotypes for the base of the Cenomanian and Turonian
Stages. Bulletin of the Geological Society of Denmark. 33, 123-128.
100 I. Jarvls et al.

Hancock, J. h,l. 1986. Cretacrous. In Introductron to the Petroleum Geo1og.v of the A’orth SPU
(Ed. K. W. Glennie), Second Edition. Oxford: Blackwell, 161 178.
Hancock, J. M. & Kauffman. E. C;. 1979. The great transgressions of the Late Cretaceous. Jownal of the
Geological Societ_v, 136, 175 186.
Haq, B. Il.* Hardenbol, J, & Vail. P. R. 1987. Chronology of fluctuating sea levels since the Triassic.
Science, 235, 1156-l 167.
Harland. m”. B., Cos, A. r., Llewellyn, I’. G., Pickton, C. .4. G., Smith, A. G. & Walters, R. 1982. il
Geologic Time Scale. Cambridge: Cambridge University Press, 13 1 pp.
Hart. M. B. 19800. The recognition of mid-Cretaceous sea-level changes by means of foraminifera.
Cretaceous Research. 1, 289-297.
Hart, >l. B. 1YXOb. A water depth model for the evolution of the planktomc Foraminiferida. Nature, 286,
252 -254.
liart, .\I. H. lY82. ‘I’uronian forammiferal biostratigraphy of southern England. In (‘olloque SUY le
?icronirn, Mtknoires du Mzmkm National d’Histoire Natuvelfe SPrie C, 49, 203-207.
Hart, Xl. B. 1985. Oceanic anoxic event 2 on-shore and off-shore SW England. Proceedings ofthe Ussher
Society, 6, 1X3-190.
IIart, X1. B. & Bailey, H. W. 1979. The distribution of planktonic Foraminiferida in the mid-Cretaceous
of NW Europe. In Asp&e der Kreide Europas (Ed. J. Wiedmann). IUGS Series A 6, 527-542.
Hart, >I. B. & Ball, K. C., 1986. Late Cretaceous anoxic events, sea-level changes and the evolution of
the planktonic foraminifera. In North Atlantic Palaeoceanography (Eds. C. P. Summerhayes and
N. J. Shackleton). Geological Society Special Publication 21, 67~ 78.
Hart, hI. B. & Bigg, I’. J, 1981, An&c events in the late Cretacrous chalk seas of north-west Europe. In
Rlicrofossilsfrom Recent and Fossil Shelf&as (Eds J. W. Nealr and M. D. Brasier). Ellis Horwood for
the British ~Iicropalarontolc,Rlcal Society, 177 1X5.
Hart, 11. I~. & Swiecicki, .A. 10x7. Foraminifera of the chalk facirs. In :I~icropn/aeo,rfol~/g~ of (‘rrrbonatr
Ewironments (Ed. hl. B. Hart). Ellis H orwood British Micropalaeontological Society Series.
121 137.
Hart, M. B.. Bailey, H. W., Fletcher. B.. Price, R. & Sweicicki, .A. 1981. Cretaceous. In Stmtigraphzcul
.qtlas of Fossil Foraminifpra (Eds. D. G. Jenkyns and J. \V. Murray). Ellis Horwood British
Micropalaeontological Society Series, 149~ 227.
Hebert. E. 1874. Comparaison de la Craie des Cotes d’Anglrterrr avcc celle de France. Bulletin de la
Sorit;ti gPologique de France 3e Se%. 2, 416642Y.
Hrrbin, J. P., Montadert, L., XIuller, C., Gomez, R., Thurow, J. 8r q’iedmann. J, 1986. Organic-rich
sedimentation at the Cenomanian-Turonian boundary in oceanic and coastal basins in the North
Atlantic and ‘Iethys. In North Atlantic Palaeoceanography (Eds C. P. Summerhayes &
N. J. Shackleton). Geological Society Special Publication, 21, 3W 422, Oxford: Blackwell.
Hilbrecht, H. 1986. On the correlation of the cpper Cenomanian and Lower Turonian of England and
Germang (Boreal and N-Tethys). on Stratigraphy, 15, 115-138.
:Vrzus/ettcw
Hilbrecht, H. & Hoefs, J. 19X6. Geochemical and palaeontological studies of the 6i3C anomaly in Boreal
and North Tcthyan Cenomanian-Turonian sediments in Germany and adjacent areas. Palneogeo-
graphs Palaeoclimatology artd Palaeoecology, 53, 169-l 89.
Hill, W. 1886. On the beds between the Upper and Lower Chalk of Dover, and their comparison with the
RIiddle Chalk of Cambridgeshire. Quavterl>,Jourrralof the Geological Society of London. 42,232~246.
Hill, W. & Jukes-Brownc. A. J. 1886. The Melbourn Rock and Zone of Belemnitella plena from
Cambridge to the Chiltern Hills. QuwtwIy Journal of the Geological So&v of London, 42, 216.~231.
Hudson, J. D. 1977. Stable isotopes and limestone lithification. Journal of the Geological Society,
133,637 660.
Irwin, H., Coleman, M. L. & Curtis, C. I>. lY77. Isotopic evidence for source of diagenetic carbonates
formed during burial of organic-rich sediments, Nature, 269, 2099213.
Jarvis, I. & Gale, A. S. 1984. The late Cretaccous transgression in the SW Anglo-Paris Basin:
Stratigraphy of the Crair de Villedieu Formation. C’rrtaceous Research, 5, 195 224.
Jar\is, I. & ‘I‘ocher. II. A. lYX3. ‘I’hr Crnomanian-‘I’uronlan boundary in SE I>evon, England.
In C’retcweous Stage Boundaries Sywposiwn .-lbstractr (Eds T. Birkelund, R. G. Bromley,
\V. Ii. Christensen, E. Hikansson & F. Surlyk). Copenhagen: University of Copenhagen, 94-97.
Jarvis, 1. h ‘I’ochcr, B. A. 1087. Field Meetinp: the Cretaceous of SE Devon. l+ 16th March, 10X0.
Proceedings of the Geologists‘ .-lssociation. 98, 5 1 -66.
Jarvis. I. S: Woodroof, P. B. 1084. Stratigraph? of the Ccnomanian and basal Turonian (Upper
Cretaceous) between Branscomhe and Seaton, SE De\ on. England. Proceedings of the Geologists’
.-lssociation, 95, 193 215.
Jarvis. I., Lear?, I’. N. & Tochrr. H. A. 1987. 1lid Creraceous (Albian-Turonian) stratigraphy of
Shapwick Grange Quarry, SE Devon, England. Mesozoic Research, 1, 119-134.
Jarvis, I.. Carson, G.. Hart, .\,I. B., Lcwy. I’. r\. & Tocher, B. A. 1988. The Cenomanian-Turonian (late
Crrtaceous) anosic event in SW England: evidence from Hooken Cliffs near Beer. S.E. Devon.
,Vrws/rtterscmStrotigrcrpl~j,. 18, 147 164.
Microfossil Assemblages and the Oceanic Anoxic Event 101

Jefferies, R. P. S. 1962. The palaeoecology of the Actinocamax plenus subzone (lowest Turonian) in the
Anglo-Paris Basin. Palaeontology, 4, 6099647.
Jefferies, R. P. S. 19630. The stratigraphy of the Actinocamax plenus subzone (Turonian) in the Anglo-
Paris Basin. Proceedings of the Geologists’ Association, 74, l-33.
Jrfferies, R. P. S. 19636. Fauna and environment in the Lowest Turonian (Actinocamarplenus Subzone)
of the Anglo-Paris Basin. In Problems in Palaeociimatology (Ed. A. E. M. Nairn). London: Wiley.
600-608.
Jenkyns, H. C. 1980. Cretaceous anoxic events: from continents to oceans. Journal of the Geological
Society, 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. & Clayton, C. J. 1986. Black shales and carbon isotopes in pelagic sediments from the
Tethyan Lower Jurassic. Sedimentology, 33, 877106.
Jsrgensen, N. 0. 1987. Oxygen and carbon isotope compositions of Upper Cretaceous chalk from the
Danish sub-basin and the North Sea Central Graben. Sedimentology, 34, 559~570.
Jukes-Browne, A. J. 1880. The subdivisions of chalk. Geological Magazine, 7, 2488257.
Jukes-Browne, A. J. &Hill, W. 1903. The Cretaceous Rocks of Britain Volume 2: The Lower and Middle
Chalk of England. Memoir of the Geological Society of the United Kingdom. London: H.M.S.O.,
568 pp.
Kennedy, W. J. 1969. The correlation of the Lower Chalk of south-east England. Proceedings of the
Geologists’ Association, 81, 6133677.
Kennedy, W. J. 1984. Ammonite faunas and the ‘standard zones’ of the Cenomanian to Maastrichtian
Stages in their type areas, with some proposals for the definition of the stage boundaries by
ammonites. Bulletin of the Geological Society of Denmark, 33, 1477161.
Kennedy, W. J., Wright, C. W. & Hancock, J. M. 1983. Ammonite zonation and correlation of the
uppermost Cenomanian and Turonian of southern England and the type areas of Sarthe and
Touraine in France. Me’moires du Museum National d’Histoire Naturelle Paris Serie C, 49, 175-181.
Keupp, H. 1979. Lower Cretaceous Calcisphaerulida and their relationship to calcareous dinoflagellate
cysts. Bulletin des Centres de Recherches d’Exploration-Production Elf-Aquitaine, 3, 651-663.
Kuhnt, W., Thurow, J., Wiedmann, J. & Herbin, J. P. 1986. Oceanic anoxic conditions around the
Cenomanian/Turonian boundary and the response of the biota. In Biogeochemistry of Black Shales
(Eds E. T. Degens, P. A. Meyers and S. C. Brassell). Mitteilungen aus dem Geologisch-
Palaontologischen Institut der Uniz’ersitat Hamburg SCOPE/ UNEP Sonderband Heft, 60, 2055246.
Landry, M. R. 1977. A review of important concepts in the trophic organisation of pelagic ecosystems.
Helgoldnder wissenschaftliche Meeresuntersuchunge, 30, 8-l 7.
Leary, P. N. 1987. The late Cenomanian Oceanic Anoxic Event: implications for foraminiferal evolution.
Ph.D. thesis. CNAA, Plymouth Polytechnic.
Letolle, R. & Pomerol, B. 1980. Mix en &idence dans les craies du Bassin de Paris d’un accident dans la
&partition du 13C d’dge CCnomanien Terminal. Compte rendus des Seances de I’Academie des Sciences,
Paris, 291, Dl33-136.
McCrae, J. M. 1950. On the isotopic chemistry of carbonates and a paleotemperature scale. Journal of
Chemistry and Physics, 18, 849-857.
Mortimer, C. P. 1987. Upper Cretaceous calcareous nannofossil biostratigraphy of the southern
Norwegian and Danish North Sea area. In Proceedings of the International Nannoplankton
>4ssociation : Vienna Meeting 1985, Abhandlungen der Geologischen Bundesanstalt, 39, 143-l 47.
>2lortimore, R. N. 1986. Stratigraphy of the Upper Cretaceous White Chalk of Sussex. Proceedings of the
Geologists’ Association, 97, 97 139.
Neale, J. W. 1978. The Cretaceous. In A Stratigraphical Index of British Ostracoda (Ed. R. Bate &
E. Robinson). Geological Journal Special Issue, 8, 325-384.
Owen, M. 1970. Turonian foraminifera from southern England. Ph.D. Thesis. University of London.
Penrose, N. L. & Kennet, J. P. 1979. Anoxic and aerobic basins in the northern Gulf of Mexico:
comparison of microfossil preservation. Abstracts of the Conference Publication for the Geological
Society of America 92nd Annual Meeting, 493.
Perch-Nielsen, K. 1979. Calcareous nannofossils from the Cretaceous between the North Sea and the
Mediterranean. In Aspekte dpr Kreide Europas (Ed. J. Wiedmann). IUGS Series A Number 6,
223-272.
Perch-Nielsen, K. 1985. Mesozoic calcareous nannofossils. In Plankton Stratigraphy (Ed. H. M. Bolli,
J. B. Saunders & K. Perch-Neilsen). Cambridge: Cambridge University Press, 427-554.
Peypouquet, J. P. 1983. Krithe and Parakrithe in the Kef section (northeast Tunisia) around the
Cretaceous-Tertiary boundary: paleohydrological implications. In Applications of Ostracoda
(Ed. R. F. Maddocks). llnie~ersity of Houston Geosciences, 510~519.
Phillips, W. 1818. A selection of facts from the best authorities arranged so far as to form an outline of the
geology of England and Wales. London.
Phillips, W. 1821. Remarks on the chalk cliffs in the neighbourhood of Dover, and on the Blue Marie
covering the Green Sand, near Folkestone. Transactions of the Geological Society of London, 5, 16 -51.
102 I Jarvls et a/

Pomerol. B. 1983. Geochemistry of the late Crnomanian: early ‘l‘uronian chalks of the f’arls Basin.
XLlanganese and carbon isotopes in carbonates as palaeoceanographic indicators. C’retuceous Research.
4,85- 93.
Porkomq, V. 19630. Karsteneis gen. n. (Ostracoda, Crustacea) from the Lpper C’retaceous of Bohemia.
C’asopis pro mineralogii a geologC. 8, 3OU4.
Porkorn-, V., 19636. The rer.ision of C’ythereis ornatissimn (Keuss, 1846) (Ostracoda, Crustaceal.
Rozpmvy Ceskoslownskt! Akademie 13d Prugue 73, 1 -5Y.
Price, F. C;. H. 1877. On the beds between the Gault and I!pper Chalk near Folliestone. Quarterly Journnl
of the Geological Society of London. 33, 43lGM8.
Rawson, I’. F., Curry, D., Dilley, 1;. C., Hancock, J. M., Kennedy, 51;. J., Neale, J. &l., Wood, C. J. &
iV;orssam, B. C. 1978. A correlation of Cretaceous rocks in the British Isles. Special Report of the
Geological Society, 9, 70 pp.
Rhoads, D. C. & Morse, J. \$‘. 1971. Evolutionary and ecological significance of oxygen deficient marine
basins. Lethaia, 4, 413428.
Robaszynski, F. & Amtdro, F. 1986. The C re t aceous of the Boulonnais (France) and a comparison with
the Cretaceous of Kent (United Kingdom). Proceedings of the Geologists’ ilssociation, 97, 171-208.
Robaszynski. F. & Caron. 11. i97Ya. Atlas de Foraminif&es planctoniques du C&ace moyen (Mer
BorGale et Ttthys). Premitre Partie. Cahiers de ,2~icropalPontologie, 1, 185 pp.
Robaszynski, F. & Caron, M. 19796. Atlas de Foraminif&es planctoniques de C&a& moyen (Xler
BorGale et TPthys). Dew&me Partie. Cnhiers de Micropal~ontologie, 2, 181 pp.
Robaszynski, F., AmCdro. F., Foucher, J.-C., Gaspard. D., Magniez-Jannin, F., Illanivit, H. & Sornay. J.
1980. Synth&e biostratigraphique de 1’Aptien au Santonien du Boulonnais a partir de sept groupes
pal&ontologiques: foraminif&es, nannoplancton, dinoflagell&s et macrofaunes: zonations micro-
pal&mtologiques intGgre& dans le cadre du C&a& bortal nerd-europ&en. Rwrte de MirrtJ-
palPontolo&. 22, 195 321.
Robaszynski, F., AlcaydP. G., Ambdro, F., Badillet, G., Ijamotte. R., Foucher, J.-C.. Jardin&, S.,
Legoux, O., Manwit, H., Monciardini. C. & Sornay, J, 1982. Le Turonien de la r&ion-type:
Saumurois et Touraine: Stratigraphie, biozonations, &dimentologie. Bulletin des Cmtres de RP-
rherches d’Explorntion Production Elf-Aquitaine. 6, 11 Y-225.
Robinson. N. D. 1985. Field meeting: the Chalk of the Kent coast . 2 I st August 1983. Proceedings of the
Geologists’ Association, 96, YS-Y6.
Robinson, NV. D. 1986. Lithostratiyraphy of the Chalk Group of the North Downs, south-east England.
Proceedings of the Geologists’ Association, 97, IJl -170.
Rosenberg, R. 1977. Benthic macrofossil dynamics, production and dispersion in an oxygen-drticient
estuary of West Su-eden. Journal of Experimental Marine Biology and Ecology, 36, 107-133.
Rosenberg, R. 1980. Effect of oxygen deficiency on benthic macrofauna in fjords. In Fjord Oceanograp/tJ
(Eds H. J. Freeland, I). 31. Farmer and C. D. Ixvings). ,~.+?‘fO Conference Series 4.9, Plenum Press,
109%-l15.
Roth, P. II. 1981. YIid-Cretaceous calcareous nanno-plankton from the central Pacific: implications for
p&oceanography. IGtial Reports of the Deep Sea Drilling Project, 62, 471-489.
Roth, P. II. & Bowdler, J. L. 1981. Middle Cretaceous calcareous nanno-plankton biostratigraphy and
oceanography of the Atlantic Ocean. Special Rchlication of the Society ofEconomic Paleontologists and
IVIineralogists, 32, 517~~5%.
Rowe, A. L%‘. 1900. The zones of the White Chalk of the English coast I: Kent and Sussex. Proceedings of
the Geologists’ Association, 16, 289-368.
Sa\Gn, S. M. 1977. The history of the Earth’s surface temperature during the past 100 million years.
dnnual Reviezc of the Earth and Planetar?, Sciences, 5, 319-355.
Schlangar, S. 0. 8r Jenkyns, H. C. 1976. Cretaceous oceanic anosic events: causes and consequences,
Geo1o.gie en Mijnbouw, 55, ! 79~ 1%.
Schlanger, S. O., Arthur, 41. A., Jenkyns. H. C. & Scholle, I’. A. 1987. The Crnomanian-Turonian
Oceanic Anoxic Event, I: Stratigraphy and distribution of organic carbon-rich beds and the marine
“C excursion. In Marine Petroleum Source Rocks (Eds J. Brooks and A. Fleet). Geological Soriet)
Special Alblication, 26, Oxford: Blackwell, 371 399.
Scholle, P. I-\. 1974. Diagenesis of [Jpper Cretaceous chalks from England, northern Ireland, and the
North Sea. In Peiagic Sediments on Lartd and under the Sea (Eds K. J H su and H. C. Jenkyns). Sprciai
Publication of the Ilrternatzonal .Issociation of Sedimentologists, 1, 177 210.
Scholle, P. A. 1977. Chalk diagenesis and its relation to petroleum exploration: Oil from chalks, a modern
mlraclr? Amerirnn Association of Petroleum Geologists’ Bulletin. 64, 67-87.
Scholle, I’. A. & Arthur, $1. A. 1980. Carbon isotope fluctuations in Cretaceous pelagic limestones:
potential stratigraphic and petroleum exploration tool. American Association of Petroleum Geologists’
Bulletin, 64, 67--X7.
Scholle. P. A., Arthur, WI. A. & Ekdale, A. -4. 19X3. Pelagic Environment. In Carbonate Deposition&
Erwironments (Eds P. A. Scholle, I>. G. Bebout and C. H. Moore). American Association of Petroleum
Geologists’ Memoir, 33, 61 Y%69 1.
Microfossil Assemblages and the Oceanic Anoxic Event 103

Smart, J. G. O., Bisson, G. 8i \Vorssam, B. C. 1966. Geology of the country around Canterbur> and
Folkestone. Memoirs of the Geological Suwey of Great Britain. London: H.M.S.O.. 114-174.
Stein, R. 1986~. Organic carbon and sedimentation rate: further evidence for anoxic deep-water
conditions in the Cenomanian/Turonian Atlantic Ocean. Marine Geology, 72, 199-209.
Stein, R. 19866. Surface-water paleo-productivity as inferred from sediments deposited in oxic and
anoxic deep-water environments of the Mesozoic Atlantic Ocean. In Biogeot-hem&r_%, of Black Shales
(Eds E. T. Degens, P. A. Meyers and S. C. Brassell). Mitteilungen aus dem Geologisch-
Paliiontologischen Institut der Unisersitiit Hamburg SCOPE/UNEP Sonderband Heft, 60, 55-70.
Stein, R., Rullkotter, J. &W&e. D. W. 1986. Accumulation of organic-carbon-rich sediments in the late
Jurassic and Cretaceous Atlantic Ocean. Chemical Geology, 56, 1 32.
Sues% E., Kulm, L. D. & Killingley, J. S. 1987. Coastal upuelling and a history of orgamc-rich mudstone
deposition off Peru. In Marine Petroleum Source Rocks (Eds J. Brooks and A. J. Fleet). Geological
Society Special Publication, 26, 181- 197.
Summerhayes, C. P. 1987. Organic-rich Cretaceous sediments from the ?;orth Atlantic. In Murine
Petroleum Source Rocks (Eds J. Brooks and A. J, Fleet). (;eological Society Special Publication. 26,
301 -316.
‘I’lrcher, B. A. & Jarvis, 1. 1987. Dinofagellate cysts and stratigraph! of the ‘I‘uronian (Upper Cretacrous)
chalk near Beer, southeast Devon, England. In Micropalaeontology oj’ Carbo~zate Environments (Ed.
41. B. Hart). Ellis Horwood British Micropalaeontological Society Series, 138 -175.
Troger. K.-A. 1981. Zu problemen der biostratigraphie der lnoceramcn und der untergliederung des
Cenomans und Turons in Mittel- und Osteuropa. Newslettrus on Stratigraphy, 9, 139- 156.
‘I‘yson, R. \‘. 1987. The genesis and palynofacies characteristics of marine petroleum source rocks. In
Marine Petroleum Source Rocks (Eds J. Brooks and .A. J. Fleet). Geological Society Special Publicntion,
26,47-67.
Verwutere, T. L.. Mullins, H. ‘l‘.. McDougall, K. & Thompson, J. B. 19X7. Sedimentation across the
central California owgen minimum zone: an alternatisc coastal upwelling sequence. Jonrnrrl af
Sedimentary Petrology, 57, 709 722.
\V;lples. r). W. 8r Cunningham, R. 1985. Shipboard organic geochemistry, I,cg 80. Deep Sea Drilling
Project. Initial Reports of the Deep Sea Drilling Proyrt, 80, 9-V-~968.
\Veaver, I’. P. E. 1981. The distribution of Ostracoda in the British Cenomanian. In ,2Iicrofossi/s,fronr
Recent and Fossil Shelf Seas (Eds J. W. Ncale and 1\I. D. Bras&). Ellis IIorwood British
>ltcropalaeontological Society Series, 156- 162.
\Vravrr. I’. I’. E. 1982. Ostracoda from the British Lov.er Chalk and I’lenus xlarls. .2lonogrupb of the
I’cllaroiltograp/lical Society, 135, l-127.
\Vhitaker. W.. Bristow, H. \V. & Hughes. M. 1872. The geology of the London Basin Part 1: The Chalk
and the Eocene beds of the southern and wcstcrn tracts. _~Ienloir of thr Grologicnt Suwe_v of En,qlnnd
mnd IVales. 4, 30-36.
\Vhitc. H. J. 0. 1909. Geology of the Country around Basingstoke. illemoiv of t/w Geological Suuw~~ c!f
Great Britain, 119 pp.
\Vhite, I I. J. 0. 1928. The geology of the country near Ramsgate and Do\ er. .IIeemoirs of the Geo/o,qic,cd
SUI-7.e~ of England and Wales. London: H.1I.S.O., 9X pp.
Wonders, A. .A. H. 1980. Middle and late Cretaceous planktonic I;oraminifera of the x\estern
1Iediterranean area. Utrecht Micropalaeontolgical Bulletin, 24, 1 158.
\Vorrd. C. J. & Smith, E. G. 197X. Lithostratigraphical classification of the Chalk in North \-orkshirr,
Clumberside and Lincolnshire. Proceedings of the Ii,rkshire Geological Society. 42, 263-287.
\Voodroof. P. B. 1981. Fauna1 and stratigraphic studies in the Turonian of the Anglo-Paris Basin. PhD.
Thesis. University of Oxford.
\Vright. C. \V\;. 8 Kennedy, \VI‘. J. 1981. The Ammonoidea of the Plenus \Iarls and the Middle (‘hdk.
Zlonograph of the Palaeontogvuphiral Society, 134 (560), 148 pp.
\‘un, H.-S. 1981. Dinoflagellaten aus der Oberkreide (Santon) van \Vestfalen. Palrreontogyaphi~~~ Aht. F%
177. 1 x0.

You might also like