ArticlePDF Available

Palaeoenvironmental significance of carbon- and oxygen-isotope stratigraphy of marine Triassic-Jurassic boundary sections in SW Britain

Authors:

Abstract and Figures

Carbon-isotope stratigraphy is a useful tool for stratigraphic correlation, especially for strata deposited during major perturbation of the carbon cycle that affected the marine, terrestrial and atmospheric reservoirs. For the Triassic-Jurassic boundary, effectively defined by a first-order mass extinction, major fluctuations in carbon-isotope values have been well documented, but these datasets have generally been derived from bulk-rock samples. Hence, the extent to which features of the isotopic curve reflect diagenetic alteration or changing proportions of constituent materials in unconstrained, Here, carbon- and oxygen-isotope data are presented from well-preserved oyster shells (Liostrea) comprising low-magnesium calcite, a mineral species relatively resistant to diagenetic alteration. Samples were obtained from Lavernock Point, Glamorgan, Wales, a coastal section close to a candidate stratotype for the base of the Jurassic at St. Audrie's Bay. Somerset, England. The carbon-isotope signature from St Audrie's Bay, previously defined on the basis of analysis of bulk organic matter, is confirmed by our new data. Major features are (1) the upper part of an 'initial' negative isotope excursion in the lowest part of the section, followed by (2) a pronounced positive excursion, and (3) an extended 'main' negative isotope excursion in the highest part of the section. The data confirm that the carbon-isotope stratigraphy previously documented from bulk organic matter in SW England records the chemical composition of the contemporaneous seawater. Bulk carbonates sampled over the same interval near Lyme Regis, England, show similar trends to those from oyster calcite in the lower part of the study section, but there are more (13)C-depleted values up-section. These lower values probably result from an admixture of primary and diagenetic carbonate. Palaeotemperatures calculated from oxygen-isotope values from Lavernock Point oyster shells are relatively cool at the beginning of the positive carbon-isotope excursion, and increased by up to 10 degrees C during the main negative carbon-isotope excursion. The new results are compatible with the view that positive carbon-isotope excursions correspond to times of low atmospheric carbon dioxide content, whereas negative carbon-isotope excursions correspond to times of high atmospheric carbon dioxide content, as is also found to be the case during the Early Jurassic (Toarcian) Oceanic Anoxic Event. The Mg/Ca and Sr/Ca ratios and delta(18)O of investigated Liostrea hisingeri show no correlation, supporting data from modern bivalves that indicate that incorporation of Mg and Sr is controlled mainly by factors other than temperature.
Content may be subject to copyright.
Journal of the Geological Society, London, Vol. 166, 2009, pp. 431–445. doi: 10.1144/0016-76492007-177.
431
Palaeoenvironmental significance of carbon- and oxygen-isotope stratigraphy of
marine TriassicJurassic boundary sections in SW Britain
CHRISTOPH KORTE
1
*, STEPHEN P. HESSELBO
1
, HUGH C. JENKYNS
1
,
ROSALIND E. M. RICKABY
1
& CHRISTOPH SPO
¨
TL
2
1
Department of Earth Sciences, University of Oxford, Parks Road, Oxford OX1 3PR, UK
2
Institut fu
¨
r Geologie und Pala
¨
ontologie, Universita
¨
t Innsbruck, Innrain 52, A-6020 Innsbruck, Austria
*Corresponding author (e-mail: korte@geo.ku.dk)
Abstract: Carbon-isotope stratigraphy is a useful tool for stratigraphic correlation, especially for strata
deposited during major perturbations of the carbon cycle that affected the marine, terrestrial and atmospheric
reservoirs. For the TriassicJurassic boundary, effectively defined by a first-order mass extinction, major
fluctuations in carbon-isotope values have been well documented, but these datasets have generally been
derived from bulk-rock samples. Hence, the extent to which features of the isotopic curve reflect diagenetic
alteration or changing proportions of constituent materials is unconstrained. Here, carbon- and oxygen-isotope
data are presented from well-preserved oyster shells (Liostrea) comprising low-magnesium calcite, a mineral
species relatively resistant to diagenetic alteration. Samples were obtained from Lavernock Point, Glamorgan,
Wales, a coastal section close to a candidate stratotype for the base of the Jurassic at St Audrie’s Bay,
Somerset, England. The carbon-isotope signature from St Audrie’s Bay, previously defined on the basis of
analysis of bulk organic matter, is confirmed by our new data. Major features are (1) the upper part of an
‘initial’ negative isotope excursion in the lowest part of the section, followed by (2) a pronounced positive
excursion, and (3) an extended ‘main’ negative isotope excursion in the highest part of the section. The data
confirm that the carbon-isotope stratigraphy previously documented from bulk organic matter in SW England
records the chemical composition of the contemporaneous seawater. Bulk carbonates sampled over the same
interval near Lyme Regis, England, show similar trends to those from oyster calcite in the lower part of the
study section, but there are more
13
C-depleted values up-section. These lower values probably result from an
admixture of primary and diagenetic carbonate. Palaeotemperatures calculated from oxygen-isotope values
from Lavernock Point oyster shells are relatively cool at the beginning of the positive carbon-isotope
excursion, and increased by up to 10 8C during the main negative carbon-isotope excursion. The new results
are compatible with the view that positive carbon-isotope excursions correspond to times of low atmospheric
carbon dioxide content, whereas negative carbon-isotope excursions correspond to times of high atmospheric
carbon dioxide content, as is also found to be the case during the Early Jurassic (Toarcian) Oceanic Anoxic
Event. The Mg/Ca and Sr/Ca ratios and ä
18
O of investigated Liostrea hisingeri show no correlation,
supporting data from modern bivalves that indicate that incorporation of Mg and Sr is controlled mainly by
factors other than temperature.
The TriassicJurassic (TJ) transition, c. 200 Ma ago, was a
time of mass extinction that affected both the marine and
continental biota (Hallam & Wignall 1997, and references there-
in). The patterns of environmental change that accompanied the
extinction, and the cause or causes, are vigorously debated (see,
e.g. Hesselbo et al. 2007a, for summary). As is the case for
many other episodes of major environmental change, the TJ
boundary is characterized by major perturbations to the carbon
cycle, demonstrated by carbon-isotope profiles generated from
several parts of the world (McRoberts et al. 1997; Pa´lfy et al.
2001, 2007; Ward et al. 2001, 2007; Hesselbo et al. 2002, 2004;
Guex et al. 2004; Galli et al. 2005, 2007; Kuerschner et al.
2007; Williford et al. 2007). Although there is generally good
agreement between the shapes of curves from different regions, a
shortcoming of the carbon-isotope profiles published hitherto is
that they are almost entirely based on analyses of bulk organic
matter or bulk carbonate, where the exact nature, origin and
proportion of the components is indeterminate. The few existing
data from diagenetically unaltered marine skeletal low-Mg
calcite are not sufficient to construct stratigraphically extended
isotope curves (e.g. van de Schootbrugge et al. 2007). Here,
analyses of shell material from the oyster Liostrea hisingeri
(Nilsson) are presented, carefully screened for diagenetic altera-
tion, and collected from numerous horizons across and above the
TJ boundary in SW Britain.
Similarities between the TJ boundary events and those
occurring during Phanerozoic Oceanic Anoxic Events have been
highlighted by a number of workers (e.g. Cohen & Coe 2002;
Jenkyns et al. 2002). Apart from the development of large-
amplitude carbon-isotope excursions, similarities include anoma-
lies and trend reversals in Sr and Os isotope records (Korte et al.
2003; Cohen & Coe 2007), and rapid turnover or crises amongst
shallow-marine carbonate-producing organisms (Hautmann 2006;
Kiessling et al. 2007; Tomasˇovy
´
ch & Siblı´k 2007). It has been
shown for other intervals of major environmental change that
carbon-isotope fluctuations are correlated with marine and atmo-
spheric palaeotemperature fluctuations, and such co-variations
have to be taken into account during consideration of the forcing
functions involved (e.g. Jenkyns 2003). Although data for
palaeotemperature changes accompanying TJ boundary events
are debated (e.g. McElwain et al. 1999; Hubbard & Boulter
2000), van de Schootbrugge et al. (2007) have demonstrated the
potential for oxygen-isotope and Mg/Ca analyses of the oysters
from St Audrie’s Bay to indicate palaeo-seawater temperature, in
this case documenting a temperature increase coincident with a
shift to lighter carbon-isotope values over some c. 3.5 m of strata.
Herein, oxygen-isotope and elemental concentration data are
presented for well-preserved low-Mg-calcite oysters for three
sections in SW Britain spanning the TJ boundary. These new
records greatly extend the dataset from the St Audrie’s Bay
section (van de Schootbrugge et al. 2007) both in terms of
quantity and temporal range.
Geological setting
Oysters have been collected at the highest level of stratigraphic
resolution possible from the upper Lilstock Formation and lower
Blue Lias Formation at Lavernock Point, Glamorgan, South
Wales (Fig. 1; Table 1), as well as from St Audrie’s Bay and
Watchet, Somerset, SW England. The lithostratigraphy of these
sections has been described by Richardson (1905, 1911), Whit-
taker (1978), Whittaker & Green (1983), Waters & Lawrence
(1987), Hesselbo et al. (2004) and Hounslow et al. (2004).
Lavernock Point represents the best location to obtain oysters at
a high resolution across the TJ interval and through lower
Hettangian strata. This section is biostratigraphically well cali-
brated (Trueman 1920; Hodges 1994) and is easily compared
with the succession at St Audrie’s Bay where the ä
13
C
org
curve
of Hesselbo et al. (2002) was generated. Additionally, whole-
rock carbonates were sampled from the Pinhay Bay section, near
Lyme Regis (Devon, SW England; Fig. 1). The stratigraphy of
this section has been described by Richardson (1906), Lang
(1924), Hallam (1960a), Hesselbo & Jenkyns (1995), Wignall
(2001a) and Hesselbo et al. (2004).
Materials and methods
Although it was not possible in every case to identify the genus
and/or species of the sampled shell fragments, most probably all
isotope data originate from Liostrea hisingeri Nilsson because
this form is the only oyster described from the sampled levels
(Richardson 1905; Trueman 1920; Hodges, pers. comm.). Oyster
shells consist generally of an outer simple prismatic layer and a
prominent inner foliated layer (Carter 1990; Hautmann 2006).
Both layers consist, at least for Jurassic oysters, primarily of
calcite and, particularly in the case of foliated layers, of low-Mg
calcite (LMC). Aragonitic elements may also have been present
in some oyster shells, but their expression appears to have been
suppressed after the earliest Jurassic (Hautmann 2006). In the
present study, only material from the foliated layers was
analysed.
It is thought that bivalves secrete their shells extracellularly
(Lowenstam & Weiner 1989), although studies of artificially
damaged specimens of the modern oyster Crassostrea virginica
(Gmelin) have shown that living hemocytes are present during
the rapid growth of prismatic and foliated LMC layers during
Fig. 1. (a) TriassicJurassic
palaeogeography (modified from Ziegler
1990; McHone 2000; Hesselbo et al. 2002).
CAMP, Central Atlantic Magmatic
Province. White square indicates UK study
area. (b) Location map for Lavernock Point,
Watchet, St Audrie’s Bay, and Lyme Regis
(modified after Hesselbo et al. 2004). (c)
Summary of pre-existing bulk organic
matter carbon-isotope data from St Audrie’s
Bay (Hesselbo et al. 2002) showing the
principal features discussed in the text. TJ
boundary is positioned on the basis of
correlation using carbon isotopes with the
proposed GSSP at Kuhjoch, Austria (details
provided by Hillebrandt, pers. comm.).
C. KORTE ET AL.432
regeneration (Mount et al. 2004), a phenomenon that could be
intracellular. However, whatever the nature of the process
involved, it has been observed that C. virginica precipitates the
prismatic and foliated shell layers close to oxygen- and carbon-
isotope equilibrium with ambient water (Surge et al. 2001).
Low-Mg calcite is relatively resistant to diagenetic alteration,
thus minimizing the potential resetting of the primary geochem-
ical signal from seawater (e.g. Veizer et al. 1997a,b, 1999;
Jenkyns et al. 2002). Observed ä
18
O and ä
13
C variations for
modern (Andrus & Crowe 2000; Surge et al. 2001) and ancient
(Kirby et al. 1998; Kirby 2000) oysters reflect climatic and/or
environmental changes. Despite the inherent resistance of low-
Mg calcite to post-depositional alteration, in the present study
each oyster was screened by chemical and optical techniques.
The methods used include optical microscopy, scanning electron
microscopy (SEM) and elemental analysis by inductively coupled
plasma-atomic emission spectroscopy (ICP-AES). The textural
observations and elemental data have been utilized to evaluate
the isotopic data.
Oysters were prepared by flaking shell fragments using a
needle. Splinters of the foliated layers were inspected and
handpicked under a binocular microscope. Weathered fragments,
attached sedimentary grains, and crack fillings were rejected.
Optically well-preserved shell material was checked by SEM to
obtain information on microstructural characteristics. Only fo-
liated shell layers with smooth surfaces were regarded as being
well preserved. The preservation of foliated structure indicates
that post-depositional dissolution and reprecipitation were negli-
gible (Fig. 2).
The Sr and Mn elemental concentrations (Table 1) were
determined at the Ruhr-University Bochum by ICP-AES, either
on aliquots of phosphoric acid remaining after reaction of shell
splinters for evolution of CO
2
(GasBench method), or on
additional separated shell splinters generated during sample
preparation for the isotope work (Prism II method). International
limestone standard reference materials (CCH-1, GBW 03105)
were analysed for Sr and Mn along with the samples, and the
accuracy and precision of these elemental analyses was better
than 3% and 1%, respectively. Obtained Sr and Mn concentra-
tions were utilized to evaluate the pristine preservation of LMC
shell material (Brand & Veizer 1980). Although variations of
these elements exist for palaeo-seawater throughout the Phaner-
ozoic (Steuber & Veizer 2002), and are concentrated to different
degrees in modern bivalves (Vander Putten et al. 2000), the
content of Sr and Mn provide further information regarding the
degree of preservation of the shells. With increasing diagenetic
influence, Sr is removed and Mn is added (Brand & Veizer
1980). To be consistent with a number of previous studies (Korte
et al. 2003, 2005a,b, 2006, 2008), samples with less than 400
ìgg
1
Sr and/or more than 250 ìgg
1
Mn were classified as
being altered, and their ä
18
O values are considered suspect. It
should be emphasized that none of the selection criteria are
perfectly reliable and the use of only one criterion (e.g. trace-
element concentrations) might be misleading when evaluating
the extent of alteration. The final evaluation of samples as either
pristine or altered was therefore based on optical and micro-
structural criteria, as well as trace-element chemistry (Table 1).
Only carbon- and oxygen-isotope data from well-preserved
oysters will be considered in the following presentation, discus-
sion and interpretation.
Samples of about 2 mg of shell fragments (oysters) or fine-
grained carbonate (whole rocks) were analysed isotopically for
ä
18
O and ä
13
C at the Department of Earth Sciences, University
of Oxford, using a VG Isogas Prism II mass spectrometer with an
on-line VG Isocarb common acid-bath preparation system. In the
instrument, the powdered sample is reacted with purified phos-
phoric acid (H
3
PO
4
)at908C. Calibration to the V-PDB standard
via NBS-19 is made daily using the Oxford in-house (NOCZ)
Carrara Marble standard. Reproducibility of replicated standards
is typically better than 0.1‰ for ä
13
C and ä
18
O.
Single shell splinters (,0.5 mm) with a mass of about 0.1 mg
were prepared in c. 1 mm steps in the growth direction for six
well-preserved oysters from the Lavernock Point section to
examine intra-shell variation in ä
13
C and ä
18
O values. Surface
layers were removed and sampling started about 1 mm below the
surface. These considerably smaller samples were analysed for
ä
18
O and ä
13
C on a GasBench II linked to a ThermoFinnigan
Delta
plus
XL mass spectrometer at the Institut fu
¨
r Geologie und
Pala¨ontologie of the Universita¨t Innsbruck (Spo¨tl & Vennemann
2003). The long-term precision (1 s.d.) is better than 0.06‰ for
ä
13
C and better than 0.08‰ for ä
18
O. Carbon- and oxygen-
isotope values were calibrated against V-PDB and are reported in
the standard notation (Tables 1 and 2).
Some carbonate samples (Table 1; c. 0.2 mg) were analysed at
the Institut fu
¨
r Geologie, Mineralogie und Geophysik, Ruhr-
Universita¨t Bochum also using a GasBench II linked to a
ThermoFinnigan Delta S mass spectrometer. The results from the
carbonate samples were corrected to the nominal values for the
carbon- and oxygen-isotope standards CO-1 and CO-8. Reprodu-
cibility was better than 0.1‰ for both ä
13
C and ä
18
O. Carbon-
and oxygen-isotope values were calibrated against V-PDB and
are reported in the standard notation (Table 1). Tests in all
three laboratories demonstrated that NBS-19 and NOCZ showed
similar results and are within 0.1‰ for both ä
13
C and ä
18
O.
Shell splinters of about 0.25 mg were powdered for the Mg/Ca
ratio measurements. To be consistent with previous work, the
samples were cleaned by a procedure developed and revised by
Boyle & Keigwin (1985) whereby ferromanganese oxides and
organic matter were removed (see Rickaby & Halloran 2005).
After a weak acid leach (0.001M HNO
3
), the samples were
dissolved in 0.5 ml 0.22M HNO
3
and centrifuged for 5 min at
5000 r.p.m. Analysis was carried out by inductively coupled
plasma-mass spectrometry (ICP-MS) at Oxford University, using
a PerkinElmer PESCIEX ELAN 6100 DRC Quadrupole ICP-MS
system fed by a low-uptake (100 ìl min
1
) spray-chamber
nebulizer (see Rosenthal et al. 1999). Data were collected for
four masses:
46
Ca,
26
Mg,
88
Sr, and
55
Mn with the ICP-MS
system set to peak hopping mode using a method adapted from
Harding et al. (2006).
Results
Carbon and oxygen isotopes
Most of the oysters originate from the Lavernock Point section
(Fig. 1) and, for this locality, the lithology, stratigraphy, and
oxygen- and carbon-isotope data are shown in Figure 3. The
lowest well-preserved oysters retrieved originated from the lower
Langport Member and collection was continued up through the
Langport Member and the Blue Lias Formation, the latter
including the P. planorbis, C. johnstoni and the W. portlocki
ammonite subzones of the P. planorbis and A. liasicus ammonite
zones of the Hettangian (Fig. 3). The lowest well-preserved
oysters originated from the base of Bed A, from the lower
Langport Member. A correlation from Lavernock to the proposed
Global Stratotype and Stratigraphic Point (GSSP) at Kujoch,
Austria (Hillebrandt, pers. comm.), cannot be accomplished on
the basis of biostratigraphy because the UK sections lack
TRIASSICJURASSIC BOUNDARY 433
Table 1. Positions, material analysed, carbon- and oxygen-isotope values, Mg/Ca and Sr/Ca ratios, and Mn and Sr concentrations of the analysed
samples
Sample Height above
base Langport
Mbr (cm)
Material ä
13
C
(‰ V-PDB)
ä
18
O
(‰ V-PDB)
Isotope
laboratory
Mg/Ca
(mM M
1
)
Sr/Ca
(mM M
1
)
Mn
(ìgg
1
)
Sr
(ìgg
1
)
Pristine oysters
LAV 109 36 Oyster 2.87 0.42 Ox 9.03 0.58 237 431
LAV 109 36 Oyster 3.30 0.46 Bo 8.42 0.55 237 431
LAV 109 b 36 Oyster 3.00 0.39 Ox nd nd 237 431
LAV 109 36 Oyster 3.76 0.18 Inn nd nd 237 431
LAV 110 91 Oyster 2.24 0.09 Bo 5.08 0.55 242 570
LAV 110 91 Oyster 2.89 0.12 Ox 5.44 0.63 242 570
LAV 261 91 Oyster 3.18 0.88 Bo 4.08 0.55 126 597
LAV 110-1 91 Oyster 2.83 0.10 Inn 4.98 0.55 216 472
LAV 110-1 91 Oyster 2.86 0.34 Inn 7.10 0.55 216 472
LAV 261 91 Oyster 3.36 0.96 Inn 4.03 0.52 126 597
LAV 261 91 Oyster 3.29 0.55 Inn 4.13 0.57 126 597
LAV 257 171 Oyster 3.51 0.19 Bo nd nd 174 530
LAV 257 171 Oyster 3.55 0.12 Bo nd nd 174 530
LAV 258 171 Oyster 4.63 1.62 Bo 2.28 0.48 126 570
LAV 257 171 Oyster 3.94 0.04 Inn nd nd 174 530
LAV 262 275 Oyster 3.62 0.49 Bo nd nd 80 560
LAV 262 275 Oyster 4.11 0.35 Bo 5.50 0.56 44 538
LAV 264 285 Oyster 3.62 0.08 Bo nd nd 81 482
LAV 204 290 Oyster 4.04 0.05 Bo 4.69 0.55 185 585
LAV 204 290 Oyster 4.04 0.05 Bo 6.29 0.50 185 585
LAV 204 290 Oyster 4.04 0.05 Bo 5.39 0.57 185 585
LAV 204 290 Oyster 4.04 0.05 Bo 6.29 0.50 185 585
LAV 203-1 303 Oyster 3.02 0.17 Bo nd nd 32 518
LAV 203-1 303 Oyster 3.31 0.02 Ox nd nd 32 518
LAV 200 332 Oyster 4.34 0.81 Ox 11.57 0.69 57 646
LAV 200 332 Oyster 4.53 0.65 Bo 9.92 0.61 57 646
LAV 201 332 Oyster 4.77 1.02 Ox 4.48 0.52 53 608
LAV 200 332 Oyster 4.45 0.36 Inn 5.23 0.56 57 646
LAV 205 362 Oyster 3.68 0.87 Ox nd nd 83 515
LAV 206 403 Oyster 3.88 0.80 Ox 6.87 0.52 64 465
LAV 206 403 Oyster 3.88 0.80 Ox 13.33 0.62 64 465
LAV 207-1 403 Oyster 3.69 0.66 Ox 9.56 0.53 48 576
LAV 207-2 403 Oyster 4.00 0.34 Bo 8.55 0.54 64 566
LAV 207-2 403 Oyster 4.16 0.48 Ox nd nd 64 566
LAV 207-1 403 Oyster 2.93 1.06 Inn 5.59 0.55 48 576
LAV 208 425 Oyster 4.25 0.22 Bo 6.55 0.50 31 594
LAV 209 448 Oyster 2.98 0.93 Ox nd nd 189 547
LAV 210-1 451 Oyster 3.33 1.17 Ox 6.34 0.46 128 531
LAV 210-1 451 Oyster 3.33 1.17 Ox 7.89 0.51 128 531
LAV 210-2 451 Oyster 3.43 1.09 Inn 5.43 0.50 67 496
LAV 210-2 451 Oyster 3.55 1.18 Ox nd nd 67 496
LAV 218 477 Oyster 3.17 1.25 Bo 5.63 0.50 93 574
LAV 218 477 Oyster 3.38 1.18 Ox nd nd 93 574
LAV 212 497 Oyster 3.26 0.38 Ox 7.62 0.52 50 580
LAV 212 497 Oyster 3.26 0.38 Ox 9.59 0.58 50 580
LAV 220 500 Oyster 3.95 0.82 Ox nd nd 41 537
LAV 213-1 515 Oyster 2.54 0.43 Bo nd nd 75 527
LAV 213-3 515 Oyster 3.24 0.24 Bo 5.39 0.46 44 965
LAV 214-2 526 Oyster 3.51 0.27 Ox 5.73 0.51 17 498
LAV 214-2 526 Oyster 3.52 0.16 Inn 4.65 0.51 17 498
LAV 215-1 537 Oyster 2.76 0.46 Ox nd nd 37 480
LAV 215-2 537 Oyster 3.02 0.54 Ox 7.44 0.51 53 501
LAV 215-2 537 Oyster 3.02 0.54 Ox 9.23 0.54 53 501
LAV 216 605 Oyster 2.93 1.19 Bo nd nd 15 638
LAV 216 605 Oyster 3.07 1.05 Inn 6.07 0.55 15 638
LAV 217-1 624 Oyster 2.15 1.79 Ox 4.30 0.56 29 583
LAV 217-1 624 Oyster 2.51 1.25 Bo 5.76 0.63 29 583
LAV 219 648 Oyster 2.45 0.69 Inn 5.77 0.54 110 434
LAV 232 877 Oyster 2.16 1.09 Bo nd nd 5 506
LAV 232 877 Oyster 1.86 1.06 Inn 3.96 0.51 5 506
LAV 230 882 Oyster 2.36 1.78 Bo nd nd 52 517
LAV 233 898 Oyster 1.99 1.74 Bo nd nd 10 590
LAV 235-A 968 Oyster 1.71 0.97 Bo nd nd 15 500
LAV 235-B 968 Oyster 2.01 0.06 Bo nd nd 46 430
LAV 237 1102 Oyster 2.48 0.99 Bo nd nd 21 461
( continued)
C. KORTE ET AL.434
relevant ammonites. However, on the basis of the carbon-isotope
data available from St Audrie’s Bay, together with the new data
presented here, the TJ boundary should lie within the Langport
Member at Lavernock.
The general trend in ä
13
C values through the Lavernock Point
section is marked by a prominent positive excursion, which starts
from c. 3‰ in the lower Langport Member, increasing by about
2‰ to highest values of c. 5‰ in the lower Blue Lias, followed
by a 3‰ decrease to c. 2‰ just below the base of the planorbis
Zone (Fig. 3). The carbon-isotope values remain relatively low
up to the portlocki Subzone of the liasicus Zone (with slight
variations in the johnstoni Zone).
Highly resolved ä
13
C profiles in the growth directions of the
shells for six well-preserved oysters from Lavernock Point are
shown in Figure 4 and Table 2: two of these samples were also
profiled perpendicular to growth direction. Carbon-isotope values
show a range of about 1‰ in LAV 110 (mean: 2.8 0.2‰,
n ¼ 44), less than 1‰ in LAV 258 (mean: 4.5 0.2‰, n ¼ 8),
about 0.6‰ in LAV 201 (mean: 4.9 0.2‰, n ¼ 8), more than
1.1‰ in LAV 208 (mean: 3.8 0.4‰, n ¼ 8), nearly 1.3‰ in
LAV 210 (mean: 3.5 0.3‰, n ¼ 26) and about 1.4‰ in LAV
219 (1.8 0.4‰, n ¼ 28). The ä
13
C values perpendicular to the
growth direction vary much less: about 0.6‰ in both LAV 201
(mean: 4.9 0.2‰, n ¼ 7) and LAV 210 (mean: 3.6 0.1‰,
n ¼ 16).
Carbon-isotope values from the Lyme Regis section (Fig. 5;
Table 3), based on analysis of whole-rock carbonate, are about 3
0.5‰ within the lower Langport Member. A decrease in
values of these samples starts in the upper Langport Member,
falling to about 1‰ just below the base of the planorbis Zone.
This declining trend continues, reaching 0.2‰ at the top of the
planorbis Subzone. The carbon-isotope values remain low (c.
0‰) in the johnstoni and portlocki Subzones, but it should be
noted that only one ä
13
C value was obtained for each of these
two units.
The ä
18
O values from Lavernock Point oysters are about 0‰
(1‰) in the lower Langport Member, increase somewhat to
highest values of more than 1.6‰ in the upper Langport Member
and decline sharply (with some reversals) to values of about
2‰ in the portlocki Subzone of the Blue Lias Formation. The
ä
18
O variations in shell-growth direction and perpendicular-to-
growth direction of each of the six well-preserved oysters show
the features described below (Fig. 4, Table 2). The oxygen-
isotope values vary by about 2.1‰ in LAV 110 (mean: 0.2
0.5‰, n ¼ 44), by more than 2‰ in LAV 258 (mean: +0.6
0.9‰, n ¼ 8), by only 0.2‰ in LAV 201 (mean: 1.0 0.1‰,
n ¼ 8), by about 1.1‰ in LAV 208 (mean: 0.3 0.4‰,
n ¼ 8), by 0.7‰ in LAV 210 (mean: 1.0 0.2‰, n ¼ 26) and
by more than 1.6‰ in LAV 219 (1.5 0.5‰, n ¼ 28). The
ä
18
O variations perpendicular to the growth direction are 0.2‰
Table 1. ( continued )
Sample Height above
base Langport
Mbr (cm)
Material ä
13
C
(‰ V-PDB)
ä
18
O
(‰ V-PDB)
Isotope
laboratory
Mg/Ca
(mM M
1
)
Sr/Ca
(mM M
1
)
Mn
(ìgg
1
)
Sr
(ìgg
1
)
LAV 238 1102 Oyster 1.69 1.46 Bo nd nd 84 449
LAV 237 1102 Oyster 1.73 1.12 Inn 9.14 0.53 21 461
LAV 243 1239 Oyster 2.04 0.89 Bo nd nd 80 456
LAV 244 1249 Oyster 2.73 1.52 Bo nd nd 48 607
LAV 246 1473 Oyster 2.59 1.41 Bo nd nd 16 480
LAV 246 1473 Oyster 2.78 1.26 Inn 3.43 0.55 16 480
LAV 247 1552 Oyster 2.32 2.05 Bo nd nd 173 560
LAV 249 1708 Oyster 1.61 2.03 Bo nd nd 84 433
LAV 249 1708 Oyster 2.01 1.24 Bo nd nd 89 425
LAV 256 1748 Oyster 1.26 1.79 Bo nd nd 127 466
LAV 255-C 1781 Oyster 1.89 1.88 Bo nd nd 190 520
LAV 255-C 1781 Oyster 1.98 1.39 Inn 2.91 0.55 190 520
LAV 252 2075 Oyster 1.37 1.98 Bo nd nd 103 525
LAV 251-A 2080 Oyster 1.82 1.81 Bo nd nd 180 480
LAV 251-B 2080 Oyster 1.70 2.23 Bo nd nd 61 500
SAB 165-1 415 Oyster 2.23 1.29 Ox 5.14 0.45 7 471
Wa 1 285 Oyster 3.64 0.71 Ox nd nd 49 550
Wa 2 325 Oyster 3.22 0.35 Ox 8.69 0.52 57 574
Altered oysters
LAV 107-1 4 Oyster 0.59 1.45 Ox nd nd 1373 687
LAV 210-3 451 Oyster 3.19 1.31 Ox nd nd nd nd
LAV 211 470 Oyster 4.06 0.10 Ox nd nd 440 570
LAV 213-2 515 Oyster 2.03 2.03 Bo nd nd 338 411
LAV 214-1 526 Oyster 2.20 1.00 Bo nd nd 227 386
LAV 231 885 Oyster 2.25 1.49 Bo nd nd nd nd
LAV 245-B 1473 Oyster 2.37 1.17 Bo nd nd 53 388
LAV 255-B 1781 Oyster 1.80 1.62 Bo nd nd 267 519
SAB 100-1 (below) 265 Oyster 2.72 2.76 Ox nd nd 1636 1100
SAB 100-2 (below) 265 Oyster 4.11 3.68 Ox nd nd 1392 681
SAB 179 (below) 1115 Oyster 3.49 10.35 Bo nd nd 817 463
Wa BBL 0 Oyster 5.59 6.22 Ox nd nd 743 347
Bulk rocks
LAV 110-1 91 Bulk rock 0.43 5.19 Bo nd nd nd nd
LAV 210 451 Bulk rock 0.87 4.67 Bo nd nd nd nd
LAV 201 332 Bulk rock 0.10 5.19 Bo nd nd nd nd
LAV 219 648 Bulk rock 0.77 3.50 Bo nd nd nd nd
LAV, Lavernock Point; SAB, St Audrie’s Bay; Wa, Watchet; Ox, Oxford; Bo, Bochum; Inn, Innsbruck; nd, not determined.
TRIASSICJURASSIC BOUNDARY 435
in LAV 201 (mean: 0.9 0.1‰, n ¼ 7) and more than 0.4‰
in LAV 210 (mean: 1.2 0.1‰, n ¼ 16).
To aid discussion of the oxygen- and carbon-isotope trends, all
ä
13
C and ä
18
O values from well-preserved oysters at Lavernock
Point and from the St Audrie’s Bay area (including the previously
published data of van de Schootbrugge et al. (2007)) are
compiled in Figure 6, together with the whole-rock carbonate
ä
13
C data from the Lyme Regis section.
The isotopic values of well-preserved oyster shells collected
from the same stratigraphic level, and therefore geologically
coeval, show spreads of more than 1‰ for both ä
13
C and ä
18
O
(Figs 3, 4 and 6). Such variation is also seen in single shells and
is the norm for modern oysters (Surge et al. 2001), reflecting
seasonal environmental and temperature change. Therefore, when
discussing long-term changes in environmental parameters, the
general trends enclosed in the envelopes shown in Figure 6
should be considered. It should be noted that the scatter of
carbon- and oxygen-isotope values for the six oysters fit perfectly
within the general trends, constituting evidence that both ä
13
C
and ä
18
O values represent a primary seawater signal.
Mg/Ca and Sr/Ca ratios
Mg/Ca and Sr/Ca ratios have been generated for 48 well-
preserved oyster shells (Table 1) and these vary in the range 2.3
13.3 mmol mol
1
and 0.40.7 mmol mol
1
, respectively. It has
been reported that Mg and Sr concentrations in bivalve shells
depend on ambient seawater temperature and increase with
warming (Dodd 1965; see also Stecher et al. 1996). Notably,
Klein et al. (1996a) documented that for the modern bivalve
Mytilus trossulus (Gould) the Mg/Ca ratios tend to reach higher
values with rising seawater temperature. However, in a more
extensive study Vander Putten et al. (2000) found that Mg/Ca
ratios in Mytilus edulis (Linnaeus) shell calcite showed consider-
able deviations from those expected from water temperature
alone, implying that metabolic processes are involved in Mg
incorporation into the calcite framework. For strontium also it is
most likely that water temperature is not the unique control on
the Sr/Ca ratios in bivalve shells because, in modern bivalves,
metabolic or kinetic effects and/or seawater chemistry have also
been shown to be important (Klein et al. 1996b; Vander Putten et
al. 2000; Lorrain et al. 2005). Work on other modern bivalve
groups has added further weight to the view that their Mg/Ca
and Sr/Ca ratios are unreliable palaeotemperature proxies (Freitas
et al. 2006).
The temperature dependence of Mg/Ca and Sr/Ca ratios in
Early Jurassic oysters was assessed by plotting these ratios
against ä
18
O (Fig. 7), a procedure that assumes that ä
18
Ois
overwhelmingly controlled by seawater temperature (discussed
further below). There are no significant correlations, and there-
Fig. 2. SEM images of well-preserved foliated layers of oyster shells. (a) LAV 110; (b) LAV 204; (c) LAV 219; (d) LAV 251. Scale bars represent 2 ìm.
C. KORTE ET AL.436
Table 2. Highly resolved ä
13
C and ä
18
O profiles of six well-preserved
Liostrea samples
Sample Section Direction Distance
(mm)
ä
13
C
(‰ V-PDB)
ä
18
O
(‰ V-PDB)
LAV 110-1 1 Grow. 0.50 3.13 0.34
LAV 110-1 1 Grow. 1.25 3.13 0.83
LAV 110-1 1 Grow. 1.90 2.96 0.46
LAV 110-1 1 Grow. 8.25 2.89 0.58
LAV 110-1 1 Grow. 8.71 2.56 1.26
LAV 110-1 1 Grow. 9.14 3.07 0.24
LAV 110-1 1 Grow. 9.14 3.12 0.14
LAV 110-1 1 Grow. 9.48 3.08 0.07
LAV 110-1 1 Grow. 9.56 3.11 0.14
LAV 110-1 1 Grow. 9.86 3.05 0.40
LAV 110-1 1 Grow. 10.20 2.91 0.14
LAV 110-1 1 Grow. 10.97 3.04 0.26
LAV 110-1 1 Grow. 11.48 3.00 0.48
LAV 110-1 1 Grow. 11.90 2.56 0.39
LAV 110-1 1 Grow. 12.45 2.39 0.59
LAV 110-1 1 Grow. 12.96 2.43 0.40
LAV 110-1 1 Grow. 13.73 2.74 0.26
LAV 110-1 1 Grow. 14.24 2.82 0.37
LAV 110-1 1 (b) Grow. 1.80 2.93 0.54
LAV 110-1 1 (b) Grow. 5.00 2.72 0.20
LAV 110-1 1 (b) Grow. 6.15 2.99 0.45
LAV 110-1 1 (b) Grow. 6.25 2.81 0.16
LAV 110-1 1 (b) Grow. 7.25 2.88 0.76
LAV 110-1 1 (b) Grow. 9.27 2.85 0.46
LAV 110-1 1 (b) Grow. 10.29 2.61 0.69
LAV 110-1 1 (b) Grow. 10.75 2.96 0.55
LAV 110-1 1 (b) Grow. 11.01 2.49 0.29
LAV 110-1 1 (b) Grow. 11.65 2.35 0.12
LAV 110-1 1 (b) Grow. 12.67 2.44 0.53
LAV 110-1 1 (b) Grow. 13.43 2.79 0.49
LAV 110-1 1 (b) Grow. 13.81 2.76 0.86
LAV 110-1 2 Grow. 0.85 2.82 0.24
LAV 110-1 2 Grow. 1.75 2.98 0.54
LAV 110-1 2 Grow. 2.65 2.26 0.77
LAV 110-1 2 Grow. 3.65 2.57 0.01
LAV 110-1 2 Grow. 5.15 2.74 0.65
LAV 110-1 2 Grow. 5.30 2.83 0.69
LAV 110-1 2 Grow. 6.10 2.83 0.58
LAV 110-1 2 Grow. 6.50 2.96 0.02
LAV 110-1 2 Grow. 7.40 2.87 0.31
LAV 110-1 2 Grow. 7.65 2.63 0.50
LAV 110-1 2 Grow. 8.50 2.97 0.07
LAV 110-1 2 Grow. 10.50 2.71 0.04
LAV 110-1 2 Grow. 11.50 2.25 0.39
LAV 201 1 Grow. 0.75 4.74 0.94
LAV 201 1 Grow. 1.40 4.65 1.05
LAV 201 1 Grow. 1.80 4.55 1.01
LAV 201 1 Grow. 2.75 4.92 0.92
LAV 201 1 Grow. 4.25 4.89 0.98
LAV 201 1 Grow. 4.50 5.14 0.89
LAV 201 1 Grow. 5.10 4.95 1.08
LAV 201 1 Grow. 6.05 5.05 0.87
LAV 201 2 Perp. 0.80 5.07 0.79
LAV 201 2 Perp. 2.20 4.69 1.02
LAV 201 2 Perp. 3.75 4.50 0.99
LAV 201 2 Perp. 3.90 4.98 0.89
LAV 201 2 Perp. 5.05 4.99 0.82
LAV 201 2 Perp. 6.00 4.83 0.84
LAV 201 2 Perp. 7.50 5.05 0.81
LAV 208-p 1 Grow. 3.50 3.56 0.57
LAV 208-p 1 Grow. 5.50 3.89 0.46
LAV 208-p 1 Grow. 7.30 3.61 0.73
LAV 208-p 1 Grow. 9.00 3.62 0.60
LAV 208-p 1 Grow. 10.00 3.14 0.56
LAV 208-p 1 Grow. 12.50 3.91 0.02
LAV 208-p 1 Grow. 14.50 4.18 0.17
LAV 208-p 1 Grow. 17.00 4.26 0.33
LAV 210-2 1 Grow. 0.90 3.78 0.93
LAV 210-2 1 Grow. 0.90 3.71 1.31
LAV 210-2 1 Grow. 1.00 3.07 0.75
LAV 210-2 1 Grow. 2.00 3.46 1.02
LAV 210-2 1 Grow. 2.50 3.35 1.21
LAV 210-2 1 Grow. 2.60 3.12 1.19
LAV 210-2 1 Grow. 3.80 3.46 0.87
LAV 210-2 1 Grow. 4.40 2.97 1.15
LAV 210-2 1 Grow. 4.85 3.47 0.99
LAV 210-2 1 Grow. 4.90 3.04 1.23
LAV 210-2 1 Grow. 5.80 3.57 0.92
LAV 210-2 1 Grow. 6.00 4.13 0.88
LAV 210-2 1 Grow. 7.50 3.85 0.92
LAV 210-2 1 Grow. 7.50 3.31 1.00
LAV 210-2 1 Grow. 7.60 3.67 0.70
LAV 210-2 1 Grow. 8.50 3.82 0.84
LAV 210-2 1 Grow. 9.50 3.55 0.98
LAV 210-2 1 Grow. 10.05 4.01 0.63
LAV 210-2 1 Grow. 11.00 3.63 0.79
LAV 210-2 1 Grow. 11.30 3.79 0.96
LAV 210-2 1 Grow. 12.50 3.52 0.93
LAV 210-2 1 Grow. 12.50 3.70 1.09
LAV 210-2 1 Grow. 13.60 3.65 0.92
LAV 210-2 1 Grow. 14.35 2.88 0.92
LAV 210-2 1 Grow. 14.40 3.54 0.96
LAV 210-2 1 Grow. 15.50 3.51 0.96
LAV 210-2 2 Perp. 3.00 3.78 0.96
LAV 210-2 2 Perp. 4.50 3.69 1.13
LAV 210-2 2 Perp. 6.00 3.64 1.13
LAV 210-2 2 Perp. 6.90 3.53 1.34
LAV 210-2 2 Perp. 8.25 3.52 1.11
LAV 210-2 2 Perp. 4.50 3.65 1.14
LAV 210-2 2 Perp. 3.20 3.80 0.94
LAV 210-2 2 Perp. 3.85 3.74 1.05
LAV 210-2 2 Perp. 9.20 3.52 1.08
LAV 210-2 2 Perp. 10.50 3.59 1.16
LAV 210-2 2 Perp. 11.50 3.48 1.17
LAV 210-2 2 Perp. 11.10 3.50 1.25
LAV 210-2 2 Perp. 12.50 3.54 1.17
LAV 210-2 2 Perp. 13.50 3.55 1.27
LAV 210-2 2 Perp. 15.15 3.40 1.03
LAV 210-2 2 Perp. 15.00 3.24 1.41
LAV 219 1 Grow. 0.80 2.30 0.93
LAV 219 1 Grow. 1.70 1.64 2.08
LAV 219 1 Grow. 3.00 1.69 1.65
LAV 219 1 Grow. 4.25 1.43 2.30
LAV 219 1 Grow. 5.50 2.63 0.67
LAV 219 1 Grow. 6.00 1.89 1.10
LAV 219 1 Grow. 7.30 1.94 1.56
LAV 219 1 Grow. 6.50 1.56 1.80
LAV 219 1 Grow. 7.50 2.09 1.37
LAV 219 1 Grow. 8.50 1.75 1.82
LAV 219 1 Grow. 9.85 2.43 1.04
LAV 219 1 Grow. 11.00 1.79 1.84
LAV 219 1 Grow. 11.50 2.40 0.84
LAV 219 1 Grow. 13.50 2.40 0.71
LAV 219 1 Grow. 12.50 2.01 1.24
LAV 219 2 Grow. 0.85 2.20 0.71
LAV 219 2 Grow. 1.50 1.51 1.32
LAV 219 2 Grow. 2.60 1.43 1.84
LAV 219 2 Grow. 3.50 1.38 2.07
LAV 219 2 Grow. 5.50 1.57 1.62
LAV 219 2 Grow. 6.50 1.28 1.82
LAV 219 2 Grow. 4.50 1.31 2.21
LAV 219 2 Grow. 7.50 1.36 1.94
LAV 219 2 Grow. 8.50 1.24 1.95
LAV 219 2 Grow. 9.50 1.70 1.44
LAV 219 2 Grow. 10.50 1.82 1.17
LAV 219 2 Grow. 11.50 2.34 0.73
LAV 219 2 Grow. 12.50 2.30 0.86
LAV 258 1 Grow. 4.50 4.89 1.64
LAV 258 1 Grow. 6.50 4.69 1.45
LAV 258 1 Grow. 7.50 4.46 1.36
LAV 258 1 Grow. 8.50 4.58 0.89
LAV 258 1 Grow. 14.50 4.29 0.23
LAV 258 1 Grow. 15.00 4.18 0.11
LAV 258 1 Grow. 22.50 4.47 0.43
LAV 258 1 Grow. 24.50 4.21 0.52
Grow., growth direction; Perp., perpendicular to growth direction.
TRIASSICJURASSIC BOUNDARY 437
fore it can be inferred that Mg/Ca and Sr/Ca ratios of the
investigated Mesozoic Liostrea hisingeri are determined mainly
by factors other than temperature. It is notable that these results
are in contrast to those of van de Schootbrugge et al. (2007),
who reported a distinct increase of oyster-calcite Mg/Ca and Sr/
Ca ratios with decreasing ä
18
O values, but this was based on a
much smaller dataset.
Discussion
Significance for correlation of TJ boundary strata
Carbon-isotope shifts reflect perturbations in the Earth’s carbon
cycle at a variety of spatial scales (Kump & Arthur 1999). It is
understood that seawater ä
13
C is controlled by the burial and re-
oxidation of
12
C-enriched organic matter within the ocean
Fig. 3. Stratigraphic section for Lavernock Point, showing sample locations and carbon- and oxygen-isotope values for well-preserved oyster (Liostrea)
shell calcite. TJ boundary position is less precisely definable in comparison with St Audrie’s Bay, but on the basis of carbon-isotope stratigraphy must lie
somewhere within the marly upper beds of the Langport Member.
C. KORTE ET AL.438
atmosphere system, linked to several factors, such as atmospheric
CO
2
levels, nutrient supply, sedimentation rates, net primary
productivity, biological isotope fractionation or sea-level changes
(Scholle & Arthur 1980; Jenkyns 1996; Hayes et al. 1999; Kump
& Arthur 1999; Jarvis et al. 2006). Additional mechanisms
suggested to add significant masses of isotopically light carbon
include input of volcanic mantle-derived CO
2
into the ocean
atmosphere system (Hansen 2006), sudden release of methane
from gas hydrates (Dickens et al. 1997; Hesselbo et al. 2000),
thermal metamorphism of organic-rich sediments (Svensen et al.
2004; McElwain et al. 2005) or overturn of
12
C-enriched
(anoxic) oceanic bottom waters (e.g. Ku
¨
spert 1982; Knoll et al.
1996).
Many carbon-isotope excursions are global in scale, and are
registered in a wide range of marine and continental deposits
such as platform carbonates, calcareous pelagic sediments,
organic-rich shales, terrestrial palaeosols and lacustrine deposits.
The coeval nature of the ä
13
C fluctuations makes it possible to
use the peaks and troughs for trans-continental stratigraphic
correlation for many stage boundaries (e.g. PermianTriassic
boundary; Korte & Kozur 2005), and this has also been proposed
for the TJ boundary through the correlation of ä
13
C
org
fluctua-
tions of several sections in North America and Europe (Hesselbo
et al. 2002, 2004; McRoberts et al. 2007). The carbon-isotope
stratigraphy from the well-preserved oysters from TJ boundary
sections presented here (Fig. 6), show the same features as
illustrated by Hesselbo et al. (2002, 2004) for St Audrie’s Bay
(SW England), Kuerschner et al. (2007) for the Tiefengraben
section (Austria), Ward et al. (2007) for New York Canyon
Ferguson Hill (Nevada, USA), and Williford et al. (2007) for
Kennecott Point (Queen Charlotte Islands, British Columbia,
Canada). The same major carbon-isotope fluctuations in the well-
preserved marine oyster carbonate described here are also seen
in bulk organic material (see Fig. 1), namely: (1) a sharp ‘initial’
negative excursion (here seen only in relatively low carbon
isotope values at the base of the Langport Member), followed by
(2) a ‘boundary’ positive excursion, and (3) an extended ‘main’
negative excursion.
It is noteworthy that the amplitudes of the carbon-isotope
excursions are much larger (commonly twice the size) for the
marine organic matter than for the carbonates, a common feature
of carbon-isotope anomalies characterizing several Mesozoic
Cenozoic events (Arthur et al. 1988; Pagani et al. 2006; Hesselbo
et al. 2007a). The general shape of the curve has been identified
in other TriassicJurassic boundary sections for marine carbo-
nates (e.g. McRoberts et al. 1997; Pa´lfy et al. 2001; Galli et al.
2005; van de Schootbrugge et al. 2007), but insufficient sampling
density, lack of organic-matter-based records for the same strata,
the occurrence of depositional hiatuses, or the effects of signifi-
cant sediment redeposition, have made the identification of the
negativepositivenegative geometry of the excursion in these
sections uncertain.
The new data presented here demonstrate that the isotopic
signature described is a seawater signal, and the similarity of the
ä
13
C records from carbonate and organic matter from different
sections in the Tethyan and in North American regions strongly
suggests that the carbon-isotope excursions are of global extent.
Consequently, further definition of a high-resolution carbon-
isotope stratigraphy has significant potential for improved corre-
lation of a time interval beset with problems using only
biostratigraphy.
In addition to insufficient sampling density and/or depositional
hiatuses, diagenetic alteration must be considered when using the
carbon-isotope curve for stratigraphic correlation. For the Lang-
port Member and the lower Blue Lias Formation, carbon-isotope
values from whole-rock carbonate at Lyme Regis show super-
ficially the same ä
13
C trend as oysters (Fig. 6), but the bulk-rock
data are much more
13
C-depleted in the planorbis, johnstoni and
portlocki subzones higher in the Blue Lias (about 2‰ lower than
stratigraphically equivalent oysters). These isotopically lighter
Fig. 4. ä
18
O and ä
13
C variations of foliated
layers in growth direction for three well-
preserved Liostrea shells. Two cut sections
have been analysed from sample LAV 110
(lower Langport Member), and one of these
has been sampled for different depths below
the surface (section 1: open circles, c.1mm
below the surface; grey circles, c. 1.3 mm
below the surface). All other sections have
been sampled c. 1 mm below the surfaces
to avoid contamination.
TRIASSICJURASSIC BOUNDARY 439
whole-rock ä
13
C data are most probably diagenetically compro-
mised because these carbonate beds are partly concretionary in
origin, with
12
C-enriched carbon originating, at least in part (Fig.
8: dark grey field), from oxidized organic matter (Hallam 1960b;
Weedon 1986; Sheppard et al. 2006). On the other hand, the data
from the organic-lean Langport Member can be interpreted as
more accurately representing the positive ä
13
C excursion, and
these relatively heavy values show that bulk-rock data, in this
case deriving from massive micritic limestones, can preserve the
primary seawater signature. Such carbonates were deposited as
lime mud and peloids and lithified during diagenesis in an
essentially closed system, such that the final ä
13
C values of these
limestones were probably derived from the dissolving solid phase
of the original lime mud (see Veizer 1983). A correlation
between Lyme Regis, St Audrie’s Bay and Lavernock Point,
based on carbon-isotopes, implies that the Langport Member at
Lyme Regis is coeval with argillaceous facies in the Langport
Member and Blue Lias at the other two localities (see Hesselbo
et al. 2004).
Palaeoenvironmental change at the TJ boundary
The relationship between the TJ environmental crisis and flood-
basalt volcanism of the Central Atlantic Magmatic Province (Fig.
1) has been much discussed, including the possibility that
volcanogenic CO
2
was responsible for the negative carbon-
isotope excursions (McHone 1996; Marzoli et al. 1999; Pa´lfy et
al. 2000; Wignall 2001b; Courtillot & Renne 2003; Pa´lfy 2003;
Nomade et al. 2007). Bulk-rock organic-carbon data from St
Audrie’s Bay show an abrupt ‘initial’ negative ä
13
C excursion
(Fig. 6). Such apparently sudden negative shifts could be
accounted for by a local hiatus obscuring a more gradual change
(see Hesselbo et al. 2002; Lucas et al. 2007), but the preponder-
ance of records showing abrupt change (e.g. Hesselbo et al.
2002; Guex et al. 2004; Ward et al. 2007; Williford et al. 2007)
implies either a relatively sudden event or an extraordinary
synchroneity of sedimentary condensation.
Rather than being produced directly by volcanogenic CO
2
(see
Self et al. 2006), such sudden negative carbon-isotope shifts are
probably accounted for by the release of isotopically light carbon
through other, less direct, mechanisms, such as those proposed
for the PalaeoceneEocene boundary and Toarcian Oceanic
Anoxic Event: namely, massive methane release from gas
hydrates (Dickens et al. 1995, 1997; Hesselbo et al. 2000; Kemp
et al. 2005), and/or the intrusion of large sills into organic
carbon-rich sedimentary strata (Svensen et al. 2004, 2007;
McElwain et al. 2005).
In the case of the TJ boundary, intrusion of Central Atlantic
Fig. 5. Lithology and bulk ä
13
C
carb
data for the section at Pinhay Bay, Lyme Regis, SW England.
C. KORTE ET AL.440
Magmatic Province magma into organic-rich lacustrine sediments
of the deep and extensive Triassic rift systems could have caused
similar effects. A large proportion of the Central Atlantic
Magmatic Province intrusive bodies are either eroded or deeply
buried beneath later Mesozoic deposits on the continental
margins (McHone 1996) and the nature of the material into
which the magma was intruded is generally unknown. In the
Newark Basin, one of these Triassic rift basins, dolerite sills are
locally intruded into Carnian lacustrine black shales of the
Lockatong Formation, which have undergone intense thermal
metamorphism (Van Houten 1969, 1971). Thus, although the
extent to which this may have been an effective process in
generation of isotopically light thermogenic methane is uncer-
tain, it is known that such processes occurred on at least a local
scale during TJ boundary time.
Positive carbon-isotope excursions are usually explained by
accelerated burial of organic carbon (Scholle & Arthur 1980;
Kump & Arthur 1999). For the pronounced earliest Jurassic
positive excursion (Fig. 6), an origin through globally accelerated
carbon burial cannot at present be demonstrated, largely because
the stratigraphic record for this time interval is so fragmentary.
At St Audrie’s Bay, although the equivalent strata are not
Table 3. Whole-rock carbon- and oxygen-isotope data for the Pinhay
Bay, Lyme Regis section
Sample Height above base of
Langport Mbr (cm)
ä
13
C
(‰ V-PDB)
ä
18
O
(‰ V-PDB)
LM 1 0 3.28 3.35
LM 2 39 3.53 2.93
LM 3 60 3.61 2.79
LM 4 93 3.36 3.41
LM 5 110 3.47 3.36
LM 6 138 3.55 3.16
LM 7 161 3.31 3.44
LM 8 185 4.03 1.76
LM 9 207 3.39 2.96
LM 10 243 3.88 2.23
LM 11 280 3.56 2.86
LM 12 319 3.58 2.94
LM 13 332 2.92 3.46
LM 14 354 2.92 3.93
LM 15 369 3.88 2.34
LM 16 420 3.70 2.46
LM 17 431 3.77 1.90
LM 18 459 3.73 2.19
LM 19 507 3.32 3.19
LM 20 520 2.86 4.56
LM 21 601 3.07 2.92
LM 22 628 3.16 2.86
LM 23 640 2.39 3.14
LM 24 649 2.43 2.81
LM 25 674 2.21 3.04
LM 26 709 1.60 4.58
LM 27 736 1.89 2.90
LM 28 759 1.39 2.08
LM 29 783 1.59 2.17
LM 30 796 1.52 1.98
LM 31 811 1.52 2.86
LM 32 848 1.58 3.16
LM 33 858 1.55 3.09
LM 34 869 1.60 2.79
LM 35 884 1.02 3.93
LM 36 909 1.11 3.09
LM 37 921 1.36 2.30
LM 38 948 1.19 2.23
LM 39 975 0.77 2.36
LM 40 994 0.80 2.32
LM 41 1083 0.45 1.53
LM 42 1114 0.01 2.62
LM 43 1182 -0.18 1.90
LM 44 1315 0.01 2.12
LM 45 1447 -0.14 1.89
Isotope laboratory: Oxford.
Fig. 6. ä
13
C and ä
18
O values for all well-preserved oysters for the
TriassicJurassic transition from Lavernock Point, Watchet and St
Audrie’s Bay (including data from van de Schootbrugge et al. 2007), and
carbon-isotope data from whole-rock carbonates from Lyme Regis. Bulk
organic ä
13
C data (Hesselbo et al. 2002, 2004) are also plotted for
comparison. Correlations between St Audrie’s Bay, Watchet and
Lavernock are based on lithostratigraphy and biostratigraphy. Correlation
between Lyme Regis and the other localities is based upon
biostratigraphy for the Blue Lias and carbon-isotope data for the
Langport Member. Ages are assigned to the data points based on the
position of samples at each locality within the recognized ammonite
subzones (see Jenkyns et al. 2002), with the addition of a notional
subzone to span the gap between the base of the planorbis Subzone and
the proposed (lower) base of the Hettangian. Data points in the Triassic
are plotted assuming a constant sedimentation rate through the earliest
Jurassic at Lavernock Point and St Audrie’s Bay.
TRIASSICJURASSIC BOUNDARY 441
markedly enriched in organic carbon, they are strongly laminated
(Hesselbo et al. 2004), indicating bottom-water anoxia and the
potential for enhanced organic matter burial at other localities.
Oxygen-isotope values in the calcitic shells of oysters are
controlled by the seawater ä
18
O, pH and temperature (Zeebe &
Wolf-Gladrow 2001). Assuming that seawater pH was similar to
the present-day value and seawater ä
18
O was about 1.2‰ in an
ice-free world (Zachos et al. 2001), bottom-water temperatures,
using the equation of O’Neil et al. (1969), were between ,7 and
14 8C for deposition of the upper Langport Member and between
c. 12 and 22 8C for the planorbis, johnstoni and portlocki
ammonite Subzones. Assuming no change in the local seawater
ä
18
O over the interval in question, these data indicate a tempera-
ture increase of more than 8 8C for the bottom waters in the
vicinity of the Lavernock Point locality. Starting from relatively
cool initial temperatures, a warming occurred in concert with the
shift towards negative carbon-isotope values.
It might be argued that the decreasing oxygen-isotope trend
that occurs leading into the planorbis Zone was caused, at least
in part, by a lowering of the local or global seawater ä
18
O rather
than a rise in temperature. Changes in global seawater ä
18
O
related to interaction of seawater with the lithosphere are on
multi-million year time scales (Gregory 1991; Veizer et al. 1999)
and can be excluded as an explanation. Short-term lowering of
the seawater ä
18
O could be caused by melting of continental ice,
in the manner observed for the post-Eocene ‘icehouse’ (Shackle-
ton & Opdyke 1973). The extent of continental ice during
supposed ‘greenhouse’ times is a matter of much speculation and
few data (Price 1999; Miller et al. 2005). Climate modelling
studies that have been carried out for the TJ boundary interval
have been aimed at testing the environmental response of the T
J world to increased rather than decreased atmospheric CO
2
partial pressures; nevertheless, polar regions of the Earth experi-
ence polar ice-forming conditions in simulations with 2 3 pre-
industrial atmospheric CO
2
content (Huynh & Poulsen 2005). It
is notable, however, that the lightening in oxygen-isotope values
observed in the present study would have corresponded to a
eustatic sea-level rise of some 200 m if generated by waning ice
sheets (see Miller et al. 2005) and, although a sea-level rise is
compatible with sequence stratigraphic interpretations for this
time interval, the large magnitude is not (Hesselbo 2008, and
references therein).
Lastly, a localized change in salinity may be postulated, either
from normal marine to freshwater conditions, or from hypersa-
line to normal marine conditions. An upward trend towards
greater freshwater influence is implausible because ammonites
appear at the culmination of the trend to lighter isotope values,
the exact opposite of what would be expected. An upward trend
from hypersaline to normal marine waters can also be ruled out
because Jurassic oysters and their modern counterparts are
intolerant of hypersaline conditions, occupying brackish to
normal marine habitats (Hendry & Kalin 1997; Fu
¨
rsich 1993). In
an isotopic study of Middle Jurassic oysters from England,
Hendry & Kalin (1997) explained heavy oxygen isotopes in
shells from nearshore settings as a result of evaporative concen-
tration within low-salinity waters of hydrodynamically closed
lagoons, a palaeogeographical setting that is very different from
the one considered here.
Faunal evidence, particularly the occurrence of corals, echino-
derms and conodonts, led Swift (1995) and Swift and Martill
(1999) to argue in favour of a normal marine environment for
the Langport Member in southern Britain. The absence of other
Fig. 7. Cross-plots for ä
18
O against Mg/Ca
and Sr/Ca.
Fig. 8. Cross-plot for ä
13
C against ä
18
O values of all pristine and altered
oysters (Lavernock Point, Watchet and St Audrie’s Bay) as well as bulk-
rock samples (Lyme Regis and Lavernock Point). It should be noted that
altered oysters as well as bulk-rock samples tend to show lower ä
18
O and
partly also lower ä
13
C values than unaltered oysters.
C. KORTE ET AL.442
typically normal marine taxa such as brachiopods and ammonites
may be due to the severe ecological disturbance that occurred
during the TriassicJurassic boundary mass extinction rather
than simple salinity control. Celestine in the Langport Member
of Devon was interpreted by Hesselbo & Jenkyns (1995) as a
possible replacement of evaporitic gypsum. However, this ob-
servation is of only marginal relevance to interpretation of the
Lavernock oyster oxygen-isotope data, which come from a
different facies and a distant location.
With regard to a detailed comparison between the carbon-
isotope and oxygen-isotope curves presented in this study, a
small difference in the position of the positive peaks is detected,
with the peak in oxygen-isotope values occurring a few deci-
metres lower in the section than the peak in carbon-isotope
values. However, these differences may reasonably be regarded
as a consequence of the variability exhibited within single shells
and between shells taken from the same horizon. The data do not
clearly define a phase lag between the oxygen- and carbon-
isotope curves.
The relationship between palaeotemperature change (as in-
ferred from oxygen-isotope values) and carbon-isotope fluctua-
tions is similar to that observed for the Toarcian Oceanic Anoxic
Event (Hesselbo et al. 2007b), during which high temperatures
coincided with strongly negative carbon-isotope values, and
lower temperatures with more positive carbon-isotope values
(and enhanced organic-carbon burial causing drawdown of atmo-
spheric CO
2
). The TJ boundary results are compatible with the
interpretation that positive carbon-isotope excursions correspond
to times of low atmospheric carbon dioxide content, and negative
carbon-isotope excursions correspond to times of high atmo-
spheric carbon dioxide content, implying significant swings in
the balance between carbon drawdown through organic-matter
burial and recycling of isotopically light carbon from endogenic
or exogenic sources.
Conclusions
Major carbon-isotope fluctuations are preserved within low-Mg
calcite shells of oysters collected at TJ boundary sections in
SW Britain. These fluctuations follow trends previously estab-
lished from bulk organic-matter samples from the same sedimen-
tary basin, and thus confirm that the observed isotopic signals
capture characteristics of the local water masses. Furthermore,
the similarity to carbon-isotope curves generated from other
basins suggests that the signal is global. The amplitude of the
carbonate carbon-isotope fluctuations is about half that observed
from organic matter, a feature common to other Mesozoic and
Early Cenozoic events affecting the global carbon cycle. The T
J boundary isotopic excursions occurred coincident with Central
Atlantic Magmatic Province continental flood-basalt volcanism,
and may have been triggered by the intrusion of mantle-derived
melts into carbon-rich sedimentary deposits and/or dissociation
of gas hydrates triggered by a global rise in temperature. The
pronounced positive carbon-isotope anomaly is associated with
relatively heavy seawater oxygen-isotope values, indicating cool
conditions, possibly associated with enhanced organic-carbon
burial and drawdown of atmospheric carbon dioxide. The subse-
quent trend towards lighter seawater oxygen-isotope values
indicates returning warmer temperatures that parallel a return to
lighter carbon-isotope values and inferred build-up of atmo-
spheric carbon dioxide. Mg/Ca and Sr/Ca ratios show no signifi-
cant correlation to ä
18
O, indicating that Mg and Sr incorporation
in the investigated Liostrea hisingeri is controlled mainly by
factors other than temperature.
We acknowledge P. Hodges (Cardiff) for scientific discussion in the field,
and N. Charnley, J. Arden, C.-J. de Hoog (all Oxford), U. Schulte, W.
Gosda (both Bochum) and M. Wimmer (Innsbruck) for stable-isotope and
elemental analyses. We thank the Deutsche Akademie der Naturforscher
Leopoldina (BMBF-LPD 9901/8-116) for contributions to the financing
of this project. P. Olsen is thanked for scientific discussion. Finally, we
thank J. Pa´lfy, I. Jarvis and an anonymous referee for their insightful
critical comments.
References
Andrus, C.F.T. & Crowe, D.E. 2000. Geochemical analysis of Crassostrea
virginica as a method to determine season of capture. Journal of Archae-
ological Science, 27, 3342.
Arthur, M.A., Dean, W.E. & Pratt, L.M. 1988. Geochemical and climatic
effects of increased marine organic carbon burial at the Cenomanian/Turonian
boundary. Nature, 335, 714717.
Boyle, E.A. & Keigwin, L.D. 1985. Comparison of Atlantic and Pacific
paleochemical records for the last 215,000 years: changes in deep ocean
circulation and chemical inventories. Earth and Planetary Science Letters,
76, 135150.
Brand, U. & Veizer, J. 1980. Chemical diagenesis of a multicomponent carbonate
system—1: Trace elements. Journal of Sedimentary Petrology, 50, 12191236.
Carter, J.G. 1990. Evolutionary significance of shell microstructure in the
Palaeotaxodonta, Pteriomorphia and Isofilibranchia. In: Carter, J.G. (ed.)
Skeletal Biomineralization, Patterns, Processes and Evolutionary Trends.Van
Nostrand Reinhold, New York, 135412.
Cohen, A.S. & Coe, A.L. 2002. New geochemical evidence for the onset of
volcanism in the Central Atlantic magmatic province and environmental
change at the TriassicJurassic boundary. Geology, 30, 267270.
Cohen, A.S. & Coe, A.L. 2007. The impact of the Central Atlantic Magmatic
Province on climate and on the Sr- and Os-isotope evolution of seawater.
Palaeogeography, Palaeoclimatology, Palaeoecology, 244, 374390.
Courtillot, V.E. & Renne, P.R. 2003. On the ages of flood basalt events.
Comptes Rendus Ge
´
oscience, 335, 113140.
Dickens, G.R., O’Neil, J.R., Rea, D.K. & Owen, R.M. 1995. Dissociation of
oceanic methane hydrate as a cause of the carbon isotope excursion at the
end of the Paleocene. Paleoceanography, 10, 965971.
Dickens, G.R., Castillo, M.M. & Walker, J.C.G. 1997. A blast of gas in the
latest Paleocene: simulating first-order effects of massive dissociation of
oceanic methane hydrate. Geology, 25, 259262.
Dodd, J.R. 1965. Environmental control of strontium and magnesium in Mytilus.
Geochimica et Cosmochimica Acta, 29, 385398.
Freitas, P.S., Clarke, L.J., Kennedy, H., Richardson, C.A. & Abrantes, F.
2006. Environmental and biological controls on elemental (Mg/Ca, Sr/Ca and
Mn/Ca) ratios in shells of the king scallop Pecten maximus. Geochimica et
Cosmochimica Acta, 70, 51195133.
Fu
¨
rsich, F.T. 1993. Palaeoecology and evolution of Mesozoic salinity-controlled
benthic macroinvertebrate associations. Lethaia, 26, 327346.
Galli, M.T., Jadoul, F., Bernasconi, S.M. & Weissert, H. 2005. Anomalies in
global carbon cycling and extinction at the Triassic/Jurassic boundary:
evidence from a marine C-isotope record. Palaeogeography, Palaeoclimatol-
ogy, Palaeoecology, 216, 203214.
Galli, M.T., Jadoul, F., Bernasconi, S.M., Cirilli, S., Weissert, H. 2007.
Stratigraphy and palaeoenvironmental analysis of the TriassicJurassic
transition in the western Southern Alps (Northern Italy). Palaeogeography,
Palaeoclimatology, Palaeoecology, 244, 5270.
Gregory, R.T. 1991. Oxygen isotope history of seawater revisited: Timescales for
boundary event changes in the oxygen isotope composition of seawater. In:
Taylor, H.P., O’Neil, J.R. & Kaplan, I.R. (eds) Stable Isotope Geochem-
istry: A Tribute to Samuel Epstein. Geochemical Society Special Publications,
3, 6576.
Guex, J., Bartolini, A., Atudorei, V. & Taylor, D. 2004. High-resolution
ammonite and carbon isotope stratigraphy across the TriassicJurassic
boundary at New York Canyon (Nevada). Earth and Planetary Science
Letters, 225, 29 41.
Hallam, A. 1960a. The White Lias of the Devon coast. Proceedings of the
Geologists’ Association, 71, 47 60.
Hallam, A. 1960b. A sedimentary and faunal study of the Blue Lias of Dorset and
Glamorgan. Philosophical Transactions of the Royal Society of London,
Series B, 243, 144.
Hallam, A. & Wignall, P.B. 1997. Mass Extinctions and their Aftermath. Oxford
University Press, Oxford.
Hansen, H.J. 2006. Stable isotopes of carbon from basaltic rocks and their possible
relation to atmospheric isotope excursions. Lithos, 92, 105116.
TRIASSICJURASSIC BOUNDARY 443
Harding, D.J., Arden, J.W. & Rickaby, R.E.M. 2006. A method for precise
analysis of trace element/calcium ratios in carbonate samples using quadru-
pole inductively coupled plasma mass spectrometry. Geochemistry, Geophy-
sics, Geosystems, 7, Q06003, doi:10.1029/2005GC001093.
Hautmann, M. 2006. Shell mineralogical trends in epifaunal Mesozoic bivalves
and their relationship to seawater chemistry and atmospheric carbon dioxide
concentration. Facies, 52, 417433.
Hayes, J.M., Strauss, H. & Kaufman, A.J. 1999. The abundance of
13
Cin
marine organic matter and isotopic fractionation in the global biogeochemical
cycle of carbon during the past 800 Ma. Chemical Geology, 161, 103125.
Hendry, J.P. & Kalin, R.M. 1997. Are oxygen and carbon isotopes of mollusc
shells reliable palaeosalinity indicators in marginal marine environments? A
case study from the Middle Jurassic of England. Journal of the Geological
Society, London, 154, 321333.
Hesselbo, S.P. 2008. Sequence stratigraphy and inferred relative sea-level change
from the onshore British Jurassic. Proceedings of the Geologists’ Association,
119, 1934.
Hesselbo, S.P. & Jenkyns, H.C. 1995. A comparison of the Hettangian to
Bajocian successions of Dorset and Yorkshire. In: Taylor, P.D. (ed.) Field
Geology of the British Jurassic. Geological Society, London, 105150.
Hesselbo, S.P., Gro
¨
cke, D.R., Jenkyns, H.C., Bjerrum, C.J., Farrimond, P.,
Morgans Bell, H.S. & Green, O.R. 2000. Massive dissociation of gas
hydrate during a Jurassic Oceanic Anoxic Event. Nature, 406, 392 395.
Hesselbo, S.P., Robinson, S.A., Surlyk, F. & Piasecki, S. 2002. Terrestrial and
marine extinction at the TriassicJurassic boundary synchronized with major
carbon-cycle perturbation: A link to initiation of massive volcanism?
Geology, 30, 251254.
Hesselbo, S.P., Robinson, S.A. & Surlyk, F. 2004. Sea-level change and facies
development across potential Triassic Jurassic boundary horizons, SW
Britain. Journal of the Geological Society, London, 161, 365379.
Hesselbo, S.P., McRoberts, C.A. & Pa
´
lfy, J. 2007a. TriassicJurassic boundary
events: problems, progress, possibilities. Palaeogeography, Palaeoclimatology,
Palaeoecology, 244, 110.
Hesselbo, S.P., Jenkyns, H.C., Duarte, L.V. & Oliveira, L.C.V. 2007b. Carbon-
isotope record of the Early Jurassic (Toarcian) Oceanic Anoxic Event from
fossil wood and marine carbonate (Lusitanian Basin, Portugal). Earth and
Planetary Science Letters, 253, 455470.
Hodges, P. 1994. The base of the Jurassic system: new data on the first appearance
of Psiloceras planorbis in southwest England. Geological Magazine, 131,
841844.
Hounslow, M.W., Posen, P.E. & Warrington, G. 2004. Magnetostratigraphy
and biostratigraphy of the Upper Triassic and lowermost Jurassic succession,
St Audrie’s Bay, UK. Palaeogeography, Palaeoclimatology, Palaeoecology,
213, 331358.
Hubbard, R.N.L.B. & Boulter, M.C. 2000. Phytogeography and paleoecology in
Western Europe and Eastern Greenland near the TriassicJurassic boundary.
Palaios, 15, 120131.
Huynh, T.T. & Poulsen, C.J. 2005. Rising atmospheric CO
2
as a possible trigger
for the end-Triassic mass extinction. Palaeogeography, Palaeoclimatology,
Palaeoecology, 217, 223242.
Jarvis, I., Gale, A.S., Jenkyns, H.C. & Pearce, M.A. 2006. Secular variation in
Late Cretaceous carbon isotopes: a new ä
13
C carbonate reference curve for
the CenomanianCampanian (99.670.6 Ma). Geological Magazine, 143,
561608.
Jenkyns, H.C. 1996. Relative sea-level change and carbon isotopes: data from the
Upper Jurassic (Oxfordian) of central and Southern Europe. Terra Nova, 8,
7585.
Jenkyns, H.C. 2003. Evidence for rapid climate change in the Mesozoic
Palaeogene greenhouse world. Philosophical Transactions of the Royal
Society of London, Series A, 361, 18851916.
Jenkyns, H.C., Jones, C.E., Gro
¨
cke, D.R., Hesselbo, S.P. & Parkinson, D.N.
2002. Chemostratigraphy of the Jurassic System: applications, limitations and
implications for palaeoceanography. Journal of the Geological Society,
London, 159, 351378.
Kemp, D.B., Coe, A.L., Cohen, A.S. & Schwark, L. 2005. Astronomical pacing
of methane release in the Early Jurassic period. Nature, 437, 396 399.
Kiessling, W., Aberhan, M., Brenneis, B. & Wagner, P.J. 2007. Extinction
trajectories of benthic organisms across the TriassicJurassic boundary.
Palaeogeography, Palaeoclimatology, Palaeoecology, 244, 201222.
Kirby, M.X. 2000. Paleoecological differences between Tertiary and Quaternary
Crassostrea oysters, as revealed by stable isotope sclerochronology. Palaios,
15, 132141.
Kirby, M.X., Soniat, T.M. & Spero, H.J. 1998. Stable isotope sclerochronology
of Pleistocene and Recent oyster shells (Crassostrea virginica). Palaios, 13,
560569.
Klein, R.T., Lohmann, K.C. & Thayer, C.W. 1996a. Bivalve skeletons record
sea-surface temperature and ä
18
O via Mg/Ca and
18
O/
16
O ratios. Geology, 24,
415418.
Klein, R.T., Lohmann, K.C. & Thayer, C.W.1996b. Sr/Ca and
13
C/
12
C ratios in
skeletal calcite of Mytilus trossulus: Covariation with metabolic rate, salinity,
and carbon isotopic composition of seawater. Geochimica et Cosmochimica
Acta, 60, 42074221.
Knoll, A.H., Bambach, R.K., Canfield, D.E. & Grotzinger, J.P. 1996.
Comparative Earth history and late Permian mass extinction. Science, 273,
452 457.
Korte, C. & Kozur, H.W. 2005. Carbon isotope stratigraphy across the Permian/
Triassic boundary at Jolfa (NW-Iran), Peitlerkofel (Sas de Pu
¨
tia, Sass de
Putia), Pufels (Bula, Bulla), Tesero (all three Southern Alps, Italy) and
Gerennava´r (Bu
¨
kk Mts., Hungary). Journal of Alpine Geology, 47, 119135.
Korte, C., Kozur, H.W., Bruckschen, P. & Veizer, J. 2003. Strontium isotope
evolution of Late Permian and Triassic seawater. Geochimica et Cosmochi-
mica Acta, 67, 4762.
Korte, C., Jasper, T., Kozur, H.W. & Veizer, J. 2005a. ä
18
Oandä
13
Cof
Permian brachiopods: A record of seawater evolution and continental
glaciation. Palaeogeography, Palaeoclimatology, Palaeoecology, 224,333
351.
Korte, C., Kozur, H.W. & Veizer, J. 2005b. ä
13
C and ä
18
O values of Triassic
brachiopods and carbonate rocks as proxies for coeval seawater and
palaeotemperature. Palaeogeography, Palaeoclimatology, Palaeoecology, 226,
287 306.
Korte, C., Jasper, T., Kozur, H.W. & Veizer, J. 2006.
87
Sr/
86
Sr record of
Permian seawater. Palaeogeography, Palaeoclimatology, Palaeoecology, 240,
89107.
Korte, C., Jones, P.J., Brand, U., Mertmann, D. & Veizer, J. 2008. Oxygen
isotope values from high latitudes: Clues for Permian sea-surface temperature
gradients and Late Palaeozoic deglaciation. Palaeogeography, Palaeoclim-
atology, Palaeoecology, 269, 1–16.
Kuerschner, W.M., Bonis, N.R. & Krystyn, L. 2007. Carbon-isotope stratigra-
phy and palynostratigraphy of the TriassicJurassic transition in the
Tiefengraben section—Northern Calcareous Alps (Austria). Palaeogeography,
Palaeoclimatology, Palaeoecology, 244, 257280.
Kump, L.R. & Arthur, M.A. 1999. Interpreting carbon-isotope excursions:
Carbonates and organic matter. Chemical Geology, 161, 181198.
Ku
¨
spert, W. 1982. Environmental changes during oil shale deposition as deduced
from stable isotope ratios. In: Einsele, G. & Seilacher, A. (eds) Cyclic and
Event Stratification. Springer, Berlin, 482501.
Lang, W.D. 1924. The Blue Lias of the Devon and Dorset coasts. Proceedings of
the Geologists’ Association, 35, 169185.
Lorrain, A., Gillikin, D.P., Paulet, Y.-M., Chauvaud, L., Le Mercier, A.,
Navez, J. & Andre
´
,L. 2005. Strong kinetic effects on Sr/Ca ratios in the
calcitic bivalve Pecten maximus. Geology, 33, 965968.
Lowenstam, H.A. & Weiner, S. 1989. On Biomineralization. Mollusca. Oxford
University Press, Oxford, 88134.
Lucas, S.G., Taylor, D.G., Guex, J., Tanner, L.H. & Krainer, K. 2007. The
proposed global stratotype section and point for the base of the Jurassic
System in the New York Canyon area, Nevada, USA. In: Lucas, S.G. &
Spielmann, J.A. (eds) Triassic of the American West. New Mexico Museum
of Natural History and Science, Bulletin, 40, 139168.
Marzoli, A., Renne, P.R., Piccirillo, E.M., Ernesto, M., Bellieni, G. & De
Min, A. 1999. Extensive 200-million-year-old continental flood basalts of the
central Atlantic magmatic province. Science, 284, 616618.
McElwain, J.C., Beerling, D.J. & Woodward, F.I. 1999. Fossil plants and
global warming at the TriassicJurassic boundary. Science, 285, 13861390.
McElwain, J.C., Murphy, J.W. & Hesselbo, S.P. 2005. Changes in carbon
dioxide during an oceanic anoxic event linked to intrusion of Gondwana
coals. Nature, 435, 479483.
McHone, J.G. 1996. Broad-terrane Jurassic flood basalts across northeastern North
America. Geology, 24, 319322.
McHone, J.G. 2000. Non-plume magmatism and rifting during the opening of the
central Atlantic Ocean. Tectonophysics, 316, 287296.
McRoberts, C.A., Furrer, H. & Jones, D.S. 1997. Palaeoenvironmental
interpretation of a TriassicJurassic boundary section from Western Austria
based on palaeoecological and geochemical data. Palaeogeography, Palaeo-
climatology, Palaeoecology, 136, 7995.
McRoberts, C.A., Ward, P.D. & Hesselbo, S.P. 2007. A proposal for the base
Hettangian Stage (¼ base Jurassic System) GSSP at New York Canyon
(Nevada, USA) using carbon isotopes. International Subcommission on
Jurassic Stratigraphy Newsletter, 34, 4349.
Miller, K.G., Wright, J.D. & Browning, J.V. 2005. Visions of ice sheets in a
greenhouse world. Marine Geology, 217, 215231.
Mount, A.S., Wheeler, A.P., Paradkar, R.P. & Snider, D. 2004. Hemocyte-
mediated shell mineralization in the eastern oyster. Science, 304, 297 300.
Nomade, S., Knight, K.B., Beutel, E.,
et al.
2007. Chronology of the Central
Atlantic Magmatic Province: Implications for the Central Atlantic rifting
processes and the TriassicJurassic biotic crisis. Palaeogeography, Palaeocli-
matology, Palaeoecology, 244, 326344.
C. KORTE ET AL.444
O’Neil, J.R., Clayton, R.N. & Mayeda, T.K. 1969. Oxygen isotope fractionation
in divalent metal carbonates. Journal of Chemical Physics, 51, 5547 5558.
Pagani, M., Pedentchouk, N., Huber, M.,
et al.
2006. Arctic hydrology during
global warming at the Palaeocene/Eocene thermal maximum. Nature, 442,
671675.
Pa
´
lfy, J. 2003. Volcanism of the Central Atlantic Magmatic Province as a potential
driving force in the end-Triassic mass extinction. In: Hames, W.E., McHone,
J.G., Renne, P. & Ruppel, C. (eds) The Central Atlantic Magmatic Province:
Insights from fragments of Pangea. American Geophysical Union, Geophysi-
cal Monograph Series, 136, 255 267.
Pa
´
lfy, J., Mortensen, J.K., Carter, E.S., Smith, P.L., Friedman, R.M. &
Tipper, H.W. 2000. Timing of the end-Triassic mass extinction: First on land,
then in the sea? Geology, 28, 39 42.
Pa
´
lfy, J., Deme
´
ny, A., Haas, J., Hete
´
nyi, M., Orchard, M.J. & Veto
˝
, I. 2001.
Carbon isotope anomaly and other geochemical changes at the Triassic
Jurassic boundary from a marine section in Hungary. Geology, 29, 1047
1050.
Pa
´
lfy, J., Deme
´
ny, A., Haas, J.,
et al.
2007. TriassicJurassic boundary events
inferred from integrated stratigraphy of the Cso˝va´r section, Hungary.
Palaeogeography, Palaeoclimatology, Palaeoecology, 244, 1133.
Price, G.D. 1999. The evidence and implications of polar ice during the Mesozoic.
Earth-Science Reviews, 48, 183210.
Richardson, L. 1905. The Rhaetic and contiguous deposits of Glamorganshire.
Quarterly Journal of the Geological Society of London, 61, 385424.
Richardson, L. 1906. On the Rhaetic and contiguous deposits of Devon and
Dorset. Proceedings of the Geologists’ Association, 14, 401409.
Richardson, L. 1911. The Rhaetic and contiguous deposits of West, Mid, and part
of East Somerset. Quarterly Journal of the Geological Society of London, 67,
174.
Rickaby, R.E.M. & Halloran, P. 2005. Cool La Nin
˜
a during the warmth of the
Pliocene? Science, 307, 19481952.
Rosenthal, Y., Field, M.P. & Sherrell, R.M. 1999. Precise determination of
element/calcium ratios in calcareous samples using sector field inductively
coupled plasma mass spectrometry. Analytical Chemistry, 71, 32483253.
Scholle, P.A. & Arthur, M.A. 1980. Carbon isotope fluctuations in Cretaceous
pelagic limestones: potential stratigraphic and petroleum exploration tool.
AAPG Bulletin, 64, 6787.
Shackleton, N.J. & Opdyke, N.D. 1973. Oxygen isotope and palaeomagnetic
stratigraphy of equatorial Pacific core V28-238: Oxygen isotope temperatures
and ice volumes on a 10
5
year and 10
6
year scale. Quaternary Research, 3,
3955.
Sheppard, T.H., Houghton, R.D. & Swan, A.R.H. 2006. Bedding and pseudo-
bedding in the Early Jurassic of Glamorgan: deposition and diagenesis of the
Blue Lias in South Wales. Proceedings of the Geologists’ Association, 117,
249264.
Self, S., Widdowson, M., Thordarson, T. & Jay, A.E. 2006. Volatile fluxes
during flood basalt eruptions and potential effects on the global environment:
A Deccan perspective. Earth and Planetary Science Letters, 248, 518 532.
Spo
¨
tl, C. & Vennemann, T.W. 2003. Continuous-flow isotope ratio mass
spectrometric analysis of carbonate minerals. Rapid Communications in Mass
Spectrometry, 17, 10041006.
Stecher, H.A., Krantz, D.E., Lord, C.J., Luther, G.W. & Bock, K.W. 1996.
Profiles of strontium and barium in Mercenaria mercenaria and Spisula
solidissima shells. Geochimica et Cosmochimica Acta, 60, 34453456.
Steuber, T. & Veizer, J. 2002. Phanerozoic record of plate tectonic control of
seawater chemistry and carbonate sedimentation. Geology, 30, 11231126.
Surge, D., Lohmann, K.C. & Dettman, D.L. 2001. Controls on isotopic
chemistry of the American oyster, Crassostrea virginica: implications for
growth patterns. Palaeogeography, Palaeoclimatology, Palaeoecology, 172,
283296.
Svensen, H., Planke, S., Malthe-Sørenssen, A., Jamtveit, B., Myklebust, R.,
Eidem, T.R. & Rey, S.S. 2004. Release of methane from a volcanic basin as
a mechanism for initial Eocene global warming. Nature, 429, 542545.
Svensen, H., Planke, S., Chevallier, L., Malthe-Sørenssen, A., Corfu, F. &
Jamtveit, B. 2007. Hydrothermal venting of greenhouse gases triggering Early
Jurassic global warming. Earth and Planetary Science Letters, 256, 554566.
Swift, A. 1995. A review of the nature and outcrop of the ‘White Lias’ facies of
the Langport Member (Penarth Group: Upper Triassic) in Britain. Proceed-
ings of the Geologists’ Association, 106, 247258.
Swift, A. & Martill, D.M. 1999. Fossils of the Rhaetian Penarth Group.
Blackwell for the Palaeontological Association, London.
Tomas
ˇ
ovy
´
ch, A. & Siblı
´
k, M. 2007. Evaluating compositional turnover of
brachiopod communities during the end-Triassic mass extinction (Northern
Calcareous Alps): Removal of dominant groups, recovery and community
reassembly. Palaeogeography, Palaeoclimatology, Palaeoecology, 244,170
200.
Trueman, A.E. 1920. The Liassic rocks of the Cardiff district. Proceedings of the
Geologists’ Association, 31, 93 107.
vander Putten, E., Dehairs, F., Keppens, E. & Baeyens, W. 2000. High
resolution distribution of trace elements in the calcite shell layer of modern
Mytilus edulis: environmental and biological controls. Geochimica et
Cosmochimica Acta, 64, 9971011.
van de Schootbrugge, B., Tremolada, F., Rosenthal, Y.,
et al.
2007. End-
Triassic calcification crisis and blooms of organic-walled ‘disaster species’.
Palaeogeography, Palaeoclimatology, Palaeoecology, 244, 126141.
Van Houten, F.B. 1969. Late Triassic Newark Group, north-central New Jersey
and adjacent New York and Pennsylvania. In: Subitzky, S. (ed.) Geology of
Selected Areas in New Jersey and Pennsylvania. Geological Society of
America and Rutgers University Press, New Brunswick, NJ, 314348.
Van Houten, F.B. 1971. Contact metamorphic mineral assemblages, Late Triassic
Newark Group, New Jersey. Contributions to Mineralogy and Petrology, 30,
114.
Veizer, J. 1983. Trace elements and isotopes in sedimentary carbonates. In:
Reeder, R.J. (ed.) Carbonates: Mineralogy and Chemistry. Mineralogical
Society of America, Reviews in Mineralogy, 11, 265 299.
Veizer, J., Bruckschen, P., Pawellek, F.,
et al.
1997a. Oxygen isotope
evolution of Phanerozoic seawater. Palaeogeography, Palaeoclimatology,
Palaeoecology, 132, 159172.
Veizer, J., Buhl, D., Diener, A.,
et al.
1997b. Strontium isotope stratigraphy:
potential resolution and event correlation. Palaeogeography, Palaeoclimatol-
ogy, Palaeoecology, 132, 6577.
Veizer, J., Ala, D., Azmy, K.,
et al.
1999.
87
Sr/
86
Sr, ä
13
Candä
18
O evolution of
Phanerozoic seawater. Chemical Geology, 161, 5988.
Ward, P.D., Haggart, J.W., Carter, E.S., Wilbur, D., Tipper, H.W. & Evans,
T. 2001. Sudden productivity collapse associated with the TriassicJurassic
boundary mass extinction. Science, 292, 11481151.
Ward, P.D., Garrison, G.H., Williford, K.H., Kring, D.A., Goodwin, D.,
Beattie, M.J. & McRoberts, C.A. 2007. The organic carbon isotopic and
paleontological record across the Triassic Jurassic boundary at the candidate
GSSP section at Ferguson Hill, Muller Canyon, Nevada, USA. Palaeogeo-
graphy, Palaeoclimatology, Palaeoecology, 244, 281 289.
Waters, R.A. & Lawrence, D. 1987. Geology of the South Wales Coalfield, Part
III, the Country around Cardiff. Geological Survey of Great Britain—England
and Wales—Memoirs: British Geological Survey—BGS Reports. Stationery
Office Books, London.
Weedon, G.P. 1986. Hemipelagic shelf sedimentation and climatic cycles: the basal
Jurassic (Blue Lias) of South Britain. Earth and Planetary Science Letters,
76, 321335.
Whittaker, A. 1978. The lithostratigraphical correlation of the uppermost Rhaetic
and lowermost Liassic strata of the W. Somerset and Glamorgan areas.
Geological Magazine, 115, 6367.
Whittaker, A. & Green, G.W. 1983. Geology of the country around Weston-
Super-Mare. Memoirs of the Geological Survey of Great Britain, Sheet 279
with parts of sheets 263 and 295 (England and Wales). HMSO, London.
Wignall, P.B. 2001a. Sedimentology of the TriassicJurassic boundary beds in
Pinhay Bay (Devon, SW England). Proceedings of the Geologists’ Associa-
tion, 112, 349360.
Wignall, P.B. 2001b. Large igneous provinces and mass extinctions. Earth-Science
Reviews, 53,133.
Williford, K.H., Ward, P.D., Garrison, G.H. & Buick, R. 2007. An extended
organic carbon-isotope record across the TriassicJurassic boundary in the
Queen Charlotte Islands, British Columbia, Canada. Palaeogeography,
Palaeoclimatology, Palaeoecology, 244, 290296.
Zachos, J., Pagani, M., Sloan, L., Thomas, E. & Billups, K. 2001. Trends,
rhythms, and aberrations in global climate 65 Ma to present. Science, 292,
686 693.
Zeebe, R.E. & Wolf-Gladrow, D.A. 2001. CO
2
in Seawater: Equilibrium,
Kinetics, Isotopes. Elsevier Oceanography Book Series, 65.
Ziegler, P.A. 1990. Geological Atlas of Western and Central Europe. Shell
Internationale Petroleum Maatschappij B.V., The Hague.
Received 8 January 2008; revised typescript accepted 15 December 2008.
Scientific editing by Ian Jarvis.
TRIASSICJURASSIC BOUNDARY 445
... Determined Mn and Fe values ranged between 6 and 207 ppm (mean 59 ppm) for Mn, and between 30 and 1115 ppm (mean 318 ppm) for Fe (Appendix). Bivalves (including Gryphaea and oysters) are typically characterized by quite variable Fe and Mn contents (e.g., Milliman, 1974;Morrison and Brand, 1986;Anderson et al., 1994;Korte et al., 2009;Korte and Hesselbo, 2011;Price and Page, 2008;Denison et al., 2003). The Gryphaea and oyster samples with high Fe (>800 ppm) and Mn (250 ppm) concentrations (Morrison and Brand, 1986;Korte et al., 2009;Denison et al., 2003) were excluded from further analysis. ...
... Bivalves (including Gryphaea and oysters) are typically characterized by quite variable Fe and Mn contents (e.g., Milliman, 1974;Morrison and Brand, 1986;Anderson et al., 1994;Korte et al., 2009;Korte and Hesselbo, 2011;Price and Page, 2008;Denison et al., 2003). The Gryphaea and oyster samples with high Fe (>800 ppm) and Mn (250 ppm) concentrations (Morrison and Brand, 1986;Korte et al., 2009;Denison et al., 2003) were excluded from further analysis. Overall, most belemnites, oysters and Gryphaea are deemed well preserved. ...
... The analysis in this study focuses exclusively on sea surface temperature (SST) data from low latitudes, obtained from sources [48][49][50][51][52] . It is important to note that surface temperature anomalies at low latitudes consistently exhibit intermediate values, indicating proximity to average values 4 . ...
Article
Full-text available
This study investigates the mechanisms underlying the varied climate changes witnessed during mass extinctions in the Phanerozoic Eon. Climate shifts during mass extinctions have manifested as either predominant global cooling or predominant warming, yet the causes behind these occurrences remain unclear. We emphasize the significance of sedimentary rock temperature in comprehending these climate shifts. Our research reveals that low-temperature heating of sulfide leads to global cooling through the release of sulfur dioxide (SO2), while intermediate-temperature heating of hydrocarbons and carbonates releases substantial carbon dioxide (CO2), contributing to global warming. High-temperature heating additionally generates SO2 from sulfate, further contributing to global cooling. Different degrees of contact heating of the host rock can lead to different dominant volatile gas emissions, crucially driving either warming or cooling. Moreover, medium to high-temperature shock-heating resulting from asteroid impacts produces soot from hydrocarbons, also contributing to global cooling. Large-scale volcanic activity and asteroid impacts are both events that heat rocks, emitting the same gases and particles, causing climate changes. The findings elucidate the critical role of heating temperature and heating time in understanding major climate changes during mass extinctions.
... Break-up of the supercontinent Pangea and the opening of the proto-Central Atlantic Ocean initiated in the Late Triassic and continued throughout the Early Jurassic (McHone 1996, Marzoli et al. 1999, Jourdan et al. 2008. The Early Jurassic is marked by transitions from cold climate modes or ice ages to supergreenhouse events and vice versa (Korte et al. 2009, Korte & Hesselbo 2011, Ruebsam & Schwark 2021, changes in global sea-level (Hallam 1997, Hesselbo et al. 2004, Hesselbo 2008, changes in Fig. 1. Map of south-western Great Britain with assigned Mesozoic basins with Early Jurassic strata (blue) and selected geological structures. ...
... Because of uncertainties regarding the numerical ages of Jones et al. (1994), we not only recalculated their ages according to the Geologic Time Scale 2020 (GTS 2020) (Gradstein et al., 2020) but also employed an alternative approach, using the cyclostratigraphic and biostratigraphic age model of Weedon et al. (2019) to establish numerical ages for the 87 Sr/ 86 Sr ratios of Lyme Regis. The I-NCIE has not been identified at Lyme Regis (Korte et al., 2009); therefore, the age of the first known bed from the Tilmanni Zone (H6) is set to 201.36 Ma, that is, approximated using the age of the TJB (Hillebrandt et al., 2013; Wotzlaw (Korte et al., 2003), squares: Lyme Regis oysters (Jones et al., 1994), orange triangles: bulk carbonate from Eiberg, purple triangles: bulk carbonate from Zlambach (the Northern Calcareous Alps) (Z. , circles: bulk carbonate from Csővár (this study). ...
Article
Full-text available
The end‐Triassic extinction (ETE) is one of the most severe biotic crises in the Phanerozoic. This event was synchronous with volcanism of the Central Atlantic Magmatic Province (CAMP), the ultimate cause of the extinction and related environmental perturbations. However, the continental weathering response to CAMP‐induced warming remains poorly constrained. Strontium isotope stratigraphy is a powerful correlation tool that can also provide insights into the changes in weathering regime, but the scarcity of ⁸⁷Sr/⁸⁶Sr data across the Triassic‐Jurassic boundary (TJB) hindered the use of this method. Here we present new high‐resolution ⁸⁷Sr/⁸⁶Sr data from bulk carbonates at Csővár, a continuous marine section that spans 2.5 Myrs across the TJB. We document a continuing decrease in ⁸⁷Sr/⁸⁶Sr ratio from the late Rhaetian to the ETE, terminated by a 300 kyr interval of a flat trend and followed by a transient increase in the early Hettangian that levels off. We suggest that the first in the series of perturbations is linked to the influx of non‐radiogenic Sr from the weathering of freshly erupted CAMP basalts, leading to a delay in the radiogenic continental weathering response. The subsequent rise in ⁸⁷Sr/⁸⁶Sr after the TJB is explained by intensified continental crustal weathering from elevated CO2 levels and reduced mantle‐derived Sr flux. Using Sr flux modeling, we also find support for such multiphase, prolonged continental weathering scenarios. Aggregating the new data set with published records employing an astrochronological age model results in a highly resolved Sr isotope reference curve for an 8.5 Myr interval around the TJB.
... onset of mass mortality [37][38][39][40][41][42][43][44][45][46][47][48], and they both culminated in brief ice ages, i.e., the latest Ordovician Hirnantian [41] and end-Devonian Hangenberg glaciations [49]. Moreover, unlike the other three Big Five mass extinctions, these two biocrises were associated with large positive carbon-isotope excursions (PCIEs; Fig. 2D-E) indicating a fundamentally different response of the global carbon cycle (i.e., net carbon burial) from that of the ECME, EPME, and ETME (whose NCIEs are indicative of net carbon release) ( Fig. 2A-C). ...
Article
Full-text available
Theory regarding the causation of mass extinctions is in need of systematization, which is the focus of this contribution. Every mass extinction has both an ultimate cause, i.e., the trigger that leads to various climato-environmental changes, and one or more proximate cause(s), i.e., the specific climato-environmental changes that result in elevated biotic mortality. With regard to ultimate causes, strong cases can be made that bolide (i.e., meteor) impacts, large igneous province (LIP) eruptions, and bioevolutionary events have each triggered one or more of the Phanerozoic Big Five mass extinctions, and that tectono-oceanic changes have triggered some second-order extinction events. Apart from bolide impacts, other astronomical triggers (e.g., solar flares, gamma bursts, and supernova explosions) remain entirely in the realm of speculation. With regard to proximate mechanisms, most extinctions are related to either carbon-release or carbon-burial processes, the former being associated with climatic warming, ocean acidification, reduced marine productivity, and lower carbonate δ13C values, and the latter with climatic cooling, increased marine productivity, and higher carbonate δ13C values. Environmental parameters such as marine redox conditions and terrestrial weathering intensity do not show consistent relationships to carbon-cycle changes. In this context, mass extinction causation can be usefully classified using a matrix of ultimate and proximate factors. Among the Big Five mass extinctions, the end-Cretaceous biocrisis is an example of a bolide-triggered carbon-release event, the end-Permian and end-Triassic biocrises of LIP-triggered carbon-release events, and the Late Ordovician and Late Devonian biocrises of bioevolution-triggered carbon-burial events. Whereas the bolide-impact and LIP-eruption mechanisms appear to invariably cause carbon release, bioevolutionary triggers can result in variable carbon-cycle changes, e.g., carbon burial during the Late Ordovician and Late Devonian events, carbon release associated with modern anthropogenic climate warming, and little to no carbon-cycle impact due to certain types of ecosystem change (e.g., the advent of the first predators around the end-Ediacaran; the appearance of Paleolithic human hunters in Australasia and the Americas). Broadly speaking, studies of mass extinction causation have suffered from insufficiently critical thinking—an impartial survey of the extant evidence shows that (i) hypotheses of a common ultimate cause (e.g., bolide impacts or LIP eruptions) for all Big Five mass extinctions are suspect given manifest differences in patterns of environmental and biotic change among them; (ii) the Late Ordovician and Late Devonian events were associated with carbon burial and long-term climatic cooling, i.e., changes that are inconsistent with a bolide-impact or LIP-eruption mechanism; and (iii) claims of periodicity in Phanerozoic mass extinctions depended critically on the now-disproven idea that they shared a common extrinsic trigger (i.e., bolide impacts).
... We assumed that the local SST values are representative of intermediate values in the oceans of each age, which not is completely correct. For example, Korte et al. (2009) describe a "temperature increase" of 7-14 °C in the Rhaetian and about 12-22 °C in the Hettangian. However, we compared the Rhaetian and Hettangian taxa as a single event based on the Paleobiology Database. ...
Article
Based on the global occurrence dataset, the shift in taxonomic and functional diversity of bivalves at the Triassic/Jurassic transition was examined herein. There is a noticeable decline in diversity at many taxonomic levels (generic, family, and order) along the Triassic/Jurassic boundary. Test changes in the functional diversity (e.g., life habits, mobility levels, and feeding mode) revealed that the percentage of mobile exceeded stationary taxa after the end of the Triassic crisis, while no major changes were observed in the life habit or feeding mode. By the Sinemurian, diversity reached the pre-extinction levels. A significant difference was also found between survivors’ longevity and extinct taxa, where the Early Jurassic (Hettangian) fauna have a longer duration relative to those that became extinct. The Triassic/Jurassic boundary is marked by a marked sea-level fall and a decrease in the mean Sea Surface Temperature (SST), which is associated with increasing siliciclastic and decreasing carbonate rocks. The latter may also point to ocean acidification at the Triassic/Jurassic boundary. The geographic range size of bivalves is slightly changed by the end of the Triassic, where the taxa are slightly characterized by narrower ranges. Hence, the geographic range size, the result of ecophysiology, plays a major role in determining the extinction risk. The difference in the magnitude of the diversity loss (i.e., taxonomically vs. functionally) indicated that the shallower marine habitat destruction resulting from the sea-level fall is the primary cause of the Triassic/Jurassic mass extinction.
... The Neg-II (initial CIE) has been suggested to refl ect the trigger for the end-Triassic mass extinction event (Hesselbo et al., 2002(Hesselbo et al., , 2004Korte et al., 2009;. Based on an 8.5‰ negative CIE in n-alkane records contemporaneous with the initial CIE (Neg-II), proposed that massive release of methane from clathrates triggered the terrestrial and marine turnovers. ...
Article
The Upper Triassic–Lower Jurassic succession in the Danish Basin is penetrated by many deep wells that were drilled during former hydrocarbon exploration campaignes but it is today targeted for geothermal energy and storage of CO2. In the Stenlille salt dome on Sjaelland sandstones of the Gassum Formation, sealed by the overlying Fjerritslev Formation mudstones, have been used for decades as a seasonal storage for natural gas. With its comprehensive dataset of seismics, geophysical well logs and conventional core data from twenty wells, the Stenlille succession serves as a model for other salt domes currently evaluated as potential CO2 storage sites in the basins. Over the last decade the cored Triassic–Jurassic boundary (TJB) succession has contributed to the understanding of environmental and palynological events during the end-Triassic mass extinction. Core, sidewall core and cuttings samples from several of the closely situated Stenlille wells are here used for establishment of a high-resolution palynostratigraphic zonation scheme covering the entire Rhaetian to Sinemurian succession by integrating new analyses with previously published data. The palynological data set have allowed recognition of nine formally described spore-pollen zones of which eight are new, while two previously described dinoflagellate cysts zones are subdivided into 3 informal subzones each. The palynological zonation is integrated with a sequence stratigraphic framework and will form the basis for the dating of future well sections in the Danish Basin as well as other basins and for correlation to outcrops. The large palynological dataset further shows that the vegetation around the Danish Basin was remarkably stable during the early to middle Rhaetian, but that events related to the emplacement of the Central Atlantic Magmatic Province accelerated ecosystem changes during c. 175 kyrs in the late Rhaetian and earliest Hettangian including ∼25 kyrs of successional recovery before the terrestrial ecosystem had again stabilized.
Article
Storm deposits or tempestites are event sequences formed by storms, requiring at least a water temperature of 26.5°C. While inland lakes are unlikely to form storm deposits because of their limited width and water temperature. The Upper Triassic Xujiahe Formation in the Sichuan Basin is a set of coal-bearing, clastic sequences with dominant sedimentary facies varying from braided river delta to lacustrine settings, with storm deposits widely reported. In the Zilanba of Guanyuan area, in situ tree trunks on a palaeosol surface in Member V of the Xujiahe Formation provide new evidence of a storm event. Six fallen-down directions of nine in situ tree trunks were predominant in the NW direction, contrary to the palaeocurrent direction of the underlying strata, suggesting that the southeasterlies prevailed during the end-Triassic in the northern Sichuan Basin. Massive mud clasts were frequently recorded in sandstones of the Xujiahe Formation, as well as in the Xindianzi section. These mud clasts showed a rip-up or a plastic deformation with upside-down V-shapes, were capped on an erosional surface, showed no transport traces and were therefore interpreted as a storm lag deposit. The megamonsoonal climate prevailed during the Late Triassic, although the megamonsoons themselves could not generate a storm deposition in the Xujiahe Formation due to its low maximum surface wind speed. The driving mechanism for generating storm deposits in the Xujiahe Formation is suggested to be tropical cyclones over the Tethys Ocean moving eastward, further landfalling on the western margin of the Sichuan Basin. Statistics of storm events in the circum-Tethys region show a widespread storm surge in low latitudes during the end-Triassic. The storm deposits at the top of the Xujiahe Formation represent a sedimentary response to the end-Triassic hyperthermal event.
Article
Full-text available
Frequent wildfires associated with emplacement of the Central Atlantic Magmatic Province (CAMP) are thought to have been important drivers of two significant changes in terrestrial plant communities and diversity during the Triassic-Jurassic Mass Extinction (TJME, ca. 201.51 Ma). However, it remains to be investigated whether these two changes are potentially related to different wildfire types. To better understand this relationship, we used a new method to reanalyze fossil pollen and spores across the Triassic-Jurassic transition in the Jiyuan Basin from the North China Plate. Results show that two peaks in wildfire frequency experienced different types of wildfires, with each linked to significant changes in plant communities and diversity losses. In the first wildfire peak, canopy fires dominated and are accompanied by significant losses of canopy forming plants, while in the second wildfire peak, ground cover fires dominated accompanied by significant losses of ground cover plants. Changes in atmospheric humidity conditions were an important control on the two different wildfire peaks. Relatively humid climatic conditions corresponded to the prevalence of canopy fires and hindered the spread and development of ground cover fires in wet surface conditions. Conversely, relatively arid climatic conditions corresponded with the prevalence of ground cover fires in dry surface environments. Our results provide a potential relationship between terrestrial plant communities and wildfire types, which is important to further understanding of terrestrial environmental and floral changes driven by Large Igneous Provinces.
Article
Full-text available
Paleotemperatures have been widely deduced from skeletal 18O/16O ratios, but these are also dependent on salinity. Without an independent measure of salinity, 18O/16O ratios cannot provide accurate data on past temperature and climate. We grew marine mollusks (bivalves) in the field while real-time data on local seawater temperature and chemistry (Mg, Ca, δ18O, and salinity) were gathered. Here we show that for Mytilus trossulus, skeletal Mg/Ca ratios provide an accurate measure of temperature and that weekly sea-surface temperatures may be estimated with an apparent accuracy of approximately ±1.5°C. Thus, with analyses of both Mg/Ca and δ18O from the same specimen, it is possible to determine seawater temperature and δ18O.
Book
Focusing on the basic principles of mineral formation by organisms, this comprehensive volume explores questions that relate to a wide variety of fields, from biology and biochemistry, to paleontology, geology, and medical research. Preserved fossils are used to date geological deposits and archaeological artifacts. Materials scientists investigate mineralized tissues to determine the design principles used by organisms to form strong materials. Many medical problems are also associated with normal and pathological mineralization. Lowenstam, the pioneer researcher in biomineralization, and Weiner discuss the basic principles of mineral formation by organisms and compare various mineralization processes. Reference tables listing all known cases in which organisms form minerals are included.
Chapter
In sections through Lower Toarcian oil shales, parallel variations in the 13-C/12-C ratios of carbonate and organic matter point to pronounced changes in the oxygen level of the water column. Fauna and flora, fossil preservation, paleogeography, and certain sedimentological features support this interpretation and justify the distinction of three lithologicisotopic facies types.Carbon isotopes might be generally useful in tracing the extent of “stagnation” during deposition of bituminous sediments. Although much lower in δ 13 C than comparable Recent marine plankton, the bulk of the organic matter (as well as the oil shale carbonate) in the Posidonia Shales is ultimately derived from phytoplankton which inhabited the oxygenated surface layer of the ocean. Diagenetic processes such as oxidation, cementation, dolomitisation, isotope exchange, impregnation and migration have led to partial redistribution of carbon and oxygen isotopes within the sediment-pore water system.
Article
This is a systematic review of the major mass extinctions in the history of life. It covers all groups of organisms - plant, animal, terrestrial, and marine - that have become extinct alongside the geological and sedimentological evidence for environmental changes during the biotic crises. All proposed extinction mechanisms - climate change, meteorite impact, volcanisms - are critically assessed. In this text the demise of the dinosaurs is put into the proper context of other extinction events. This book is intended for undergraduates in Europe and graduate students in the US, studying geology, palaeontology, or evolutionary biology, and their teachers. It should also be of interest to research scientists in adjacent subjects.