You are on page 1of 16

Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

A New Zealand record of sea level rise and environmental change during the
Paleocene–Eocene Thermal Maximum
Luke Handley a,⁎, Erica M. Crouch b, Richard D. Pancost a
a
Organic Geochemistry Unit, Bristol Biogeochemistry Research Centre, School of Chemistry, University of Bristol, Cantock's Close, Bristol BS8 1TS, UK
b
GNS Science, PO Box 30368, Lower Hutt 5040, New Zealand

a r t i c l e i n f o a b s t r a c t

Article history: The global warming associated with the Paleocene–Eocene Thermal Maximum (PETM) ca. 55.5 Myr ago is the
Received 14 May 2010 most dramatic identified short-term temperature increase of the Cenozoic. One direct consequence of a
Received in revised form 2 March 2011 warming world is a rise in sea level, due primarily to the thermal expansion of water as oceans warmed. The
Accepted 7 March 2011
Kumara-2 core, South Island, New Zealand, spans the Paleocene/Eocene transition and provides a rare
Available online 11 March 2011
southern hemisphere continental margin record of the PETM. Lithology, palynology and compound-specific
Keywords:
stable isotope compositions of higher plant leaf wax n-alkanes reveal a 4.85 m PETM and a carbon isotope
Carbon isotope excursion (CIE) of ca. − 4.5‰, larger than the − 2.5 to − 3.5‰ CIEs generally recorded by deep sea
Hydrogen isotope foraminifera. There is a shift from a terrestrial to a marine, potentially anoxic, sedimentary depositional
Plant microfossil assemblage environment at the base of the PETM, interpreted as being the result of a local sea level rise. Coincident with
Palynology the onset of the CIE is the appearance of pollen associated with thermophilic conditions and the development
Organic geochemistry of Nypa mangrove swamps. Moreover, there is a reorganisation of the angiosperm pollen assemblage during
the PETM, and an initial increase in fern spores and decrease in gymnosperms. Crucially, all of these changes
occur below the horizon characterised by the most negative δ13C values, suggesting that: 1) the recorded
negative excursion of 4.5‰ may indeed reflect the shift in atmospheric CO2 isotopic composition; and 2) that
the large input of 13C-depleted carbon into the ocean–atmosphere system was not geologically instantaneous,
with at least some of the added carbon lagging warming, sea level rise and vegetation change. Furthermore,
compound-specific hydrogen isotope analyses show a large degree of variability both directly before and
during the CIE, suggesting that the PETM in New Zealand was characterised by complex and transient changes
in the hydrological regime, similar to those reported in North America, the Arctic and Eastern Africa. Thus, the
new PETM record from Kumara-2 reveals local climatic and biotic responses, driven by a combination of
global warming and the consequential change in local depositional environment induced by sea level rise.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction modelling studies suggest that such an increase cannot alone account
for the inferred warming such that other unidentified processes and/or
The Paleocene–Eocene Thermal Maximum (PETM) is one of the most feedbacks must have also contributed to the rise in global temperatures
abrupt climatic events of the Paleogene, characterised by an increase in (Zeebe et al., 2009).
global temperatures that remained elevated for ca. 170 kyr (Röhl et al., The source of this carbon dioxide remains uncertain, but evidence for
2007). This temperature increase has been estimated to be 4–6 °C for the a massive injection of 13C-depleted carbon into the ocean–atmosphere
deep ocean (Kennett and Stott, 1991; Bralower et al., 1995; Tripati and reservoir comes from the presence of a negative carbon isotope
Elderfield, 2005), 5 °C for tropical sea surface waters (Bralower et al., excursion (CIE) recognised in organic and inorganic carbon. This
1995; Thomas et al., 1999; Zachos et al., 2003), 5–8 °C for mid to high decrease in δ13C values is recorded in marine carbonates (Shackleton,
latitude surface waters (Kennett and Stott, 1991; Zachos et al., 2003; 1986; Kennett and Stott, 1991; Bains et al., 1999; Thomas et al., 2002;
Sluijs et al., 2006; Zachos et al., 2006) and 8 °C for Arctic mean annual air Zachos et al., 2003; Zachos et al., 2006), in both marine (Bolle et al., 2000;
temperatures (Weijers et al., 2007). The PETM global warming period Sluijs et al., 2006) and terrestrial total organic carbon (Magioncalda et al.,
was most likely driven by a dramatic increase in atmospheric pCO2 (e.g. 2004; Collinson et al., 2007), in paleosol carbonates (Koch et al., 1992;
Pagani et al., 2006a; Panchuk et al., 2008), although some recent Koch et al., 1995; Bowen et al., 2002; Schmitz and Pujalte, 2003) and in
individual terrestrial biomarkers (Pagani et al., 2006b; Smith et al., 2007;
Handley et al., 2008).
⁎ Corresponding author at: address: NIOZ Royal Netherlands Institute for Sea
Research, Department of Marine Organic Biogeochemistry, PO Box 59, 1790 AB Den
Sea level change may have been another consequence of PETM
Burg, Texel, The Netherlands. Tel.: +31 222369408. warming. Although Earth in the Early Paleogene is generally believed to
E-mail address: luke.handley@nioz.nl (L. Handley). have been ice-free, a number of recent studies (Speijer and Wagner,

0031-0182/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2011.03.001
186 L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200

2002; Miller et al., 2005; Hollis et al., 2009) suggest the presence of atmosphere reservoir and whether this large input was geologically
ephemeral small ice sheets in Antarctica, and melting of these ice sheets instantaneous or not, as has been suggested from some recent studies
prior to or during the PETM would result in sea level rise. Moreover, a (e.g. Sluijs et al., 2007; Sluijs et al., 2008a).
global temperature increase of 5 °C would translate into a eustatic rise of
3–5 m through thermal expansion of seawater (using the temperature– 2. Site and sample description
volume relationship of seawater of 1.9 × 10− 4 m3/°C). In coastal regions,
a sea level rise would have a direct impact on both marine and terrestrial The Kumara-2 exploration well was drilled in 1985 and is located
biota, along with the source of organic matter delivered to sediments. in central Westland, South Island (NZTopo50 map grid reference
Sluijs et al. (2008a) recently presented evidence for a sea level rise BU19/519 849), ca. 20 km south of the Greymouth township (Fig. 1).
associated with the PETM. Examining sites from the New Jersey Shelf, The well was drilled to a total depth of 1756 m, with several intervals
North Sea, Arctic Ocean and New Zealand, they observed palynological throughout the well being cored. A core interval was taken from 1753
and organic biomarker changes that indicate near-shore marine to 1197 m below the surface (mbs), and we have examined sediment
sections became located further from the coast during the PETM. The from this interval, spanning 1750 to 1710 mbs (Fig. 2). The lower
collective results from their study point towards a eustatic rise with a 40 m of core incorporates Paleocene and Eocene sediments of the
maximum flooding surface during the early stages of the PETM, possibly Brunner and Paparoa coal measures (Carter et al., 1986; Raine, 1986).
initiated prior to the associated negative CIE (Sluijs et al., 2008a). Terrestrial sediments are predominant in the lower 40 m of core and
Documenting the magnitude of the CIE and the impact of warming on consist of a mixture of sandstone, shales, mudstones and coals (Fig. 2),
local depositional environments is crucial for our understanding of the with the depositional environment interpreted as backswamp, over-
quantity of carbon released and the impacts of global warming. The bank and channel fill deposits (Carter et al., 1986). However, marine
Kumara-2 core in New Zealand (Fig. 1) spans the Late Paleocene–Early sediments occur from 1739.1 to 1734.25 mbs (Carter et al., 1986).
Eocene, including the PETM; unlike other New Zealand marine records
deposited in an outer shelf-upper slope setting (Kaiho et al., 1996; 3. Methods
Hancock et al., 2003; Hollis et al., 2005) the Kumara-2 core represents one
of the few known Southern Hemisphere continental margin records that 3.1. Palynological analyses
span this time interval. The sequence is characterised by a transient
switch, during the PETM, from a terrestrial to a shallow marine 30 samples, from 1749.98 to 1710 mbs, were processed for
depositional setting (Sluijs et al., 2008a). The Kumara-2 core contains palynological analysis. Samples were collected from mudstone, shale,
well preserved organic matter, which has allowed us to complete a multi- coal and fine sandstone sediments. Between 6 and 40 g of sediment were
proxy study of biomarkers and palynomorph assemblages across the crushed, dried in an oven and treated with hot 10% HCl and 50% HF for
PETM to provide information on the depositional environment and carbonate and siliceous removal, respectively. Oxidation was completed
associated local effects of abrupt global warming. Specifically, the CIE is and due to the variety of sediment types oxidation times ranged from 15
documented using n-alkane compound specific δ13C values, to provide a to 70 min. Samples were then sieved over a 6 μm mesh to remove small
new record with which to consider the magnitude of carbon release. organic and mineral particles. To break up clumps of residue, samples
Changes in depositional environment, including changes in organic were placed in an ultrasonic bath for a maximum of 3 min during the
matter input, and changes in the plant community structure are evaluated sieving process. Well-mixed representative fractions of the N6 μm
using organic biomarkers and palynomorphs. We use compound-specific residue were mounted on glass slides using a glycerine jelly medium
hydrogen isotopes of higher plant leaf wax n-alkanes to constrain and analysed. The quantitative examination involved a minimum count
whether environmental changes included variations in the hydrological of 200 spores and pollen and, where possible, a minimum of 100
cycle and local humidity, as recorded across the PETM at northern mid dinoflagellate cysts. Slides were also scanned for additional taxa. In
latitudes (Smith et al., 2007) and the Arctic (Pagani et al., 2006b). addition, a count of 100 palynomorphs was completed to calculate the
Moreover, we compare the timing of these changes to that of the n-alkane percentage of terrestrial and marine palynomorphs. Three samples,
recorded CIE in order to evaluate whether environmental change 1733.97, 1725.15 and 1722.96 mbs, contained very sparse palynomorph
preceded or lagged the input of 13C-depleted carbon into the ocean– assemblages and are not discussed further.

Fig. 1. Location of the Kumara-2 exploration well in the South Island, New Zealand.
L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200 187

Fig. 2. Lithological column, n-alkane δ13C records, relative abundance of marine palynomorph and the dinoflagellate genus Apectodinium, total organic carbon (TOC) content, and (Pr + Ph)
concentrations for the Kumara-2 core in New Zealand across the Paleocene/Eocene boundary. (A) Compound specific δ13C values of terrestrial higher plant derived n-alkanes. The error bars
represent one standard deviation based on duplicate analyses; (B) Relative abundance of marine palynomorphs, as a percentage of the total palynomorph assemblage, and Apectodinium
relative abundance as a percentage of the total dinocyst count; (C) TOC content expressed as % of total freeze-dried sediment weight; (D) Total TOC-normalised pristane and phytane (Pr + Ph)
concentrations, expressed as μg per g of sedimentary organic carbon. Sampling horizons are highlighted on the lithological column; Mst: mudstone; Sst: sandstone. The shaded area highlights
the PETM CIE. The arrows and horizontal dashed lines indicate the first occurrences of the angiosperm taxa Cupanieidites orthoteichus and Spinizonocolpites prominatus (☆), which mark the
base of New Zealand miospore zone PM3b and base of the Eocene, and the most abundant occurrence of the dinoflagellate genus Apectodinium (✻).

3.2. Organic and biomarker analyses detector (FID) and fitted with a Chrompack fused silica capillary column
(50 m × 0.32 mm i.d.) coated with a CP Sil-5CB stationary phase
3.2.1. Bulk organic analyses (dimethylpolysiloxane equivalent, 0.12 μm film thickness). GC–MS
28 samples were gently washed with methanol (MeOH) and analysis was performed on a Thermoquest Finnigan Trace GC interfaced
powdered with a Retsch PM100 Ball Mill. Total organic carbon, inorganic with a Thermoquest Finnigan Trace MS operating with an electron
carbon, sulphur, and total nitrogen were measured as a mass% of the total ionisation source at 70 eV and scanning over m/z ranges of 50 to 850 Da.
sediment using a Carlo Erba EA 1108 and a Coulomat 702 (Strohlein). The GC was fitted with a fused silica capillary column (50 m × 0.32 mm i.
d.) coated with a ZB1 stationary phase (dimethylpolysiloxane equiva-
3.2.2. Lipid biomarker extraction and fractionation lent, 0.12 μm film thickness). For both GC and GC–MS, 1 μl of sample was
For biomarker analyses, the powdered samples were extracted via injected at 50 °C using an on-column injector. The temperature was
Soxhlet apparatus for 24 h using dichloromethane (DCM)/MeOH (2:1 increased to 130 °C with an initial ramp of 20 °C/min, then to 300 °C at
v/v) as the organic solvent. The total lipid extracts were separated into 4 °C/min, followed by an isothermal for 20 min.
two fractions on an aminopropyl solid phase extraction column by
elution with DCM/iso-propanol (3:1 v/v; neutral fraction) and 2% (by 3.2.4. Gas chromatography–combustion–isotope ratio mass
volume) acetic acid in diethyl ether (acid fraction). The neutral spectrometry
fraction was further split using a column packed with (activated) Compound-specific stable carbon isotope ratios were determined for
alumina by elution with hexane (saturated hydrocarbon fraction), urea-adducted n-alkanes by gas chromatography–combustion–isotope
hexane/DCM (9:1 v/v; aromatic fraction) and DCM/MeOH (1:2 v/v; ratio mass spectrometry (GC–C–IRMS) using a Varian 3400 GC coupled
polar fraction). n-alkanes were selectively adducted from the to a Finnigan MAT Delta-S IRMS via a modified Finnigan MAT Type I
saturated hydrocarbon fraction via urea adduction by adding 200 μl combustion interface; the combustion furnace comprised Cu and Pt
of urea (saturated in methanol) and 200 μl of acetone, followed by wires (0.1 mm o.d.) in an alumina reactor (0.5 mm i.d.). GC column and
200 μl of hexane. The resulting urea crystals were washed with conditions were the same as described above for GC–MS. Values,
hexane to remove non-adducts and then dissolved in 500 μl MeOH measured in duplicate, are reported in standard per mil notation relative
and 500 μl double distilled water. The adductable hydrocarbons were to Vienna PeeDee Belemnite (VPDB); analytical accuracy, on the basis of
then extracted with three consecutive hexane washes. replicate analysis of standards, was typically ±0.3‰ and precision, as
represented by 1 standard deviation, was generally b0.3‰ (see
3.2.3. Gas chromatography and gas chromatography–mass spectrometry supplementary material).
Individual compounds were identified and quantified relative to
internal standards (5α-androstane, apolar fraction; hexadecan-2-ol, 3.2.5. Gas chromatography–thermal conversion–isotope ratio mass
polar fraction; n-C19 alkane, acid fraction) using gas chromatography spectrometry
(GC) and gas chromatography–mass spectrometry (GC–MS). Typical Compound-specific stable hydrogen isotope ratios were determined
error in absolute quantification was ±20%. GC analyses were performed for urea-adducted n-alkanes by gas chromatography–thermal conver-
on a CarloErba Gas Chromatograph equipped with a flame ionisation sion–isotope ratio mass spectrometry (GC–TC–IRMS) using an Agilent
188 L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200

6890 gas chromatograph equipped with an Agilent split/splitless injector 4.2. Palynology
(splitless mode; 300 °C; purge time=2 min) coupled to a ThermoQuest
Finnigan DeltaPlus XL isotope ratio mass spectrometer via a thermal Palynomorph assemblages are well-preserved and diverse through-
conversion reactor (300× 0.5 mm i.d.; Al2O3; 1450 °C) and ThermoQuest out most of the interval studied, although they are rare to absent from a
Finnigan GC Combustion III interface. An H+3 factor, to correct for H3 ions, ca. 12.5 m interval within the 16.05 m hard sandstone-dominated unit
was calculated daily and was consistently lower than 5 ppm V− 1. The GC that begins at 1734.25 mbs. Only terrestrial palynomorphs were
column was the same as described above, but the temperature recorded in the studied core, except for an interval between 1739 and
programme comprised an initial isothermal of 1 min at 40 °C followed 1733.43 mbs where marine dinocysts are also recorded (Fig. 2). Within
by a ramp of 10 °C/min to 300 °C and a final isothermal for 13 min. The the marine mudstone unit, marine palynomorphs comprise up to 31% (at
analytical accuracy and precision of the system were monitored using a 1735.92 mbs) of the total palynomorph assemblage (Sluijs et al., 2008a).
standard suite of 15 n-alkanes (C16–C30; Mixture B, Arndt Schimmel- Pollen of angiosperm taxa associated with thermophilic conditions,
mann, Biogeochemical Laboratories, Indiana University), which was such as Cupanieidites orthoteichus and Spinizonocolpites prominatus (Nypa
analysed before and after each sample. The root-mean-square error mangrove), are first recorded at 1737.97 mbs (Figs. 2 and 3). These
(RMS) for hydrogen isotopic measurements of the standard mixture was species are characteristic of New Zealand miospore zone PM3b (Raine,
calculated for each sample analysed to monitor analytical accuracy and 1984), which is regionally correlated to the base of the New Zealand
was typically b5‰. Analytical precision, as represented by 1 standard Waipawan Stage and equivalent to the base of the Eocene (Hancock et al.,
deviation, was generally b2‰ (see supplementary material). δD values, 2003; Cooper, 2004). In addition, the dinocyst genus Apectodinium is
based on duplicate analyses and reported in standard per mil notation abundant within the marine mudstone (Fig. 2), in agreement with other
relative to Vienna Standard Mean Ocean Water (VSMOW), were coastal marine PETM sequences in New Zealand and other regions of the
calculated against a calibrated H2 gas. world (e.g. Crouch et al., 2001; Sluijs et al., 2008a).
In the lower part of the core (1749.98–1739.21 mbs), angiosperm
4. Results taxa are the most abundant terrestrial palynomorphs, with percentages
ranging from 42 to 60% of the assemblage (Fig. 3). Gymnosperm taxa,
Based on lithological, geochemical and palynological examination, such as Phyllocladidites mawsonii, range from 16 to 44% and fern spores
the onset of the PETM is recorded in the Kumara-2 core between constitute 10–39% of the plant microfossil assemblage. At 1737.97 mbs,
samples at 1739.76 and 1737.97 mbs, and is placed at the base of the concurrent with the first occurrence of thermophilic angiosperm
mudstone interval that begins at 1739.1 mbs (Fig. 2). The n-alkane species, there is a change in the overall spore/pollen assemblage that
compound specific δ13C record clearly shows the shift to 13C-depleted continues in the overlying ca. 3.5 m section. The relative abundance of
values, which are characteristic of the PETM negative carbon isotope gymnosperms declines to 10% in the lower part of the 3.5 m section and
excursion, and the mudstone interval coincides with the first evidence then begins to increase at the top of this interval (Fig. 3). Conversely,
of marine palynomorphs and the dinoflagellate cyst (dinocyst) genus fern spores show an increase in relative abundance (to 45%) in the lower
Apectodinium. The top of the PETM is placed at 1734.25 mbs, at the part and then decrease in the upper part. Angiosperm taxa slightly
unconformity between the mudstone unit and the overlying sandstone- decrease in the lower part of the interval (to 41%) but then become more
dominated lithology (Fig. 2). Therefore, the PETM interval recorded in abundant, up to 60% of the assemblage, in the upper part (Fig. 3). Within
the Kumara-2 core is 4.85 m thick, here defined as spanning from 1739.1 the angiosperm assemblage, there is a notable change in composition
to 1734.25 mbs (Fig. 2). Lithological, geochemical and palynological through the 1737.97–1734.5 mbs interval, including a decrease in
variations associated with the PETM, and pre- and post-PETM, are Proteaceae and Triorites species and the common presence of several
discussed in more detail below. species that first occur within this interval — Malvacipollis, Myrtaceae,
Cupanieidites orthoteichus and Spinizonocolpites prominatus (Fig. 3).
4.1. Lithology and bulk geochemistry In samples taken at 1733.43 and 1732.74 mbs, the plant micro-
fossil assemblage returns to a composition similar to that seen in the
From the base of the examined core, 1749.98 mbs, to 1739.1 mbs, the latest Paleocene part of the core (1749.98–1739.21 mbs), with more
sequence is predominantly composed of a non-calcareous shale and abundant angiosperms (up to 63%), an increase in gymnosperms
mudstone, interspersed with coal lenses and layers, and silty and fine (30%) and lower abundances of fern spores (down to 5%). Moreover,
sandstone (Fig. 2), which is consistent with a terrestrial depositional the relative abundance of angiosperm taxa associated with thermo-
environment (Carter et al., 1986). A 10 cm coal layer is present at the top philic conditions, particularly Malvacipollis and Spinizonocolpites
of this interval, from 1739.2 to 1739.1 mbs (Fig. 2), and palynomorph prominatus, is lower than in underlying interval from 1737.97 to
assemblages indicate a terrestrial depositional environment. At 1734.5 mbs (Fig. 3).
1739.1 mbs, there is a shift to a dark non-calcareous mudstone dominated In the upper part of the core, from 1721.54 to 1710 mbs,
lithology with common pyrite, which persists as the dominant facies for angiosperm taxa become more abundant, ranging from 61 to 80% of
the following 4.85 m upsection (Fig. 2). The first marine dinocysts are the spore/pollen assemblage. The abundance of gymnosperm pollen
recorded at 1737.97 mbs and are present throughout the mudstone unit. remains relatively stable (17–33%) and fern spores comprise less than
From 1734.25 to 1718.2 mbs is a hard sandstone dominated interval that 11% of the assemblage. The most prominent vegetation change
contains carbonaceous laminae and mudstone clasts in the upper part. through the entire studied interval occurs at 1712.98 mbs and is
Above 1718.2 mbs to the top of the studied section at 1710 mbs, the marked by a significant increase in Casuarina (Myricipites harrisii)
sediment consists of interbedded sandstone, shale, mudstone and coal pollen (Fig. 3). This change in the plant microfossil record has been
layers and lenses (Fig. 2), interpreted to represent a terrestrial consistently recognised throughout New Zealand in the Early Eocene
depositional environment similar to that in the lower part of the section and marks a notable long-term shift in vegetation records (e.g. Raine,
(Carter et al., 1986). 1984; Pocknall, 1990; Crouch et al., 2003).
Total organic carbon (TOC) content varies throughout, ranging from In the lower part of the marine mudstone interval (1737.97–
0.4 to 36% (Fig. 2). Sediments with very high organic carbon content 1737.02 mbs), dinocysts of the genus Spiniferites and Fibrocysta type
contain minor to large amounts of coal (e.g., sample at 1743.26 mbs). In species are most common. The dinocyst assemblage is dominated by
the dark non-calcareous mudstone layer, between 1739.1 and Apectodinium (ca. 54%) in the interval with the highest percentage of
1734.25 mbs, TOC contents are uniformly low (b1%) and are much marine palynomorphs, from 1735.92 to 1735.5 mbs (Fig. 2). The upper
lower than the rest of the section, which suggests a significant change in horizon of the marine mudstone interval (1735.04–1734.5 mbs)
depositional environment from over- and under-lying strata. comprises common Operculodinium and Kenleyia species.
L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200 189

Fig. 3. Relative abundance of selected species and groups of plant microfossils, major plant microfossil groups (fern spores, gymnosperms and angiosperms), along with the n-alkane δ13C
record from the Kumara-2 core. Plant microfossil abundances are shown as a percentage of the total spore and pollen count. The shaded area highlights the PETM CIE.

4.3. Biomarkers mudstone interval, before returning to near pre-excursion initial values
above 1733.97 mbs (Fig. 5).
Functionalised compounds are only found in trace amounts The n-alkane terrigenous to aquatic ratio (TAR = (C27 + C29 + C31)/
throughout, suggesting the sequence is relatively thermally mature, (C15 + C17 + C19)) varies throughout the section (Fig. 5). At the base of
likely near the onset of the oil generation window. This is confirmed the section TARs are generally high, being greater than 7.6 and with a
by the distribution of steranes and hopanes, as discussed below. maximum value of 14.1. Ratios then decrease from 1745.88 to
Consequently, only the saturated hydrocarbon fraction contained 1743.26 mbs due to a decrease in the concentration of HMW n-
compounds detectable in appreciable quantities for analysis using alkanes, before increasing again to values greater than ca. 8 between
GC–MS and GC–C–IRMS. 1741.75 and 1739.76 mbs. At 1737.97 mbs, ratios decrease to 1.3 and
remain lower than 4.2 up to 1733.97 mbs, with a slight increase at the
4.3.1. n-alkanes upper end of this interval (Fig. 5). The initial decrease is caused by
The saturated fraction is dominated by a homologous n-alkane both a decrease in the concentration of long chain n-alkanes but also
series; all samples contain n-C17 to n-C32 alkanes, and some have an increase in the concentration of the short chain homologues. Above
measurable quantities of n-C14 to n-C16 and n-C33 to n-C41 (Fig. 4). The the marine mudstone interval, TAR ratios increase once more to an
series is dominated by high molecular weight (HMW) odd-carbon- average value of 5, with a peak value of 10 at 1717.36 mbs (Fig. 5).
number homologues, a signature typical of terrestrial higher plant leaf
wax n-alkanes (Eglinton et al., 1962; Eglinton and Hamilton, 1967; 4.3.2. n-alkane compound-specific carbon isotopes
Kolattukudy, 1976). However, the odd-over-even predominance (OEP), δ13C values were measured for C27, C29 and C31 n-alkanes, all
an expression of the preference for odd carbon number n-alkanes derived primarily from terrestrial higher plant leaf waxes (Eglinton
compared to their even carbon number homologues (Scalan and Smith, and Hamilton, 1967). Throughout most of the section there is an offset
1970; OEP = (Cn − 2 + 6 × Cn + Cn + 2) / (4 × Cn − 1 + 4 × Cn + 1), with n in δ13C values between the 3 n-alkanes: C31 is the most 13C-depleted
being an odd integer) ranges from 0.6 to 2 (Fig. 5), which is markedly in all but one sample, with values typically ca. 1‰ lower than C29,
lower than that found in fresh leaf waxes (Collister et al., 1994b) and which in turn has values typically ca. 0.5‰ more negative than C27
suggests a potential diagenetic or catagenetic overprint of the original (Fig. 2). Despite this small offset in isotope composition, the δ13C
biological signal. The OEP values for n = 27/29/31 all have identical values of all three n-alkanes describe almost identical trends
trends; throughout most of the section, OEP values are N1, which throughout the core.
reflects a predominantly terrestrial source for the n-alkanes. At Below 1737.97 mbs, values fluctuate between − 29.4 and − 31.1‰
1737.97 mbs, coincident with the presence of marine dinocysts, OEP for C31, −28.2 and − 29.9‰ for C29 and −27.5 and −29.1‰ for C27
values decrease dramatically to a minimum of 0.7 and remain lower (Fig. 2). At 1737.97 mbs, δ13C values for all three start to shift toward
than 1 for the overlying 2.5 m before increasing gradually (Fig. 5). Above more negative values. This trend continues in the following 5 sampled
1733.97 mbs, OEPs return to near pre-excursion values. The Carbon sediments from the marine mudstone interval, culminating in the
Preference Index (CPI; Bray and Evans, 1961), which uses 9 adjacent n- most 13C-depleted values recorded for the core between 1735.04 and
alkanes from the homologous series, can also be used to express the 1734.5 mbs: − 33.4, − 33.6 and − 35.0‰ for C27, C29 and C31,
preference for odd or even n-alkanes: CPI = 2 × (C25 + C27 + C29 + C31) / respectively (Fig. 2). This negative isotope excursion is interpreted
[C24 + 2 × (C26 + C28 + C30) + C32]. As with the OEP, the CPI decreases as the PETM CIE and the magnitude of the excursion is 4.5‰ for all
from 1.5 to values b1 at 1737.97 mbs and remains low over the marine three n-alkanes. At the peak of the CIE, the offset between n-alkane
190 L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200

Fig. 4. Partial m/z 57 mass chromatograms of the saturated hydrocarbon fractions for representative sediments below and within the PETM CIE, highlighting acyclic hydrocarbons.
(A) Sample Kum-4, 1747.1 mbs depth, and (B) sample Kum-13, 1737.0 mbs depth: •, n-alkanes; Pr, pristane; Ph, phytane; IS, internal standard. Numbers indicate the number of
carbon atoms.

records observed in the lower part of the core becomes less interval and CIE, with the exception of a sample at 1733.43 mbs, δD
pronounced. At 1733.43 mbs, above the marine mudstone, there is a values increase gradually upsection from −169‰ to −137‰ (Fig. 6).
sharp positive shift in all δ13C values, with an abrupt return toward
pre-CIE values and the re-establishment of an offset between records.
However, these values are slightly 13C-depleted relative to pre- 4.3.4. Pristane and phytane
excursion values and remain relatively constant throughout the Although the saturated fraction is dominated by n-alkanes, other
remainder of the section (Fig. 2). biomarkers are present. Pristane (Pr) and phytane (Ph), C19 and C20
isoprenoidal hydrocarbons typically inferred to derive from the phytyl
sidechain in chlorophyll a (e.g. Rowland, 1990; Rontani and Volkman,
4.3.3. n-alkane compound-specific hydrogen isotopes 2003), are present in all sediments. Concentrations track the TOC
δD values were also measured for the C27 and C29 n-alkanes; content, but concentrations normalised to TOC vary throughout the core
concentrations of the C31 n-alkane were too low for accurate δD (Fig. 2). In the lower part (1749.98 to 1739.76 mbs), TOC-normalised
determination. Both n-alkanes have similar values and exhibit similar concentrations are highly variable, ranging from 0.82 to 4.1 μg g− 1
TOC.
trends (Fig. 6). In the lower part of the section (1749.98 to 1739.76 mbs), From 1737.97 to 1734.5 mbs, coinciding with the CIE and marine
prior to the PETM, δD values increase from ca. −160‰ to −140‰ before mudstone unit, TOC-normalised concentrations are relatively constant
decreasing once more to −155‰ (Fig. 6). Samples at 1737.97 and at ca. 1.8 μg g− 1
TOC. Above 1734 mbs, they initially increase but are then
1737.02 mbs, the first two geochemical samples within the CIE, have δD consistently low (b1.5 μg g− 1
TOC) above 1722 mbs (Fig. 2).
values of −165‰, a slight negative shift compared to immediate pre- The pristane to phytane (Pr/Ph) ratio has been invoked to reflect
excursion values. The sample at 1735.92 mbs records a positive shift of ca. depositional conditions, such as changes in salinity and oxicity (Didyk
+20‰, which is followed by a decrease throughout the remainder of the et al., 1978), but must be used with caution as many factors, such as the
CIE; δD values reach a PETM minimum at 1734.5 mbs of −166 and organic matter source and thermal maturity, can also affect the ratio
−165‰ for C27 and C29, respectively (Fig. 6). Above the marine mudstone (Burnham et al., 1982; Ten Haven et al., 1987). Pr/Ph ratios have an
L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200 191

Fig. 5. n-alkane distribution ratios for the Kumara-2 core in New Zealand. (A) n-alkane δ13C values (B) HMW n-alkane odd-over-even predominance (OEP). OEP = (Cn − 2 + 6 × Cn + Cn + 2) /
(4× Cn − 1 + 4 × Cn + 1). (C) HMW n-alkane carbon preference index (CPI). CPI = 2 × (C25 + C27 + C29 + C31) / [C24 + 2 × (C26 + C28 + C30) + C32]. (D) n-alkane terrigenous to aquatic ratio
(TAR). TAR = (C27 + C29 + C31)/(C15 + C17 + C19). The shaded area highlights the PETM CIE and the horizontal dashed lines indicate the first occurrence of thermophilic angiosperm pollen
and the most abundant occurrence of the dinoflagellate genus Apectodinium, respectively. The arrows highlight the inferred increase in terrigenous organic matter input during the CIE,
within the marine interval.

average value of 5.5 for most of the Kumara-2 core (Fig. 7). A marked quantification. Oleanane concentration varies greatly and parallels
decrease occurs at the onset of the CIE, with values ≤2.5 occurring in fluctuations in TOC content (Fig. 2). While TOC-normalised oleanane
the four lowermost samples within the CIE. Values then increase in concentrations are very low in the majority of samples (Fig. 9), there is a
the upper part of the recovered CIE and reach a pre-CIE ratio of 6.0 ca. 3 m interval associated with the CIE and marine mudstone unit
at 1733.4 mbs (Fig. 7). Ratios remain greater than 4.5 for the where concentrations are notably higher. An initial increase at
remainder of the core except for the topmost sediment at 1710 mbs 1737.97 mbs is followed by a maximum at ca. 1735.92 mbs, which
(Pr/Ph = 2.8). marks an overall increase of nearly 2 orders of magnitude compared to
background values (Fig. 9). After this spike, concentrations decrease
4.3.5. Oleananes sharply, returning at 1734.5 mbs to values close to those found in the
Various saturated triterpanes are present in the Kumara sediments, lower part of the core.
with oleanane, a diagenetic product of various triterpenoids found in
angiosperm species (Rullkötter et al., 1994), being the predominant 4.3.6. Hopanes and related indices
non-hopanoid. It is present in the majority of sediments in both its Hopanes, diagenetic products of bacterial bacteriohopanepolyols
18α(H),22β(H) and 18β(H),22α(H) isomeric configurations (Figs. 8 (Ourisson et al., 1979; Rohmer et al., 1984), occur in all Kumara
and 9). However, due to low concentrations, the isomer peaks were not sediments. As with other biomarkers, only the fully saturated defunc-
fully resolved in all chromatograms and the two were summed for tionalised hopanes were observed. All samples contain a suite of C29–C35
17α(H),21β(H)-hopanes (Fig. 8), and C31 to C35 homohopanes are
present in both their 22R and 22S isomeric configurations. The C27
hopanes, trisnorneohopane and trisnorhopane also occur throughout.
Moretanes, hopanes with the 17β(H),21α(H) isomeric configuration,
are also present, albeit in low concentrations, and hopanes with the
biologically produced ββ configuration were not detected.
The homohopane isomerisation ratio (22R to 22S) for C31
homohopane has an average value of 0.5 with a standard deviation
of 0.02 for the entire study section. Similarly, the ratio of βα to αβ for
the C30 hopane varies between 0.4 and 0.6 with an average value of 0.5
and a standard deviation of 0.07. Such ratios suggest an equivalent
vitrinite reflectance (R0) no greater than 0.5 (Seifert and Moldowan,
1980; Waples and Machihara, 1991; Peters et al., 2005), and the
persistence of an OEP in the n-alkane distribution suggests that the
section has only just entered the oil window. The relatively uniform
values for both thermal maturity indicators show that the degree of
maturity remains unchanged throughout the interval examined in the
Kumara-2 core.
The homohopane index is defined as the ratio of C35 αβ homohopane
Fig. 6. δD values of terrestrial higher plant derived n-alkanes for the Kumara-2 core in to the total C31–35 homohopane concentration (Peters and Moldowan,
New Zealand. The error bars represent one standard deviation based on duplicate
analyses. The shaded area highlights the PETM CIE and the horizontal dashed lines
1991). Throughout most of the core, the homohopane index is b0.04
indicate the first occurrence of thermophilic angiosperm pollen and the most abundant (such that the C35 component represents less than 4% of the total
occurrence of the dinoflagellate genus Apectodinium, respectively. homohopanes), with typical values of 0.02 (2%; Fig. 7). Between 1737.97
192 L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200

Fig. 7. Indicators of depositional redox conditions for the Kumara-2 core in New Zealand. (A) n-alkane δ13C excursions. (B) Homohopane index, the ratio of C35 αβ-homohopane to
total C31–35 homohopane concentration. (C) Ratio of trisnorneohopane (Ts) to total C27 hopane (Ts + Tm). (D) Pr/Ph ratio. The shaded area highlights the PETM CIE and the
horizontal dashed lines indicate the first occurrence of thermophilic angiosperm pollen and the most abundant occurrence of the dinoflagellate genus Apectodinium, respectively.

and 1734.5 mbs, however, homohopane indices increase markedly and occurring changes in organic matter source, the CIE was evaluated using
abruptly, with a maximum of 28% at ca. 1735.92 mbs before decreasing the C27, C29 and C31 n-alkanes, typically derived from terrestrial higher
upcore to b4% at depths above 1734.5 mbs (Figs. 7 and 8). plant leaf waxes (Eglinton et al., 1962). The pronounced CIE of 4.5‰,
17α-22,29,30-trisnorhopane (Tm) and 18α-22,29,30-trisnorneo- which is identified in sediment samples taken from 1737.97 to
hopane (Ts) are both C27 hopanes that differ with respect to the 1734.5 mbs, combined with the plant microfossil zonation and
position of a D-ring methyl group; Tm is substituted at C-18 and Ts at presence of abundant Apectodinium dinocysts, confirm that the
C-17. The ratio of Ts to total C27 hopane, (Ts + Tm), varies between section spans the Paleocene/Eocene boundary and the PETM. Based
0.86 and 0.96 throughout most of the section. While Ts is the on lithological, geochemical and palynological observation, the
predominant isomer (Fig. 7), at 1737.97 mbs, the ratio decreases recovered CIE spans 4.85 m (Fig. 2). The CIE is terminated by a very
abruptly to 0.64 and then returns to near pre-excursion values above sharp return to near pre-excursion values at an unconformity
1734.5 mbs (Fig. 7). recognised at 1734.25 mbs (Fig. 2). This suggests the PETM is
truncated and the latter part of the interval is missing from the
5. Discussion sedimentary record, as many existing PETM records reveal a more
gradual return (e.g. Kennett and Stott, 1991; Zachos et al., 2003;
5.1. Stratigraphy of the Kumara-2 core Zachos et al., 2005; Sluijs et al., 2006).
The gradual decrease in n-alkane δ13C values at the onset of the
5.1.1. Maturity and implications CIE, which is not often observed in PETM records, is interpreted as a
As indicated by biomarker distributions, organic matter (OM) in gradual input of 13C-depleted carbon into the earth–atmosphere
the Kumara-2 core has just entered the early stages of the oil window. system, suggesting that the early stages of the event were recovered
Hydrogen isotopic enrichments have been shown to occur in at this site. It is possible that this gradual decrease was caused by the
n-alkanes in thermally mature sedimentary OM and oils (Li et al., reworking of pre-CIE n-alkanes, particularly as the transition from a
2001; Schimmelmann et al., 2004; Dawson et al., 2005; Radke et al., terrestrial to marine depositional environment could have encour-
2005; Pedentchouk et al., 2006; Dawson et al., 2007). However, aged the resuspension, mixing and redeposition of pre-CIE material,
diagenetic n-alkane hydrogen exchange may only occur significantly with the reworked component gradually decreasing into the early
in early to post mature sediments (R0 values of ca. 0.7 to 1.6; stages of the CIE. However, the onset interval spans more than 2 m of
Pedentchouk et al., 2006; Dawson et al., 2007). The effects of core and this would therefore require the resuspension of older
thermal maturation on the carbon isotopic composition of individual material for a significant period of time. Also, there is a significant
n-alkanes are not believed to be as significant as for hydrogen, negative shift in n-alkane δD values at the onset of the CIE. If pre-CIE
occurring most significantly in the oil generation window, during and n-alkanes were being extensively reworked then this negative
following petroleum formation (e.g. Clayton, 1991; Clayton and excursion must have been even larger. This is followed by a rapid
Bjorøy, 1994; Rooney et al., 1998; Harris et al., 2003). With an positive shift in δD values within the onset interval, which does not
inferred R0 value of ca. 0.5, the Kumara-2 sediments are only in the seem compatible with the idea of a gradual decrease in pre-CIE n-
early stages of thermal maturity and, therefore, the n-alkane carbon alkane input.
and hydrogen isotope compositions are assumed to record original The unconformity at 1734.25 mbs, between the marine mudstone
values. Furthermore, maturity remains constant throughout the core unit and the overlying sandstone-dominated lithology, precludes the
and thus trends may be preserved even if absolute values are not. determination of accurate sedimentation rates. However, the most
recent age model established by Röhl et al. (2007), indicates that peak
5.1.2. Bio- and chemostratigraphic recognition of the PETM negative values are reached ca. 60 kyr (3 precession cycles) after the
One of the defining characteristics of hyperthermals, such as the onset of the CIE. Thus, the recovered portion of the PETM could span
PETM, is a pronounced negative CIE. Due to the complex lithological up to ca. 60 kyr. If this is the case, then the sedimentation rate during
changes that occur in the Kumara-2 core, along with the inferred co- the recovered CIE can be estimated at ~ 4.85/60 = ~ 8.08 × 10− 2 m/kyr.
L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200 193

Fig. 8. Partial m/z 191 mass chromatograms of the saturated hydrocarbon fractions, highlighting hopanes and oleanane, for representative sediments from below and within the
Kumara PETM CIE. (A) sample Kum-4, 1747.1 mbs depth, and (B) sample Kum-13, 1737.0 mbs depth: Ts, trinorneohopane; Tm, trisnorhopane. For the αβ configurations, when two
isomers are shown in the chromatogram the first corresponds to the 22R(H) isomer and the second to the 22S(H) isomer. Numbers indicate the number of carbon atoms.

5.2. Depositional environment at Kumara-2 core

5.2.1. Changes in depositional environment and sea level rise


Lithological, palynological and organic biomarker analyses all
indicate marine depositional conditions between 1739.1 and
1734.25 mbs. Dinocysts are recorded in all samples from this interval,
but are absent above and below it; the one exception occurs at
1733.4 mbs, in the lower part of the sandstone-dominated lithology,
where dinocysts comprise 2% of total palynomorphs, reflecting either a
minor marine influence in the base of the sandstone unit or reworking of
older CIE material. The first occurrence of dinocysts coincides with the
presence of a dark non-calcareous mudstone (Fig. 2). Moreover, coal
layers and lenses indicative of a terrestrial depositional environment are
interspersed throughout the rest of the core, but are absent from the
mudstone unit.
Many of the aforementioned biomarker proxies can also reflect the
change in depositional environment. In particular, the shift in TAR
(Fig. 5), marking an increase in the relative concentration of low
molecular weight n-alkanes, possibly derived from algal precursors (e.g.
Gelpi et al., 1970; Naraoka and Ishiwatari, 1999; Duan and Wang, 2002), Fig. 9. Timing of higher plant community change recorded at the Kumara-2 core in New
Zealand. Oleanane concentration is expressed in μg per g of total organic carbon (TOC).
likely reflects the shift to marine conditions. Similarly, the shift to lower The shaded area highlights the PETM CIE and the horizontal dashed lines indicate the
CPI and OEP values (Fig. 5) is probably a consequence of the change in first occurrences of thermophilic angiosperm pollen and the most abundant occurrence
depositional environment, reflecting a decrease in the sedimentary of the dinoflagellate genus Apectodinium, respectively.
194 L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200

concentration of terrestrial leaf waxes and a possible dilution by other n- that are at least partially dependant on the organic matter source and
alkane sources (Rieley et al., 1991; Collister et al., 1994a; Collister et al., host rock lithology (e.g. Rullkötter et al., 1984; Sinninghe Damsté et al.,
1994b). We note that the higher molecular weight n-alkanes likely still 1988; Sinninghe Damsté and de Leeuw, 1990). Thus, the trends also
have a higher plant origin in the marine interval, but that the leaf waxes reflect the transition from a fluvial to marine depositional environment.
could have been more degraded during transport and diagenetic While we cannot interpret trends in biomarker ratios, we can
reactions could introduce n-alkanes via the degradation of n-alkyl consider the absolute values from the marine mudstone interval,
moieties derived from a range of biological precursors, resulting in lower between 1739.1 and 1734.25 mbs. The very high homohopane indices
CPI/OEP values. and low Pr/Ph ratios suggest that the marine depositional environment
The above results indicate a transient rise in sea level during the was suboxic to anoxic. For example, Peters et al. (1995) interpreted a
PETM. The Late Paleocene is generally believed to have been an ice- homohopane index of 12% as representing anoxic source-rock deposi-
free Earth, although recent studies support the presence of transient tional conditions, whereas in the Kumara-2 marine interval C35
ice sheets in the Early Paleogene (DeConto and Pollard, 2003; Miller et homohopanes comprise up to 28% of total homohopanes. The high
al., 2005). However, in the absence of evidence for extensive ice sheets abundance of Apectodinium dinocysts, a presumed heterotrophic taxon,
directly preceding the PETM, the most probable explanation for this suggests nutrient-rich surface waters during the CIE could have been
change is a sea level rise caused by the thermal expansion of sea water associated with eutrophication, consistent with other studies in New
(see Sluijs et al., 2008a for a review). Warming of ocean waters during Zealand and elsewhere (Crouch et al., 2003; Sluijs et al., 2005; Sluijs et al.,
the PETM was remarkably uniform for both surface and deep waters, 2008a). Eutrophic conditions and a more anoxic water column have been
with an average temperature increase of ca. 5 °C (Kennett and Stott, reported for the PETM from the New Jersey Shelf, where it is thought to
1991; Bralower et al., 1995; Thomas et al., 1999; Zachos et al., 2003; be a result of enhanced terrestrial runoff and nutrient delivery (Lippert
Tripati and Elderfield, 2005), which, given the temperature–density and Zachos, 2007). These conditions could promote the lowering of the
relationship of seawater, should cause a 3–5 m rise in global sea level. redox potential of sedimentary pore water (Lippert and Zachos, 2007)
The results from Kumara-2 agree with other evidence for sea level rise and are consistent with what is observed at Kumara-2.
in PETM sections from other geographical locations, including the
North Sea (e.g. Bujak and Brinkhuis, 1998), Arctic Ocean (Sluijs et al., 5.3. Changes in terrestrial plant community
2006), Egyptian Tethys (Speijer and Morsi, 2002), New Jersey margin
(Sluijs et al., 2008a) and the Tawanui section in the North Island of Palynomorph assemblages show that the proportions of fern spores,
New Zealand (Crouch and Brinkhuis, 2005). gymnosperm and angiosperm taxa change at the onset of the PETM.
Throughout the CIE at Kumara-2, despite the marine depositional Through the lower part of the observed CIE, the gymnosperm relative
environment, there is a significant terrestrial contribution to buried abundance declines, fern spore abundance increases and the percentage
organic matter and palynomorph assemblages: the n-alkane TAR of angiosperm taxa drops slightly (Fig. 3). Within the angiosperm
ranges from 1.3 to 4.2 and terrestrial palynomorphs comprise ≥69% of assemblage, there is a noticeable change in the composition with
the total palynomorph assemblage (Figs. 2 and 5). The upper part of increases in several taxa, including Spinizonocolpites prominatus (Nypa
the recovered CIE is also characterised by an increase in the terrestrial mangrove) and Malvacipollis. Both angiosperm and gymnosperm relative
component of sedimentary organic matter relative to marine OM abundance increases in the upper part of the CIE interval. Above the
(increase in the n-alkane TAR from 1.3 to 4.2) and increased terrestrial unconformity at 1734.25 mbs, the Early Eocene sequence is character-
palynomorphs relative to marine palynomorphs compared to the ised by a higher relative abundance of angiosperm palynomorphs than in
onset of the CIE. This could reflect marine regression, but n-alkane the Late Paleocene, with a similar component of gymnosperms and fewer
δ13C values suggest the CIE is truncated and this interval appears to fern spores. Notably, the most significant change in the plant microfossil
precede the recovery stage of the isotope excursion, thus preceding record, an increase in Casuarina (Myricipites harrisii) pollen, occurs in the
the termination of the warming period and associated drop in sea Early Eocene sequence.
level. We speculate that a possible explanation is an increase in the The dramatic increase in oleanane concentration at the onset of the
amount of terrestrial organic matter delivered to the sediment. CIE provides further evidence for changes in the higher plant community
Indeed, previous studies show evidence for increased sedimentation structure (Fig. 9). β-amyrin, taraxerol, germanicol and oleanoic acid all
rates at other New Zealand sites during the PETM (Crouch et al., 2003; have an oleanane-type hydrocarbon skeleton, making this one of the
Hollis et al., 2005; Nicolo et al., 2007). Moreover, there is evidence most widespread fossil triterpenoids, a group of compounds largely
from many other geographical locations for an increase in weathering restricted to angiosperm plants (Ten Haven and Rullkötter, 1988;
and terrestrial runoff during the PETM, in the form of elevated Rullkötter et al., 1994). Consequently, oleananes have long been used as
sedimentation rates or increased kaolinite and terrestrial organic angiosperm biomarkers in ancient sediment and petroleum studies (e.g.
matter input (e.g. Schmitz and Pujalte, 2003; Kelly et al., 2005; Killops and Frewin, 1994; Moldowan et al., 1994). Oleanane concentra-
Schmitz and Pujalte, 2007). tion increases rapidly in the lower part of the CIE but then decreases
equally abruptly, before the termination of the CIE, suggesting a
5.2.2. Further biomarker insight into the depositional environment transient shift in the plant community, perhaps due to transitional
The distribution of hydrocarbon biomarkers can reflect thermal climatic conditions. An expansion of angiosperms, in particular the
maturity, depositional environment and lithology (Peters et al., 2005). transient migration of tropical and sub-tropical species to higher
In the Kumara-2 core, thermal maturity does govern the absolute values latitudes, has been recognised from locations such as the Bighorn
of the hopane ratios, but, because thermal maturity is constant Basin (Wing et al., 2005) and Arctic (Sluijs et al., 2006), during the PETM.
throughout the studied interval, it cannot explain the observed At Kumara, however, there is no palynological evidence for an overall
variations. Instead, these variations must reflect changes in the angiosperm expansion. Indeed, angiosperm pollen relative abundance
depositional environment. Homohopane indices increase markedly slightly decreases in the lower part of the CIE (Fig. 3). The increased
during the CIE, whereas Pr/Ph ratios and Tm/(Tm + Ts) ratios both oleanane concentrations could therefore arise from a reorganisation
decrease (Fig. 7). All of these trends appear to be consistent with a shift rather than an expansion of the angiosperm assemblage: oleanoid-type
from oxidising to relatively anoxic depositional conditions (Powell and compounds are produced in high concentrations by mangrove plants
McKirdy, 1973; Didyk et al., 1978; Moldowan et al., 1986; Peters and (e.g. Ghosh et al., 1985; Koch et al., 2003a), and the beginning of the
Moldowan, 1991). However, they cannot be interpreted so simplistically increase in oleanane concentration coincides with the appearance of
in this setting. Each of the aforementioned ratios reflects a combination Spinizonocolpites prominatus pollen (Fig. 3), which is considered to have
of source inputs and subsequent diagenetic and catagenetic reactions been produced by Nypa mangrove plants (e.g. Collinson, 1993). This
L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200 195

increase in concentration could also be due to enhanced preservation 2006a, 2006b); or 2) a lower isotope effect during evaporation and
resulting from the transition from an oxic terrestrial to marine source water vapour formation, arising from much warmer ocean
depositional environment. However, the increase in oleanane concen- surface temperatures around New Zealand. Given that there is little
tration is not synchronous with this transition and in the latter part of the precipitation induced fractionation of New Zealand meteoric water in
recovered CIE concentrations decrease whilst the depositional environ- the present day, it seems unlikely that this value was even lower during
ment remains unchanged. Whether the dramatic change in oleanane the Paleogene. Instead, we interpret the higher values as a consequence
concentrations is strictly a consequence of the rise in sea level and of warmer oceans, a hypothesis which should be tested by adding
change in depositional environment, or instead reflects more wide- hydrogen isotopic components to general circulation models.
spread ecosystem changes in New Zealand remains to be seen.
Nonetheless, there is evidence for the development of a transitional 5.4.2. Stratigraphic changes in n-alkane δD values
mangrove swamp at Kumara during the PETM. There are several variations in the n-alkane δD record, with shifts
in the upper and lower parts of the core being comparable to those
5.4. Changes in the hydrological cycle occurring at the PETM (Fig. 6). We consider each of these discreetly
and for each discuss the potential influence of vegetation change or a
5.4.1. n-alkane δD values marine contribution to the HMW n-alkane pool.
Hydrogen isotope (δD) values of terrestrial higher plant leaf waxes
record: 1) the δD values of meteoric water (Sauer et al., 2001; 5.4.2.1. Prior to the PETM. From 1744.42 to 1737.02 mbs, preceding and
Chikaraishi and Naraoka, 2003; Sachse et al., 2004; Rao et al., 2009); spanning the onset of the CIE, n-alkane δD values decrease from −145 to
2) the extent of D-enrichment during soil evaporation and leaf −165‰ (Fig. 6). This decrease in δD values starts prior to the inferred
evapotranspiration, which is dependent on humidity (Smith and changes in plant community structure and the change in depositional
Freeman, 2006); and 3) the biological fractionation that occurs during environment at 1739.1 mbs. Crucially, either an increased angiosperm
leaf wax synthesis. Thus, they can be powerful tools in the reconstruc- (Pedentchouk et al., 2007, 2008) or marine contribution would be
tion of past environmental conditions, although an important caveat is expected to result in more D-enriched rather than depleted values; this
the effect of vegetation change as both evapotranspiration and is not observed, suggesting that organic matter source changes are not
biosynthetic effects can vary amongst different taxa. In the Kumara-2 the primary control on changes in n-alkane δD values. As such, our data
core, δD values of the C27 and C29 n-alkanes range from −137 to −168‰ suggest that prior to and during the onset of the PETM meteoric water
and values for both are very similar throughout the study interval was becoming D-depleted due to more precipitation and/or that
(Fig. 6). These values are comparable in magnitude to those observed in evaporative enrichment was decreasing perhaps due to a more humid
Tanzania during the early stages of the CIE (Handley et al., 2008) and are climate. The former could arise not only from increased precipitation in
more D-enriched than those reported from northern mid latitudes in the the New Zealand region but also a shift to a more distant source, such
Bighorn Basin (−195 to −185‰; Smith et al., 2007) and the Arctic that water vapour has been fractionated to a greater degree during
(−230 to −260‰; Pagani et al., 2006a, 2006b) across the Paleocene/ transport. Distinguishing local and regional changes will require greater
Eocene boundary. geographical coverage; nonetheless, these data suggest changes in the
Changes in higher plant community can influence the isotopic hydrological cycle prior to the PETM.
composition of leaf waxes (e.g. Chikaraishi and Naraoka, 2003; Liu and
Huang, 2008; Pedentchouk et al., 2008). However, previous studies have 5.4.2.2. During the PETM. As discussed above, n-alkanes in the two
estimated average hydrogen isotopic enrichment factors between C29 n- lowermost samples (1737.97 and 1737.02 mbs) within the CIE have
alkane and meteoric source water (εC29-GW) during leaf wax lipid relatively low δD values. Above this, the n-alkane δD record becomes
biosynthesis. εC29-GW was determined to be ca. −130‰ to −120‰ for more complex, with large positive and negative shifts (Fig. 6). These
surface sediment n-alkanes across a transect of 13 European lakes, with shifts occur after the beginning of the marine mudstone unit and
lower values at higher latitudes (Sachse et al., 2004), and −140‰ to change in plant microfossil composition, suggesting that changes in n-
−130‰ for surface soil n-alkanes across 28 sampling sites in Eastern alkane source are not the sole controlling factors. Nonetheless, δD
China, again with lower values at higher latitudes (Rao et al., 2009). values in the PETM interval should be interpreted cautiously. While
These values are consistent with other reported hydrogen isotopic detailed age models are lacking from all four sites, the Kumara δD record
composition relationships between surface soil n-alkanes and meteoric is broadly similar to those from Tanzania (Handley et al., 2008), the
water from a variety of different geographical locations with varying Arctic (Pagani et al., 2006b) and the Bighorn Basin (44°N) in North
vegetation types (e.g. Chikaraishi and Naraoka, 2003; Bi et al., 2005; America (Smith et al., 2007): all of these sites exhibit a dramatic positive
Sachse et al., 2006; Smith and Freeman, 2006). Assuming an average excursion during the early part of the CIE, and at the Arctic, Bighorn
εC29-GW of ca. −130‰, Early Paleogene precipitation at Kumara Basin and Kumara sites the positive shift is shortly followed by a return
(paleolatitude 55°S; paleolongitude 170°W) would have had a mean to lower δD values; unfortunately, the Tanzanian record is truncated and
δD value of ca. −40 to −10‰. Currently, average annual precipitation it is not possible to comment on whether a similar return may also occur
δD values for New Zealand range from ca. −55‰ to −27‰, with values there. Interestingly, in both the Kumara and Tanzanian records, the
of ca. −40‰ for the Kumara region (Bowen and Revenaugh, 2003; positive δD excursion does not occur at the base of the CIE and instead
Bowen, 2009). However, global δDsw values (hydrogen isotopic lags it, in apparent contrast to the Arctic and Bighorn Basin records.
composition of seawater) were likely ca. 12‰ lower during the From the Arctic record, Pagani et al. (2006b) interpreted the positive
Paleogene, ca. 8× larger than the putative δ18Osw difference for an ice- shift in C27 and C29 n-alkane δD values during the early part of the CIE as
free world (1.5‰; e.g. Shackleton and Kennett, 1975; Shackleton, 1977; an increase in the δD of meteoric water, caused by decreased
Lear et al., 2000; Schrag et al., 2002; Zachos et al., 2006). If this is the case, precipitation, and isotopic distillation, at lower latitudes, an interpretation
precipitation δD values should be more negative than present day consistent with an observed decrease in Arctic Ocean salinity (Sluijs et al.,
estimates but the opposite is observed in the Kumara-2 core. 2006; Sluijs et al., 2008b). Arctic air masses are particularly susceptible to
There are several reasons for this potential discrepancy. First, the distillation processes due to the long pathway of vapour transport and
εC29-GW could have been lower than −130‰; that could arise from extensive precipitation that occurs along it; however, isotopic distillation
either a more arid climate state or from the specific vegetation regime. may also have affected the Kumara site. If so, the positive excursion
Alternatively, meteoric water could have been more D-enriched during reflects either a shift to more local, as opposed to tropical, water sources
the Cenozoic, which could be indicative of: 1) less precipitation locally or less precipitation during transport (Pagani et al., 2006b). Alternatively,
or during water vapour transport from the source area (e.g. Pagani et al., the positive shift could record an increase in the temperature of the
196 L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200

source waters (e.g. Gat, 1996) and less fractionation during water vapour 1991; Thomas et al., 2002). Few leaf wax lipid compound-specific δ13C
formation. A further explanation is that the shift records more arid records exist for the PETM (Kaiho et al., 1996; Hasegawa et al., 2006;
conditions, bringing about enrichment of soil or leaf water via increased Pagani et al., 2006b; Smith et al., 2007; Handley et al., 2008), yet these
evaporation (Sachse et al., 2004); indeed, this explanation was invoked to are crucial for the interpretation of the magnitude of the CIE as recorded
explain the positive excursion observed in the Bighorn Basin (Smith et al., in the terrestrial carbon pool. Various attempts have been made to
2007) and is consistent with other climate indicators from that region reconcile these records or argue which more closely represents the
(Wing et al., 2005; Smith and Freeman, 2006; Kraus and Riggins, 2007). atmospheric CIE (Bowen et al., 2004; Smith et al., 2007).
Globally and locally, the PETM is characterised by evidence for a In addition to the isotopic composition of atmospheric CO2, factors
more active hydrological cycle, including increased sedimentary which affect carbon isotopic fractionation during photosynthesis will
kaolinite contents (e.g. Robert and Chamley, 1991; Gibson et al., 1993; also influence the δ13C values of terrestrial higher plant n-alkanes (e.g.
Kaiho et al., 1996; Schmitz and Pujalte, 2003; Zachos et al., 2006), Diefendorf et al., 2010). These include changes in humidity (e.g.
suggesting either more humid conditions (Robert and Kennett, 1994; Farquhar et al., 1989) and shifts in the composition of the higher plant
Gibson et al., 2000; Bolle and Adatte, 2001) or increased soil erosion community, especially changes in the proportion of gymnosperms to
(Thiry and Dupuis, 2000; Schmitz and Pujalte, 2003), and increased angiosperms (Chikaraishi and Naraoka, 2003; Smith et al., 2007). A
sedimentation rates (e.g. Crouch et al., 2003; Lippert and Zachos, 2007; further concern in the Kumara-2 core is the potential change in the
Sluijs et al., 2008b). Notably, studies have shown that sedimentation rates source of n-alkanes, as marine organic matter can be 13C-enriched
increased at several New Zealand marine PETM sections (Hollis et al., relative to terrestrial organic matter (Popp et al., 1989). However, the
2005; Nicolo et al., 2007), while in the Kumara sequence there is an difference between the δ13C values of terrestrial and aquatic organic
increase in the proportion of terrigenous biomarkers within the marine matter is generally smaller in a high pCO2 greenhouse world such as the
mudstone unit. Thus, the simplest explanation for our δD record is Paleogene. Also, as discussed above, n-alkane hydrogen isotopes show
increased export of D-enriched water to New Zealand, analogous to that no consistent D-enrichment through the marine interval as would be
inferred by Pagani et al. (2006b) for the Arctic. However, the Tanzanian expected if they derived from marine organic matter (Fig. 6).
PETM section is also characterised by a positive δD excursion and Increases in humidity result in increased carbon isotopic fractiona-
increased sedimentation rate. It is difficult to envisage how water vapour tion during photosynthesis (e.g. Farquhar et al., 1989; Diefendorf et al.,
isotopic distillation, which is already minimal, could have decreased at 2010). Thus, a potential increase in humidity during the PETM could
this tropical site. Instead, it is argued that Tanzania experienced a more result in the amplification of a negative CIE. While it is probable that
arid (higher δD values) but more seasonal climate (increased sedimen- several factors influence n-alkane δD values in the Kumara record, the
tation rate); such a model could also apply to New Zealand and should be most negative n-alkane δ13C values coincide with some of the most
explicitly tested in future climate simulations. positive δD values (Fig. 6), which is inconsistent with elevated humidity
Regardless of mechanism, the Kumara climatic perturbation reflected enhancing carbon isotope fractionation by plants and amplifying the
by the positive δD excursion was transient and the associated CIE. These relationships are similar to observations in Tanzania
hydrological regime quickly returned to conditions similar to those (Handley et al., 2008) and the Bighorn Basin (Smith et al., 2007).
directly preceding the onset of the PETM within the CIE. This appears to be Several studies have shown that n-alkanes derived from gymnos-
a global characteristic of PETM δD records, with only the truncated record perms are typically more 13C-enriched compared to those found in
from Tanzania being an exception (Pagani et al., 2006b; Smith et al., 2007; angiosperms grown under the same climatic conditions (Leavitt and
Handley et al., 2008). Thus, it appears that any perturbation in the Newberry, 1992; Flanagan et al., 1997; Chikaraishi and Naraoka, 2003;
hydrological regime, at least with respect to isotopic partitioning, was a Diefendorf et al., 2010). Results from the Arctic also show that during the
transient response to PETM warming. PETM the magnitude of the negative CIE recorded in angiosperm-
specific biomarkers is greater than that in gymnosperm-specific
5.4.2.3. Post PETM. In the upper part of the core (1721.54 to 1710 mbs), compounds from the same sediments (Schouten et al., 2007), likely
n-alkane δD values increase from ca. −160 to −135‰ (Fig. 6) and are due to changes in water use efficiency. Therefore, it is important to be
also apparently unrelated to changes in the depositional environment or able to assess potential changes in the plant community when
vegetation groups. This increase, particularly in the upper 10 m of the interpreting terrestrial carbon isotope excursions. As described in
core, could record either increasing aridity or decreasing water vapour Section 5.3, there is evidence at Kumara-2 for an increase in the relative
isotopic distillation as discussed above. Intriguingly, it is at the top of this abundance of angiosperms compared to gymnosperms at the onset of
interval that the most dramatic vegetation change in the Kumara-2 core the CIE and an increase in fern spores in the lower part of the CIE, and
is recorded: a shift to abundant Casuarina (Myricipites harrisii) pollen, these could have affected the magnitude of the CIE. However, terrestrial
which is recorded in sequences throughout New Zealand. In modern higher plant community structure, as recorded by palynomorph relative
settings, Casuarina species are well adapted to dry habitats or prolonged abundances, remains relatively unchanged during the CIE even across
dry seasons (Heywood, 1978; Osundina, 1997), and locally drier the interval where δ13C values decrease to a minimum at 1735.92 mbs.
conditions during the Early Eocene may therefore have played a role Also, the increase in oleanane concentration occurs in the early stages of
in this notable vegetation change. the PETM (Fig. 9). At 1734.5 mbs, δ13C values remain depleted by ca.
4.5‰ relative to pre-excursion values, whilst oleanane concentrations
5.5. The carbon isotope excursion have decreased to values found in the lower part of the core. Therefore,
while the initial change in terrestrial plant composition may influence
5.5.1. Magnitude of the CIE the initial CIE values, it seems that changes in plant community do not
Assuming peak negative carbon isotope values are recovered, the significantly affect the most negative n-alkane δ13C values, which
magnitude of the CIE in the Kumara-2 core, as recorded by the δ13C suggests that the −4.5‰ shift does reflect primarily a change in the
values of higher plant leaf wax n-alkanes, is ca. 4.5‰. There is much carbon isotopic composition of atmospheric CO2. This value is very
debate concerning the fidelity with which different sedimentary similar to that of −4.6‰, obtained from plant leaf wax n-alkanes from
archives record the actual shift in atmospheric CO2 δ13C values during the Bighorn Basin (Smith et al., 2007) when attempting to account for
the PETM. The magnitude of the isotope excursion in terrestrial records changes in mean annual precipitation and changes in plant community
is often larger than that found in marine settings: terrestrial paleosols structure (Diefendorf et al., 2010).
typically record a CIE of −4 to −5‰ (e.g. Koch et al., 1992; Koch et al., A CIE of −4.5‰ has important implications for estimating the quantity
2003b; Schmitz and Pujalte, 2003), compared to a −2.5 to −3‰ of carbon added to the ocean–atmosphere system and its isotopic
excursion in deep sea foraminifera δ13C values (e.g. Kennett and Stott, composition. Such an isotopic shift would require the addition of 3600 PgC
L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200 197

derived from methane hydrates (assuming a δ13C value of −60‰), CIE. The record from Kumara-2 provides further evidence for sea level
which is inconsistent with estimates of the Paleocene hydrate reservoir rise, and possibly warming, prior to the peak of the CIE during the PETM.
(Buffett and Archer, 2004). Model simulations by Zeebe et al. (2009),
based on the magnitude of marine CaCO3 dissolution associated with the 6. Conclusions
PETM carbon pulse, estimate an input of ca. 3000 PgC with a carbon
isotopic composition lighter than −50‰, assuming a CIE of −3‰. In We have identified a new PETM section in New Zealand that occurs
contrast, simulations by Panchuk et al. (2008) require a carbon pulse of over a 4.85 m interval and is characterised by a negative CIE of 4.5‰,
at least 6800 PgC with an average δ13C value of −22‰, based on a −4‰ recorded in higher plant leaf wax n-alkanes. In the Kumara-2 core, South
excursion, a scenario which is more compatible with a CIE of −4.5‰. Island, dramatic changes in depositional environment during the PETM
Furthermore, Zeebe et al. (2009) argued that the addition of 3000 PgC to are interpreted as evidence for a sea level rise, with a transition from a
the ocean–atmosphere system was not sufficient to drive reported terrestrial to a near-shore marine setting observed in lithological,
global sea surface temperature increases of 5–8 °C (e.g. Kennett and palynological and organic biomarker records. Evidence for sea level rise
Stott, 1991; Sluijs et al., 2006; Zachos et al., 2006) and that other coincides with the initiation of the negative CIE but predates the most
unrecognised forcing mechanisms must play a significant role. In 13
C-depleted values in the terrestrial higher plant n-alkane record. Other
contrast, a larger carbon input similar in size to that invoked by Panchuk proxies suggest that this marine depositional environment was anoxic
et al. (2008), can yield an increase in pCO2 consistent with PETM or suboxic, possibly with nutrient-rich surface waters during part of the
warming and with the observed − 4.5‰ CIE. This supports the PETM. Higher plant δD values are harder to interpret due to complex
hypothesis that carbon sources that were less 13C-depleted than source changes but suggest that the PETM was characterised by notable
methane hydrates also contributed to the PETM CIE (e.g. Kurtz et al., hydrological variations that are similar to those observed in North
2003; Svensen et al., 2004; Storey et al., 2007). America, the Arctic and Eastern Africa. A negative shift in δD values
across the transition into the PETM apparently reflects an initial
decrease in precipitation-induced isotopic distillation during water
5.5.2. Timing of the CIE and sea level rise vapour transport to New Zealand. Intriguingly, dramatic changes in
The onset of the CIE is gradual in all 3 n-alkane records, with 3 higher plant δD values occur in horizons preceding and following the
intermediate δ13C values spanning over 3 m of sediment before maximum PETM. Climatic and/or depositional environment changes caused
negative values are attained (Fig. 2). This gradual onset is not a common variations in terrestrial plant assemblages, with the appearance of
feature of PETM records and could provide direct evidence that the release pollen associated with thermophilic conditions and the development of
of carbon into the atmosphere was not ‘instantaneous’. This interpretation Nypa mangrove swamps. The onset of the CIE also saw an increase in
does require caution, however, as an alternative mechanism could be fern spores and an initial decrease in gymnosperms. The above
mixing of pre-excursion n-alkanes into the excursion interval. The observations, combined with the gradual onset of the CIE, suggest that
transport of n-alkanes into the sedimentary record, particularly into the injection of 13C-depleted carbon into the ocean–atmosphere system
marine sediments, can occur with significant lags (N1000 yr) as shown by was not geologically instantaneous and that carbon continued to be
radiocarbon analyses of recent sediments (Eglinton et al., 1997; released after an initial sea level rise (and inferred temperature
Smittenberg et al., 2006). increase), inconsistent with the hypothesis of a single catastrophic
The timing of the CIE relative to the PETM temperature increase is release of carbon as the sole driver of warming during the PETM.
currently a topic of much interest. In northeastern North America,
increases in sea surface temperature (SST) may have occurred prior to Acknowledgements
the onset of the CIE and maximum warming predates the most 13C-
depleted values in the marine realm (Sluijs et al., 2007). Likewise, sea Palynological processing was completed by Roger Tremain, GNS
level rise at these sites, believed to be caused at least partially by the Science. We thank Dr. Robert Berstan and Dr. Ian Bull of the Bristol Node
thermal expansion of ocean waters, may have been initiated 20 to of the Natural Environment Research Council (NERC) Life Sciences Mass
200 kyr prior to the CIE (Sluijs et al., 2008a). Such offsets suggest that the Spectrometry Facility for technical assistance. Discussions and com-
massive carbon input invoked as the cause of the warming during the ments from Chris Hollis and Ian Raine greatly improved the manuscript.
PETM may not have been geologically instantaneous. The Kumara record EMC acknowledges research support by the Royal Society of New
allows these hypotheses to be explored further. If the sea level rise at Zealand Marsden Fast-Start Fund (contract 03-GNS-001) and the GNS
Kumara-2 is due to thermal expansion of ocean waters, then the sudden Science Global Change Through Time Programme (Foundation for
transition to marine depositional conditions would be the result of Research, Science and Technology contract C05X0701). We also thank
increased ocean water temperatures. Geochemical and palynological the NERC for funding LH's PhD. We also thank Professor Thierry Corrège,
samples at 1737.97 m and 1737.02 mbs are the first combined analyses Professor Scott Wing and two anonymous referees for their thorough
to record marine sedimentation. While both samples have lower δ13C critical evaluations and comments that helped improve this manuscript.
values than underlying sediments, they do not reflect the most negative
part of the CIE. The lack of a well-constrained age model and
Appendix A. Supplementary data
sedimentation rates makes detailed interpretation of the timing of the
changes associated with the PETM difficult. However, the evidence
Supplementary data to this article can be found online at doi:10.1016/
suggests the sea level rise, and inferred increase in temperature,
j.palaeo.2011.03.001.
occurred prior to the maximum negative carbon isotopic shift. The rise
in sea level required to flood the study site and whether such a rise was
rapid or gradual is unclear from the Kumara-2 core. These changes References
therefore either coincided with the first input of 13C-depleted carbon Bains, S., Corfieid, R.M., Norris, R.D., 1999. Mechanisms of climate warming at the end of
into the atmosphere or slightly preceded it. Apectodinium dinocysts the Paleocene. Science 285, 724–727.
Bi, X.H., Sheng, G.Y., Liu, X.H., Li, C., Fu, J.M., 2005. Molecular and carbon and hydrogen isotopic
occur in the very first marine sediments, and the increase in their relative
composition of n-alkanes in plant leaf waxes. Organic Geochemistry 36, 1405–1417.
abundance also precedes the occurrence of the lowest n-alkane δ13C Bolle, M.P., Adatte, T., 2001. Palaeocene–early Eocene climatic evolution in the Tethyan
values. Similarly, initiation of noticeable changes in the plant commu- realm: clay mineral evidence. Clay Minerals 36, 249–261.
nity, as recorded by palynomorphs and oleanane concentrations, seem Bolle, M.P., Pardo, A., Hinrichs, K.U., Adatte, T., von Salis, K., Burns, S., Keller, G., Muzylev,
N., 2000. The Paleocene–Eocene transition in the marginal northeastern Tethys
to occur prior to the most negative n-alkane δ13C values, further (Kazakhstan and Uzbekistan). International Journal of Earth Sciences 89, 390–414.
suggesting an offset between climatic perturbations and the peak of the Bowen, G.J., 2009. The Online Isotopes in Precipitation Calculator, version 2.2.
198 L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200

Bowen, G.J., Revenaugh, J., 2003. Interpolating the isotopic composition of modern Gelpi, V., Schneider, H., Mann, J., Oro, J., 1970. Hydrocarbons of geochemical significance
meteoric precipitation. Water Resources Research 39 doi:10.129/2003WR002086. in microscopic algae. Phytochemistry 9, 603–612.
Bowen, G.J., Clyde, W.C., Koch, P.L., Ting, S.Y., Alroy, J., Tsubamoto, T., Wang, Y.Q., Wang, Y., Ghosh, A., Misra, S., Dutta, A.K., Choudhury, A., 1985. Pentacyclic triterpenoids and
2002. Mammalian dispersal at the Paleocene/Eocene boundary. Science 295, 2062–2065. sterols from 7 species of mangrove. Phytochemistry 24, 1725–1727.
Bowen, G.J., Beerling, D.J., Koch, P.L., Zachos, J.C., Quattlebaum, T., 2004. A humid climate Gibson, T.G., Bybell, L.M., Owens, J.P., 1993. Latest Paleocene lithologic and biotic events
state during the Palaeocene/Eocene thermal maximum. Nature 432, 495–498. in neritic deposits of southwestern New Jersey. Paleoceanography 8, 495–514.
Bralower, T.J., Zachos, J.C., Thomas, E., Parrow, M., 1995. Late Paleocene to Eocene Gibson, T.G., Bybell, L.M., Mason, D.B., 2000. Stratigraphic and climatic implications of
paleoceanography of the equatorial Pacific Ocean: stable isotopes recorded at clay mineral changes around the Paleocene/Eocene boundary of the northeastern
Ocean Drilling Program Site 865, Allison Guyot. Paleoceanography 10, 841–865. US margin. Sedimentary Geology 134, 65–92.
Bray, E.E., Evans, E.D., 1961. Distributions of n-paraffins as a clue to recognition of Hancock, H.J.L., Dickens, G.R., Strong, C.P., Hollis, C.J., Field, B.D., 2003. Foraminiferal and
source beds. Geochimica Et Cosmochimica Acta 22, 2–15. carbon isotope stratigraphy through the Paleocene–Eocene transition at Dee
Buffett, B., Archer, D., 2004. Global inventory of methane clathrate: sensitivity to Stream, Marlborough, New Zealand. New Zealand Journal of Geology and
changes in the deep ocean. Earth and Planetary Science Letters 227, 185–199. Geophysics 46, 1–19.
Bujak, J.P., Brinkhuis, H., 1998. Global warming and dinocyst changes across the Handley, L., Pearson, P.N., McMillan, I.K., Pancost, R.D., 2008. Large terrestrial and
Paleocene/Eocene Epoch boundary. In: Aubry, M.P. (Ed.), Late Paleocene–Early marine carbon and hydrogen isotope excursions in a new Paleocene/Eocene
Eocene Climatic and Biotic Events in the Marine and Terrestrial Records. Columbia boundary section from Tanzania. Earth and Planetary Science Letters 275, 17–25.
University Press, New York, pp. 277–295 (Ed. by Aubry, M.P.). Harris, S.A., Whiticar, M.J., Fowler, M.G., 2003. Classification of Duvernay sourced oils
Burnham, A.K., Clarkson, J.E., Singleton, M.F., Wong, C.M., Crawford, R.W., 1982. Biological from central and southern Alberta using Compound Specific Isotope Correlation
markers from Green River kerogen decomposition. Geochimica Et Cosmochimica Acta (CSIC). Bulletin of Canadian Petroleum Geology 51, 99–125.
46, 1243–1251. Hasegawa, T., Yamamoto, C., Pratt, L.M., 2006. Data report: stable carbon isotope
Carter, M., Kelly, C., Hillyer, M., McDowell, P., 1986. Kumara-2, -2A, well completion report. fluctuations of long-chain n-alkanes from Leg 208 Hole 1263A across the
Petroleum Report Series PR, 1183. Ministry of Economic Development, p. 714. Paleocene/Eocene boundary. In: Kroon, D., Zachos, J.C., Richter, C. (Eds.),
Chikaraishi, Y., Naraoka, H., 2003. Compound-specific δD–δ13C analyses of n-alkanes Proceedings of the Ocean Drilling Program: Scientific Results, 208 (Ed. by Kroon,
extracted from terrestrial and aquatic plants. Phytochemistry 63, 361–371. D., Zachos, J.C., Richter, C.).
Clayton, C.J., 1991. Effect of maturity on carbon isotope ratios of oils and condensates. Heywood, V.H. (Ed.), 1978. Flowering Plants of the World. Oxford University Press.
Organic Geochemistry 17, 887–899. Hollis, C.J., Dickens, G.R., Field, B.D., Jones, C.M., Strong, C.P., 2005. The Paleocene–Eocene
Clayton, C.J., Bjorøy, M., 1994. Effect of maturity on 13C/12C ratios of individual transition at Mead Stream, New Zealand: a southern Pacific record of early Cenozoic
compounds in North Sea oils. Organic Geochemistry 21, 737–750. global change. Palaeogeography, Palaeoclimatology, Palaeoecology 215, 313–343.
Collinson, M.E., 1993. Taphonomy and fruiting biology of recent and fossil Nypa. Special Hollis, C.J., Handley, L., Taylor, K., Pancost, R.D., Schouten, S., Creech, J., Baker, J., Andrew,
Papers in Paleontology 49, 165–180. B., Nelson, C.S., Lurcock, P., Wilson, G.S., Schiøler, P., Morgans, H.E.G., Crouch, E.M.,
Collinson, M.E., Steart, D.C., Scott, A.C., Glasspool, I.J., Hooker, J.J., 2007. Episodic fire, Crampton, J.S., Strong, C.P., Huber, M., 2009. Ice in the greenhouse: New Zealand's
runoff and deposition at the Palaeocene–Eocene boundary. Journal of the evidence for Antarctic glaciation in the Late Paleocene. In: Barrel, D., Tulloch, A.
Geological Society 164, 87–97. (Eds.), Geological Society of New Zealand and New Zealand Geophysical Society
Collister, J.W., Lichtfouse, E., Hieshima, G., Hayes, J.M., 1994a. Partial resolution of sources of Joint Annual Conference, November 2009: Geological Society of New Zealand
n-alkanes in the saline portion of the Parachute Creek Member, Green River Formation Miscellaneous Publication, 128A, p. 91 (Ed. by Barrel, D., Tulloch, A.).
(Piceance Creek Basin, Colorado). Organic Geochemistry 21, 645–659. Kaiho, K., Arinobu, T., Ishiwatari, R., Morgans, H.E.G., Okada, H., Takeda, N., Tazaki, K.,
Collister, J.W., Rieley, G., Stern, B., Eglinton, G., Fry, B., 1994b. Compound-specific δ13C Zhou, G.P., Kajiwara, Y., Matsumoto, R., Hirai, A., Niitsuma, N., Wada, H., 1996. Latest
analyses of leaf lipids from plants with differing carbon dioxide metabolisms. Paleocene benthic foraminiferal extinction and environmental changes at Tawanui,
Organic Geochemistry 21, 619–627. New Zealand. Paleoceanography 11, 447–465.
Cooper, R.A. (Ed.), 2004. The New Zealand geological timescale: Institute of Geological Kelly, D.C., Zachos, J.C., Bralower, T.J., Schellenberg, S.A., 2005. Enhanced terrestrial
and Nuclear Sciences Monograph. weathering/runoff and surface ocean carbonate production during the recovery
Crouch, E.M., Brinkhuis, H., 2005. Environmental change across the Paleocene–Eocene stages of the Paleocene–Eocene thermal maximum. Paleoceanography 20,
transition from eastern New Zealand: a marine palynological approach. Marine doi:10.1029/2005PA001163 .
Micropalaeontology 56, 138–160. Kennett, J.P., Stott, L.D., 1991. Abrupt deep-sea warming, palaeoceanographic changes
Crouch, E.M., Heilmann-Clausen, C., Brinkhuis, H., Morgans, H.E.G., Rogers, K.M., Egger, and benthic extinctions at the end of the Palaeocene. Nature 353, 225–229.
H., Schmitz, B., 2001. Global dinoflagellate event associated with the late Paleocene Killops, S.D., Frewin, N.L., 1994. Triterpenoid diagenesis and cuticular preservation.
thermal maximum. Geology 29, 315–318. Organic Geochemistry 21, 1193–1209.
Crouch, E.M., Dickens, G.R., Brinkhuis, H., Aubry, M.P., Hollis, C.J., Rogers, K.M., Visscher, H., Koch, P.L., Zachos, J.C., Gingerich, P.D., 1992. Correlation between isotope records in
2003. The Apectodinium acme and terrestrial discharge during the Paleocene–Eocene marine and continental carbon reservoirs near the Palaeocene/Eocene boundary.
thermal maximum: new palynological, geochemical and calcareous nannoplankton Nature 358, 319–322.
observations at Tawanui, New Zealand. Palaeogeography Palaeoclimatology Palaeoeco- Koch, P.L., Zachos, J.C., Dettman, D.L., 1995. Stable isotope stratigraphy and paleoclimatology
logy 194, 387–403. of the Paleogene Bighorn Basin (Wyoming, USA). Palaeogeography, Palaeoclimatology,
Dawson, D., Grice, K., Alexander, R., 2005. Effect of maturation on the indigenous δD Palaeoecology 115, 61–89.
signatures of individual hydrocarbons in sediments and crude oils from the Perth Koch, B.P., Rullkötter, J., Lara, R.J., 2003a. Evaluation of triterpenols and sterols as
Basin (Western Australia). Organic Geochemistry 36, 95–104. organic matter biomarkers in a mangrove ecosystem in northern Brazil. Wetlands
Dawson, D., Grice, K., Alexander, R., Edwards, D., 2007. The effect of source and maturity on Ecology and Management 11, 257–263.
the stable isotopic compositions of individual hydrocarbons in sediments and crude Koch, P.L., Clyde, W.C., Hepple, R.P., Fogel, M.L., Wing, S.L., Zachos, J.C., 2003b. Carbon
oils from the Vulcan Sub-basin, Timor Sea, Northern Australia. Organic Geochemistry and oxygen isotope records from paleosols spanning the Paleocene–Eocene
38, 1015–1038. boundary, Bighorn Basin, Wyoming. In: Wing, S.L., Gingerich, P.D., Schmitz, B.,
DeConto, R.M., Pollard, D., 2003. A coupled climate-ice sheet modeling approach to the Thomas, E. (Eds.), Causes and Consequences of Globally Warm Climates in the Early
Early Cenozoic history of the Antarctic ice sheet. Palaeogeography, Palaeoclimatology, Paleogene: Geological Society of America Special Paper, 369, pp. 49–64 (Ed. by
Palaeoecology 198, 39–52. Wing, S.L., Gingerich, P.D., Schmitz, B., Thomas, E.).
Didyk, B.M., Simoneit, B.R.T., Brassell, S.C., Eglinton, G., 1978. Organic geochemical Kolattukudy, P.E., 1976. The Chemistry and Biochemistry of Natural Waxes. Elsevier,
indicators of palaeoenvironmental conditions of sedimentation. Nature 272, 216–222. Amsterdam.
Diefendorf, A.F., Mueller, K.E., Wing, S.L., Koch, P.L., Freeman, K.H., 2010. Global patterns in leaf Kraus, M.J., Riggins, S., 2007. Transient drying during the Paleocene–Eocene Thermal
13
C discrimination and implications for studies of past and future climate. Proceedings of Maximum (PETM): analysis of paleosols in the bighorn basin, Wyoming.
the National Academy of Sciences of the United States of America 107, 5738–5743. Palaeogeography, Palaeoclimatology, Palaeoecology 245, 444–461.
Duan, Y., Wang, Z., 2002. Evidence from carbon isotope measurements for biological Kurtz, A.C., Kump, L.R., Arthur, M.A., Zachos, J.C., Paytan, A., 2003. Early Cenozoic decoupling of
origins of individual longchain n-alkanes in sediments from the Nansha Sea, China. the global carbon and sulfur cycles. Paleoceanography 18, doi:10.1029/2003PA000908 .
Chinese Science Bulletin 47, 578–581. Lear, C.H., Elderfield, H., Wilson, P.A., 2000. Cenozoic deep-sea temperatures and global
Eglinton, G., Hamilton, R.J., 1967. Leaf epicuticular waxes. Science 156, 1322–1335. ice volumes from Mg/Ca in benthic foraminiferal calcite. Science 287, 269–272.
Eglinton, G., Gonzalez, A.G., Hamilton, R.J., Raphael, R.A., 1962. Hydrocarbon constituents of Leavitt, S.W., Newberry, T., 1992. Systematics of stable-carbon isotopic differences
the wax coatings of plant leaves — a taxanomic survey. Phytochemistry 1, 89–102. between gymnosperm and angiosperm trees. Plant Physiology 11, 257–262.
Eglinton, T.I., BenitezNelson, B.C., Pearson, A., McNichol, A.P., Bauer, J.E., Druffel, E.R.M., Li, M.W., Huang, Y.S., Obermajer, M., Jiang, C.Q., Snowdon, L.R., Fowler, M.G., 2001.
1997. Variability in radiocarbon ages of individual organic compounds from marine Hydrogen isotopic compositions of individual alkanes as a new approach to
sediments. Science 277, 796–799. petroleum correlation: case studies from the Western Canada Sedimentary Basin.
Farquhar, G.D., Ehleringer, J.R., Hubick, K.T., 1989. Carbon isotope discrimination and Organic Geochemistry 32, 1387–1399.
photosynthesis. In: Briggs, W.R., Jones, R.L., Walbot, V. (Eds.), Annual Review of Lippert, P.C., Zachos, J.C., 2007. A biogenic origin for anomalous fine-grained magnetic
Plant Physiology and Plant Molecular Biology, 40. Annual Reviews Inc, Palo Alto material at the Paleocene–Eocene boundary at Wilson Lake, New Jersey.
USA, pp. 503–537 (Ed. by Briggs, W.R., Jones, R.L., Walbot, V.). Paleoceanography 22, doi:10.1029/2007PA001471 .
Flanagan, L.B., Brooks, J.R., Ehleringer, J.R., 1997. Photosynthesis and carbon isotope Liu, Z., Huang, Y., 2008. Hydrogen isotopic compositions of plant leaf lipids are
discrimination in boreal forest ecosystems: a comparison of functional characteristics unaffected by a twofold pCO2 change in growth chambers. Organic Geochemistry
in plants from three mature forest types. Journal of Geophysical Research-Atmospheres 39, 478–482.
102, 28861–28869. Magioncalda, R., Dupuis, C., Smith, T., Steurbaut, E., Gingerich, P.D., 2004. Paleocene–Eocene
Gat, J.R., 1996. Oxygen and hydrogen isotopes in the hydrologic cycle. Annual Review of carbon isotope excursion in organic carbon and pedogenic carbonate: direct
Earth and Planetary Sciences 24, 225–262. comparison in a continental stratigraphic section. Geology 32, 553–556.
L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200 199

Miller, K.G., Wright, J.D., Browning, J.V., 2005. Visions of ice sheets in a greenhouse Rullkötter, J., Peakman, T.M., Lo ten Haven, H., 1994. Early diagenesis of terrigenous
world. Marine Geology 217, 215–231. triterpenoids and its implications for petroleum geochemistry. Organic Geochem-
Moldowan, J.M., Sundararaman, P., Schoell, M., 1986. Sensitivity of biomarker istry 21, 215.
properties to depositional environment and or source input in the Lower Toarcian Sachse, D., Radke, J., Gleixner, G., 2004. Hydrogen isotope ratios of recent lacustrine
of southwest Germany. Organic Geochemistry 10, 915–926. sedimentary n-alkanes record modern climate variability. Geochimica Et Cosmo-
Moldowan, J.M., Dahl, J., Huizinga, B.J., Fago, F.J., Hickey, L.J., Peakman, T.M., Taylor, D.W., chimica Acta 68, 4877–4889.
1994. The molecular fossil record of oleanane and its relation to angiosperms. Sachse, D., Radke, J., Gleixner, G., 2006. δD values of individual n-alkanes from
Science 265, 768–771. terrestrial plants along a climatic gradient — implications for the sedimentary
Naraoka, H., Ishiwatari, R., 1999. Carbon isotopic compositions of individual long-chain biomarker record. Organic Geochemistry 37, 469–483.
n-fatty acids and n-alkanes in sediments from river to open ocean: multiple origins Sauer, P.E., Eglinton, T.I., Hayes, J.M., Schimmelmann, A., Sessions, A.L., 2001. Compound-
for their occurrence. Geochemical Journal 33, 215–235. specific D/H ratios of lipid biomarkers from sediments as a proxy for environmental
Nicolo, M.J., Dickens, G.R., Hollis, C.J., Zachos, J.C., 2007. Multiple early Eocene and climatic conditions. Geochimica Et Cosmochimica Acta 65, 213–222.
hyperthermals: their sedimentary expression on the New Zealand continental Scalan, R.S., Smith, J.E., 1970. An improved measure of the odd-to-even predominance
margin and in the deep sea. Geology 35, 699–702. in the normal alkanes of sediment extracts and petroleum. Geochimica Et
Osundina, M.A., 1997. Nodulation and growth of mycorrhizal Casuarina equisetifolia J.R. Cosmochimica Acta 34, 611–620.
and G. First in response to flooding. Biology and Fertility of Soils 26, 95–99. Schimmelmann, A., Sessions, A.L., Boreham, C.J., Edwards, D.S., Logan, G.A., Summons, R.E.,
Ourisson, G., Albrecht, P., Rohmer, M., 1979. The hopanoids. Palaeochemistry and 2004. D/H ratios in terrestrially sourced petroleum systems. Organic Geochemistry 35,
biochemistry of a group of natural products. Pure and Applied Chemistry 51, 709–729. 1169–1195.
Pagani, M., Caldeira, K., Archer, D., Zachos, J.C., 2006a. An ancient carbon mystery. Schmitz, B., Pujalte, V., 2003. Sea-level, humidity, and land-erosion records across the
Science 314, 1556–1557. initial Eocene thermal maximum from a continental-marine transect in northern
Pagani, M., Pedentchouk, N., Huber, M., Sluijs, A., Schouten, S., Brinkhuis, H., Sinninghe Spain. Geology 31, 689–692.
Damsté, J.S., Dickens, G.R., Scientists, Expedition, 2006b. Arctic hydrology during Schmitz, B., Pujalte, V., 2007. Abrupt increase in seasonal extreme precipitation at the
global warming at the Palaeocene/Eocene thermal maximum. Nature 442, 671–675. Paleocene–Eocene boundary. Geology 35, 215–218.
Panchuk, K., Ridgwell, A., Kump, L.R., 2008. Sedimentary response to Paleocene–Eocene Schouten, S., Woltering, M., Rijpstra, W.I.C., Sluijs, A., Brinkhuis, H., Sinninghe Damsté, J.S.,
Thermal Maximum carbon release: a model-data comparison. Geology 36, 315–318. 2007. The Paleocene–Eocene carbon isotope excursion in higher plant organic matter:
Pedentchouk, N., Freeman, K.H., Harris, N.B., 2006. Different response of δD values of differential fractionation of angiosperms and conifers in the Arctic. Earth and Planetary
n-alkanes, isoprenoids, and kerogen during thermal maturation. Geochimica et Science Letters 258, 581–592.
Cosmochimica Acta 70, 2063–2072. Schrag, D.P., Adkins, J.F., McIntyre, K., Alexander, J.L., Hodell, D.A., Charles, C.D.,
Pedentchouk, N., Wagner, T., Jones, M., Pellegrini, M., Brugnoli, E., Lauteri, M., Pollegioni, McManus, J.F., 2002. The oxygen isotopic composition of seawater during the Last
P., Behling, H., 2007. Comparison of δ13C and δD values of n-alkanes from Glacial Maximum. Quaternary Science Reviews 21, 331–342.
angiosperms and gymnosperms in Western Europe. Geochimica Et Cosmochimica Seifert, W.K., Moldowan, J.M., 1980. The effect of thermal stress on source rock quality as
Acta 71, A770-A770. measured by hopane stereochemistry. Physics and Chemistry of the Earth 12, 229–237.
Pedentchouk, N., Sumner, W., Tipple, B., Pagani, M., 2008. δ13C and δD composition of n- Shackleton, N.J., 1977. Oxygen isotope stratigraphic record of Late Pleistocene.
alkanes from modern angiosperms and conifers: an experimental set up in central Philosophical Transactions of the Royal Society of London Series B-Biological
Washington State, USA. Organic Geochemistry 39, 1066–1071. Sciences 280, 169–182.
Peters, K.E., Moldowan, J.M., 1991. Effects of source, thermal maturity, and Shackleton, N.J., 1986. Paleogene stable isotope events. Palaeogeography, Palaeoclima-
biodegradation on the distribution and isomerization of homohopanes in tology, Palaeoecology 57, 91–102.
petroleum. Organic Geochemistry 17, 47–61. Shackleton, N.J., Kennett, J.P., 1975. Paleotemperature history of the Cenozoic and the
Peters, K.E., Clark, M.E., Dasgupta, U., McCaffrey, M.A., Lee, C.Y., 1995. Recognition of an initiation of Antarctic glaciation: oxygen and carbon isotope analyses in DSDP sites
Infracambrian source-rock based on biomarkers in the Bahewala-1 Oil, India. 277, 279, and 281. Initial Reports of the Deep Sea Drilling Project 29, 743–755.
American Association of Petroleum Geologists Bulletin 79, 1481–1494. Sinninghe Damsté, J.S., de Leeuw, J.W., 1990. Analysis, structure and geochemical
Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide. Cambridge significance of organically-bound sulfur in the geosphere — state-of-the-art and
University Press, Cambridge. future-research. Organic Geochemistry 16, 1077–1101.
Pocknall, D.T., 1990. Palynological evidence for the early to middle Eocene vegetation and Sinninghe Damsté, J.S., Rijpstra, W.I.C., de Leeuw, J.W., Schenck, P.A., 1988. Origin of organic
climate history of New Zealand. Review of Palaeobotany and Palynology 65, 57–69. sulfur-compounds and sulfur-containing high molecular-weight substances in
Popp, B.N., Takigiku, R., Hayes, J.M., Louda, J.W., Baker, E.W., 1989. The post-Paleozoic sediments and immature crude oils. Organic Geochemistry 13, 593–606.
chronology and mechanism of 13C depletion in primary marine organic-matter. Sluijs, A., Pross, J., Brinkhuis, H., 2005. From greenhouse to icehouse; organic-walled
American Journal of Science 289, 436–454. dinoflagellate cysts as paleoenvironmental indicators in the Paleogene. Earth
Powell, T.G., McKirdy, D.M., 1973. Relationship between ratio of pristane to phytane, Science Reviews 68, 281–315.
crude-oil composition and geological environment in Australia. Nature-Physical Sluijs, A., Schouten, S., Pagani, M., Woltering, M., Brinkhuis, H., Sinninghe Damsté, J.S., Dickens,
Science 243, 37–39. G.R., Huber, M., Reichart, G.J., Stein, R., 2006. Subtropical Arctic Ocean temperatures
Radke, J., Bechtel, A., Gaupp, R., Puttmann, W., Schwark, L., Sachse, D., Gleixner, G., 2005. during the Palaeocene/Eocene thermal maximum. Nature 441, 610–613.
Correlation between hydrogen isotope ratios of lipid biomarkers and sediment Sluijs, A., Brinkhuis, H., Schouten, S., Bohaty, S.M., John, C.M., Zachos, J.C., Reichart, G.J.,
maturity. Geochimica Et Cosmochimica Acta 69, 5517–5530. Sinninghe Damsté, J.S., Crouch, E.M., Dickens, G.R., 2007. Environmental precursors
Raine, J.I., 1984. Outline of a palynological zonation of Cretaceous to Paleogene to rapid light carbon injection at the Palaeocene/Eocene boundary. Nature 450,
terrestrial sediments in West Coast region, South Island, New Zealand. Report New 1218–1221.
Zealand Geological Survey, p. 82. Sluijs, A., Brinkhuis, H., Crouch, E.M., John, C.M., Handley, L., Munsterman, D., Bohaty, S.M.,
Raine, J.I., 1986. Palynostratigraphy of Sub-Kaiata Formation sequence, Petrocorp Zachos, J.C., Reichart, G.J., Schouten, S., Pancost, R.D., Sinninghe Damsté, J.S., Welters, N.
Kumara-2 well, Westland. New Zealand Geological Survey Report, p. 13. L.D., Lotter, A.F., Dickens, G.R., 2008a. Eustatic variations during the Paleocene–Eocene
Rao, Z., Zhu, Z., Jia, G., Henderson, A.C.G., Xue, Q., Wang, S., 2009. Compound specific δD greenhouse world. Paleoceanography 23, doi:10.1029/2008PA001615 .
values of long chain n-alkanes derived from terrestrial higher plants are indicative Sluijs, A., Röhl, U., Schouten, S., Brumsack, H.J., Sangiorgi, F., Sinninghe Damsté, J.S.,
of the δD of meteoric waters: evidence from surface soils in eastern China. Organic Brinkhuis, H., 2008b. Arctic late Paleocene–early Eocene paleoenvironments with
Geochemistry 40, 922–930. special emphasis on the Paleocene–Eocene thermal maximum (Lomonosov Ridge,
Rieley, G., Collier, R.J., Jones, D.M., Eglinton, G., Eakin, P.A., Fallick, A.E., 1991. Sources of Integrated Ocean Drilling Program Expedition 302). Paleoceanography 23,
sedimentary lipids deduced from stable carbon isotope analyses of individual doi:10.1029/2007PA001495 .
compounds. Nature 352, 425–427. Smith, F.A., Freeman, K.H., 2006. Influence of physiology and climate on δD of leaf wax
Robert, C., Chamley, H., 1991. Development of early Eocene warm climates, as inferred n-alkanes from C3 and C4 grasses. Geochimica Et Cosmochimica Acta 70, 1172–1187.
from clay mineral variations in oceanic sediments. Global and Planetary Change 89, Smith, F.A., Wing, S.L., Freeman, K.H., 2007. Magnitude of the carbon isotope excursion
315–331. at the Paleocene–Eocene thermal maximum: the role of plant community change.
Robert, C., Kennett, J.P., 1994. Antarctic subtropical humid episode at the Paleocene–Eocene Earth and Planetary Science Letters 262, 50–65.
boundary: clay-mineral evidence. Geology 22, 211–214. Smittenberg, R.H., Eglinton, T.I., Schouten, S., Sinninghe Damsté, J.S., 2006. Ongoing
Röhl, U., Westerhold, T., Bralower, T.J., Zachos, J.C., 2007. On the duration of the buildup of refractory organic carbon in boreal soils during the Holocene. Science
Paleocene–Eocene thermal maximum (PETM). Geochemistry Geophysics Geosys- 314, 1283–1286.
tems 8, doi:10.1029/2007GC001784 . Speijer, R.P., Morsi, A.M.M., 2002. Ostracode turnover and sea-level changes associated
Rohmer, M., Bouviernave, P., Ourisson, G., 1984. Distribution of hopanoid triterpenes in with the Paleocene–Eocene thermal maximum. Geology 30, 23–26.
prokaryotes. Journal of General Microbiology 130, 1137–1150. Speijer, R.P., Wagner, T., 2002. Sea-level changes and black shales associated with the
Rontani, J.F., Volkman, J.K., 2003. Phytol degradation products as biogeochemical late Paleocene thermal maximum: organic-geochemical and micropaleontologic
tracers in aquatic environments. Organic Geochemistry 34, 1–35. evidence from the southern Tethyan margin (Egypt–Israel). In: Koeberl, C.,
Rooney, M.A., Vuletich, A.K., Griffith, C.E., 1998. Compound-specific isotope analysis as a MacLeod, K.G. (Eds.), Catastrophic Events and Mass Extinctions: Impacts and
tool for characterizing mixed oils: an example from the West of Shetlands area. Beyond: Geological Society of America Special Paper, 356, pp. 533–549 (Ed. by
Organic Geochemistry 29, 241–254. Koeberl, C., MacLeod, K.G.).
Rowland, S.J., 1990. Production of acyclic isoprenoid hydrocarbons by laboratory Storey, M., Duncan, R.A., Swisher, C.C., 2007. Paleocene–Eocene thermal maximum and
maturation of methanogenic bacteria. Organic Geochemistry 15, 9–16. the opening of the northeast Atlantic. Science 316, 587–589.
Rullkötter, J., Aizenshtat, Z., Spiro, B., 1984. Biological markers in bitumens and Svensen, H., Planke, S., Malthe-Sorenssen, A., Jamtveit, B., Myklebust, R., Eidem, T.R.,
pyrolyzates of Upper Cretaceous bituminous chalks from the Gahreb Formation Rey, S.S., 2004. Release of methane from a volcanic basin as a mechanism for initial
(Israel). Geochimica Et Cosmochimica Acta 48. Eocene global warming. Nature 429, 542–545.
200 L. Handley et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 305 (2011) 185–200

Ten Haven, H.L., Rullkötter, J., 1988. The diagenetic fate of taraxer-14-ene and oleanene Wing, S.L., Harrington, G.J., Smith, F.A., Bloch, J.I., Boyer, D.M., Freeman, K.H., 2005.
isomers. Geochimica Et Cosmochimica Acta 52, 2543–2548. Transient floral change and rapid global warming at the Paleocene–Eocene
Ten Haven, H.L., de Leeuw, J.W., Rullkötter, J., Sinninghe Damsté, J.S., 1987. Restricted utility of boundary. Science 310, 993–996.
the pristane/phytane ratio as a palaeoenvironmental indicator. Nature 330, 641–643. Zachos, J.C., Wara, M.W., Bohaty, S., Delaney, M.L., Petrizzo, M.R., Brill, A., Bralower, T.J.,
Thiry, H., Dupuis, C., 2000. Use of clay minerals for paleoclimatic reconstructions: limits of the Premoli-Silva, I., 2003. Paleoclimate: a transient rise in tropical sea surface
method with special reference to the Paleocene–lower Eocene interval. Gff 122, 166–167. temperature during the Paleocene–Eocene Thermal Maximum. Science 302,
Thomas, D.J., Bralower, T.J., Zachos, J.C., 1999. New evidence for subtropical warming 1551–1553.
during the late Paleocene thermal maximum: stable isotopes from Deep Sea Zachos, J.C., Röhl, U., Schellenberg, S.A., Sluijs, A., Hodell, D.A., Kelly, D.C., Thomas, E.,
Drilling Project Site 527, Walvis Ridge. Paleoceanography 14, 561–570. Nicolo, M., Raffi, I., Lourens, L.J., McCarren, H., Kroon, D., 2005. Rapid acidification of
Thomas, D.J., Zachos, J.C., Bralower, T.J., Thomas, E., Bohaty, S., 2002. Warming the fuel the ocean during the Paleocene–Eocene Thermal Maximum. Science 308,
for the fire: evidence for the thermal dissociation of methane hydrate during the 1611–1614.
Paleocene–Eocene thermal maximum. Geology 30, 1067–1070. Zachos, J.C., Schouten, S., Bohaty, S., Quattlebaum, T., Sluijs, A., Brinkhuis, H., Gibbs, S.J.,
Tripati, A.K., Elderfield, H., 2005. Deep-sea temperature and circulation changes at the Bralower, T.J., 2006. Extreme warming of mid-latitude coastal ocean during the
Paleocene–Eocene Thermal Maximum. Science 308, 1894–1897. Paleocene–Eocene Thermal Maximum: inferences from TEX86 and isotope data.
Waples, D.W., Machihara, T., 1991. Biomarkers for Geologists. OK, Tulsa. Geology 34, 737–740.
Weijers, J.W.H., Schouten, S., Sluijs, A., Brinkhuis, H., Sinninghe Damsté, J.S., 2007. Zeebe, R.E., Zachos, J.C., Dickens, G.R., 2009. Carbon dioxide forcing alone insufficient to
Warm arctic continents during the Palaeocene–Eocene thermal maximum. Earth explain Palaeocene–Eocene Thermal Maximum warming. Nature Geoscience 2,
and Planetary Science Letters 261, 230–238. 576–580.

You might also like