You are on page 1of 35

Paleocene–Eocene Thermal Maximum

The Paleocene–Eocene thermal


maximum (PETM), alternatively
"Eocene thermal maximum 1" (ETM1),
and formerly known as the
"Initial  Eocene" or
"Late Paleocene thermal maximum",
was a time period with a more than 5–8 °C
global average temperature rise across the
event.[1][2] This climate event occurred at
the time boundary of the Paleocene and
Eocene geological epochs.[3] The exact age
and duration of the event is uncertain but it
is estimated to have occurred around 55.5 Climate change during the last 65 million years as expressed
million years ago (Ma).[4] by the oxygen isotope composition of benthic foraminifera. The
Paleocene-Eocene thermal maximum (PETM) is characterized
The associated period of massive carbon by a brief but prominent excursion, attributed to rapid warming.
release into the atmosphere has been Note that the excursion is understated in this graph due to the
estimated to have lasted from 20,000 to smoothing of data.
50,000 years. The entire warm period
lasted for about 200,000 years. Global
temperatures increased by 5–8 °C.[2]

The onset of the Paleocene–Eocene thermal maximum has been linked to volcanism[1] and uplift associated
with the North Atlantic Igneous Province, causing extreme changes in Earth's carbon cycle and a
significant temperature rise.[2][5][6] The period is marked by a prominent negative excursion in carbon
stable isotope (δ13 C) records from around the globe; more specifically, there was a large decrease in
13 C/12 C ratio of marine and terrestrial carbonates and organic carbon.[2][7][8] Paired δ13 C, δ11 B, and

δ18 O data suggest that ~12 000 Gt of carbon (at least 44 000 Gt CO2 e) were released over 50,000 years,[5]
averaging 0.24 Gt per year.

Stratigraphic sections of rock from this period reveal numerous other changes.[2] Fossil records for many
organisms show major turnovers. For example, in the marine realm, a mass extinction of benthic
foraminifera, a global expansion of subtropical dinoflagellates, and an appearance of excursion, planktic
foraminifera and calcareous nanofossils all occurred during the beginning stages of PETM. On land,
modern mammal orders (including primates) suddenly appear in Europe and in North America.[9]

Setting Paleogene graphical timeline


The configuration of oceans and continents
Neogene
was somewhat different during the early –  
Paleogene relative to the present day. The −25 — O
Panama Isthmus did not yet connect North l Chattian
i
America and South America, and this – g
o
allowed direct low-latitude circulation −30 — c
e Rupelian
between the Pacific and Atlantic Oceans. n
e First Antarctic
permanent ice
– ← permanent ice-
The Drake Passage, which now separates −35 — [10]
sheets
South America and Antarctica, was closed, Priabonian
and this perhaps prevented thermal isolation –C  
of Antarctica. The Arctic was also more e
−40 — P Bartonian
restricted. Although various proxies for past naE
–o l o
atmospheric CO2 levels in the Eocene do z c
not agree in absolute terms, all suggest that −45 — o e e Lutetian  
o
levels then were much higher than at –i gn
present. In any case, there were no c e
−50 —
significant ice sheets during this time.[12]
– Ypresian
e
Earth surface temperatures increased by −55 — n
about 6 °C from the late Paleocene through e ← PETM
[12] – P Thanetian
the early Eocene. Superimposed on this a
l
long-term, gradual warming were at least −60 — e Selandian
o
two (and probably more) "hyperthermals". c
– e
These can be defined as geologically brief n Danian
e
(<200,000 year) events characterized by −65 —
  ← K-Pg mass
extinction
rapid global warming, major changes in the M
Z Cretaceous
environment, and massive carbon addition.
Though not the first within the Subdivision of the Paleogene according to
Cenozoic, [13] the PETM was the most the ICS, as of 2021.[11]
extreme of these hyperthermals. Another Vertical axis scale: millions of years ago
hyperthermal clearly occurred at
approximately 53.7 Ma, and is now called ETM-2 (also referred to as H-1, or the Elmo event). However,
additional hyperthermals probably occurred at about 53.6 Ma (H-2), 53.3 (I-1), 53.2 (I-2) and 52.8 Ma
(informally called K, X or ETM-3).[14] The number, nomenclature, absolute ages, and relative global
impact of the Eocene hyperthermals are the source of considerable current research. Whether they only
occurred during the long-term warming, and whether they are causally related to apparently similar events
in older intervals of the geological record (e.g. the Toarcian turnover of the Jurassic) are open issues.

Global warming
A study in 2020 estimated the Global mean surface temperature (GMST) with 66% confidence during the
latest Paleocene (c. 57 Ma) as 22.3 to 28.3  °C, PETM (56 Ma) as 27.2 to 34.5  °C, and Early Eocene
Climatic Optimum (EECO) (53.3 to 49.1 Ma) as 23.2 to 29.7 °C.[15] Estimates of the amount of average
global temperature rise at the start of the PETM range from approximately 3 to 6 °C[16] to between 5 and
8 °C.[2] This warming was superimposed on "long-term" early Paleogene warming, and is based on several
lines of evidence. There is a prominent (>1‰) negative excursion in the δ18 O of foraminifera shells, both
those made in surface and deep ocean water. Because there was little or no polar ice in the early Paleogene,
the shift in δ18 O very probably signifies a rise in ocean temperature.[17] The temperature rise is also
supported by the spread of warmth-loving taxa to higher latitudes,[18] changes in plant leaf shape and
size,[19] the Mg/Ca ratios of foraminifera,[16] and the ratios of certain organic compounds, such as
TEXH86 .[20]

TEXH86 values indicate that the average sea surface temperature (SST) reached over 36 °C (97 °F) in the
tropics during the PETM, enough to cause heat stress even in organisms resistant to extreme thermal stress,
such as dinoflagellates, of which a significant number of species went extinct.[20] Oxygen isotope ratios
from Tanzania suggest that tropical SSTs may have been even higher, exceeding 40 °C.[21] Low latitude
Indian Ocean Mg/Ca records show seawater at all depths warmed by ~4-5 °C.[22] In the Pacific Ocean,
tropical SSTs increased by about 4-5 °C.[23] TEXL86
values from deposits in New Zealand, then located
between 50°S and 60°S in the southwestern
Pacific,[24] indicate SSTs of 26  °C (79  °F) to 28  °C
(82  °F), an increase of over 10  °C (18  °F) from an
average of 13  °C (55  °F) to 16  °C (61  °F) at the
boundary between the Selandian and Thanetian.[25]
Sediment core samples from the East Tasman Plateau,
then located at a palaeolatitude of ∼65 °S, show an
increase in SSTs from ~26 °C to ~33 °C during the
PETM.[26] In the North Sea, SSTs jumped by 10 °C,
reaching highs of ~33 °C.[27]

Certainly, the central Arctic Ocean was ice-free before,


during, and after the PETM. This can be ascertained
from the composition of sediment cores recovered
during the Arctic Coring Expedition (ACEX) at 87°N
on Lomonosov Ridge.[28] Moreover, temperatures
increased during the PETM, as indicated by the brief
presence of subtropical dinoflagellates,[29] and a
marked increase in TEX86 .[30] The latter record is
intriguing, though, because it suggests a 6  °C (11  °F)
rise from ~17 °C (63 °F) before the PETM to ~23 °C
(73  °F) during the PETM. Assuming the TEX86
record reflects summer temperatures, it still implies
much warmer temperatures on the North Pole
compared to the present day, but no significant
latitudinal amplification relative to surrounding time.

The above considerations are important because, in


many global warming simulations, high latitude A stacked record of temperatures and ice volume
temperatures increase much more at the poles through in the deep ocean through the Mesozoic and
Cenozoic periods.
an ice–albedo feedback.[31] It may be the case,
LPTM— Paleocene-Eocene thermal maximum
however, that during the PETM, this feedback was
OAEs— oceanic anoxic events
largely absent because of limited polar ice, so
MME— mid-Maastrichtian event
temperatures on the Equator and at the poles increased
similarly. Notable is the absence of documented
greater warming in polar regions compared to other
regions. This implies a non-existing ice-albedo feedback, suggesting no sea or land ice was present in the
late Paleocene.[4]

Proxy data from Esplugafereda in northeastern Spain shows a rapid +8 °C temperature rise, in accordance
with existing regional records of marine and terrestrial environments.[32]

Precise limits on the global temperature rise during the PETM and whether this varied significantly with
latitude remain open issues. Oxygen isotope and Mg/Ca of carbonate shells precipitated in surface waters of
the ocean are commonly used measurements for reconstructing past temperature; however, both
paleotemperature proxies can be compromised at low latitude locations, because re-crystallization of
carbonate on the seafloor renders lower values than when formed. On the other hand, these and other
temperature proxies (e.g., TEX86 ) are impacted at high latitudes because of seasonality; that is, the
"temperature recorder" is biased toward summer, and therefore higher values, when the production of
carbonate and organic carbon occurred.

Carbon cycle disturbance


Clear evidence for massive addition of 13 C-depleted carbon at the onset of the PETM comes from two
observations. First, a prominent negative excursion in the carbon isotope composition (δ13 C) of carbon-
bearing phases characterizes the PETM in numerous (>130) widespread locations from a range of
environments. Second, carbonate dissolution marks the PETM in sections from the deep sea.[2]

The total mass of carbon injected to the ocean and atmosphere during the PETM remains the source of
debate. In theory, it can be estimated from the magnitude of the negative carbon isotope excursion (CIE),
the amount of carbonate dissolution on the seafloor, or ideally both.[33][34] However, the shift in the δ13 C
across the PETM depends on the location and the carbon-bearing phase analyzed. In some records of bulk
carbonate, it is about 2 ‰ (per mil); in some records of terrestrial carbonate or organic matter it exceeds
6‰.[2][35] Carbonate dissolution also varies throughout different ocean basins. It was extreme in parts of
the north and central Atlantic Ocean, but far less pronounced in the Pacific Ocean.[34][36][37] With available
information, estimates of the carbon addition range from about 2,000 to 7,000 gigatons.[34][36][37]

Timing of carbon addition and warming


The timing of the PETM δ13 C excursion is of considerable interest. This is because the total duration of the
CIE, from the rapid drop in δ13 C through the near recovery to initial conditions, relates to key parameters
of our global carbon cycle, and because the onset provides insight to the source of 13 C-depleted CO2 .

The total duration of the CIE can be estimated in several ways. The iconic sediment interval for examining
and dating the PETM is a core recovered in 1987 by the Ocean Drilling Program at Hole 690B at Maud
Rise in the South Atlantic Ocean. At this location, the PETM CIE, from start to end, spans about 2 m.
Long-term age constraints, through biostratigraphy and magnetostratigraphy, suggest an average Paleogene
sedimentation rate of about 1.23  cm/1,000yrs. Assuming a constant sedimentation rate, the entire event,
from onset though termination, was therefore estimated at 200,000 years.[7] Subsequently, it was noted that
the CIE spanned 10 or 11 subtle cycles in various sediment properties, such as Fe content. Assuming these
cycles represent precession, a similar but slightly longer age was calculated by Rohl et al. 2000. If a
massive amount of 13 C-depleted CO2 is rapidly injected into the modern ocean or atmosphere and
projected into the future, a ~200,000 year CIE results because of slow flushing through quasi steady-state
inputs (weathering and volcanism) and outputs (carbonate and organic) of carbon.[38] A different study,
based on a revised orbital chronology and data from sediment cores in the South Atlantic and the Southern
Ocean, calculated a slightly shorter duration of about 170,000 years.[39]

A ~200,000 year duration for the CIE is estimated from models of global carbon cycling.[40]

Age constraints at several deep-sea sites have been independently examined using 3 He contents, assuming
the flux of this cosmogenic nuclide is roughly constant over short time periods. This approach also suggests
a rapid onset for the PETM CIE (<20,000 years). However, the 3 He records support a faster recovery to
near initial conditions (<100,000 years) than predicted by flushing via weathering inputs and carbonate and
organic outputs.[41]

There is other evidence to suggest that warming predated the δ13 C excursion by some 3,000 years.[42]
Some authors have suggested that the magnitude of the CIE may be underestimated due to local processes
in many sites causing a large proportion of allochthonous sediments to accumulate in their sedimentary
rocks, contaminating and offsetting isotopic values derived from them.[43] Organic matter degradation by
microbes has also been implicated as a source of skewing of carbon isotopic ratios in bulk organic
matter.[44]

Effects

Precipitation

The climate would also have become much wetter, with the
increase in evaporation rates peaking in the tropics. Deuterium
isotopes reveal that much more of this moisture was transported
polewards than normal.[45] Warm weather would have
predominated as far north as the Polar basin. Finds of fossils of
Azolla floating ferns in polar regions indicate subtropic
temperatures at the poles.[46] Central China during the PETM
hosted dense subtropical forests as a result of the significant
increase in rates of precipitation in the region, with average Azolla floating ferns, fossils of this
temperatures between 21  °C and 24  °C and mean annual genus indicate subtropical weather at
precipitation ranging from 1,396 to 1,997  mm.[47] Very high the North Pole
precipitation is also evidenced in the Cambay Shale Formation of
India by the deposition of thick lignitic seams as a consequence of
increased soil erosion and organic matter burial.[48] Precipitation rates in the North Sea likewise soared
during the PETM.[49] In Cap d'Ailly, in present-day Normandy, a transient dry spell occurred just before
the negative CIE, after which much moister conditions predominated, with the local environment
tranisitioning from a closed marsh to an open, eutrophic swamp with frequent algal blooms.[50] In the
Rocky Mountain Interior, precipitation locally declined, however,[51] as the interior of North America
became more seasonally arid.[52] East African sites display evidence of aridity punctuated by seasonal
episodes of potent precipitation, revealing the global climate during the PETM not to be universally
humid.[53] Evidence from Forada in northeastern Italy suggests that arid and humid climatic intervals
alternated over the course of the PETM concomittantly with precessional cycles in mid-latitudes, and that
overall, net precipitation over the central-western Tethys decreased.[54]

Ocean

The amount of freshwater in the Arctic Ocean increased, in part due to northern hemisphere rainfall
patterns, fueled by poleward storm track migrations under global warming conditions.[45] The flux of
freshwater entering the oceans increased drastically during the PETM, and continued for a time after the
PETM's termination.[55]

Anoxia

In parts of the oceans, especially the north Atlantic Ocean, bioturbation was absent. This may be due to
bottom-water anoxia, or by changing ocean circulation patterns changing the temperatures of the bottom
water.[36] However, many ocean basins remained bioturbated through the PETM.[56] Iodine to calcium
ratios suggest oxygen minimum zones in the oceans expanded vertically and possibly also laterally.[57]
Water column anoxia and euxinia was most prevalent in restricted oceanic basins, such as the Arctic and
Tethys Oceans.[58] Geochemical evidence from the epicontinental North Sea Basin indicates local euxinia
occurred there as well.[59] The Gulf Coastal Plain was also affected by euxinia.[60]

It is possible that during the PETM's early stages, anoxia helped to slow down warming through carbon
drawdown via organic mater burial.[61] A pronounced negative lithium isotope excursion in both marine
carbonates and local weathering inputs suggests that weathering and erosion rates increased during the
PETM, generating an increase in organic carbon burial, which acted as a negative feedback on the PETM's
severe global warming.[62]

Sea level

Along with the global lack of ice, the sea level would have risen due to thermal expansion.[30] Evidence for
this can be found in the shifting palynomorph assemblages of the Arctic Ocean, which reflect a relative
decrease in terrestrial organic material compared to marine organic matter.[30]

Currents

At the start of the PETM, the ocean circulation patterns changed radically in the course of under
5,000  years. Global-scale current directions reversed due to a shift in overturning from the southern
hemisphere to northern hemisphere overturning. This "backwards" flow persisted for 40,000 years. Such a
change would transport warm water to the deep oceans, enhancing further warming.[63] The major biotic
turnover among benthic foraminifera has been cited as evidence of a significant change in deep water
circulation.[64]

Acidification

Ocean acidification occurred during the PETM.[65] The lysocline marks the depth at which carbonate starts
to dissolve (above the lysocline, carbonate is oversaturated): today, this is at about 4 km, comparable to the
median depth of the oceans. This depth depends on (among other things) temperature and the amount of
CO2 dissolved in the ocean. Adding CO2 initially raises the lysocline, resulting in the dissolution of deep
water carbonates. This deep-water acidification can be observed in ocean cores, which show (where
bioturbation has not destroyed the signal) an abrupt change from grey carbonate ooze to red clays (followed
by a gradual grading back to grey). It is far more pronounced in North Atlantic cores than elsewhere,
suggesting that acidification was more concentrated here, related to a greater rise in the level of the
lysocline. Corrosive waters may have then spilled over into other regions of the world ocean from the
North Atlantic. Model simulations show acidic water accumulation in the deep North Atlantic at the onset
of the event. Acidification of deep waters, and the later spreading from the North Atlantic can explain
spatial variations in carbonate dissolution.[66] In parts of the southeast Atlantic, the lysocline rose by 2 km
in just a few thousand years.[56] Evidence from the tropical Pacific Ocean suggests a minimum lysocline
shoaling of around 500 m at the time of this hyperthermal.[67] Acidification may have increased the
efficiency of transport of photic zone water into the ocean depths, thus partially acting as a negative
feedback that retarded the rate of atmospheric carbon dioxide buildup.[68] As a consequence of
coccolithophorid blooms enabled by enhanced runoff, carbonate was removed from seawater as the Earth
recovered from the negative carbon isotope excursion, thus acting to ameliorate ocean acidification.[69]

Life
Stoichiometric magnetite (Fe3 O4 ) particles were obtained from PETM-age marine sediments. The study
from 2008 found elongate prism and spearhead crystal morphologies, considered unlike any magnetite
crystals previously reported, and are potentially of biogenic origin.[70] These biogenic magnetite crystals
show unique gigantism, and probably are of aquatic origin. The study suggests that development of thick
suboxic zones with high iron bioavailability, the result of dramatic changes in weathering and sedimentation
rates, drove diversification of magnetite-forming organisms, likely including eukaryotes.[71] Biogenic
magnetites in animals have a crucial role in geomagnetic field navigation.[72]

Ocean

The PETM is accompanied by a mass extinction of 35–50% of benthic foraminifera (especially in deeper
waters) over the course of ~1,000 years – the group suffering more than during the dinosaur-slaying K-T
extinction.[73][74][75] At the onset of the PETM, benthic foraminiferal diversity dropped by 30% in the
Pacific Ocean,[76] though this decline was not ubiquitous to all sites; Himalayan platform carbonates show
no major change in assemblages of large benthic foraminifera at the onset of the PETM; their decline came
about towards the end of the event.[77] The nannoplankton genus Fasciculithus went extinct as a result of
increased surface water oligotrophy.[78] A decrease in diversity and migration away from the oppressively
hot tropics indicates planktonic foraminifera were adversely affected as well.[79] Contrarily, dinoflagellates
bloomed.

The deep-sea extinctions are difficult to explain, because many species of benthic foraminifera in the deep-
sea are cosmopolitan, and can find refugia against local extinction.[80] General hypotheses such as a
temperature-related reduction in oxygen availability, or increased corrosion due to carbonate undersaturated
deep waters, are insufficient as explanations. Acidification may also have played a role in the extinction of
the calcifying foraminifera, and the higher temperatures would have increased metabolic rates, thus
demanding a higher food supply. Such a higher food supply might not have materialized because warming
and increased ocean stratification might have led to declining productivity [81] and/or increased
remineralization of organic matter in the water column, before it reached the benthic foraminifera on the sea
floor.[82] The only factor global in extent was an increase in temperature. Regional extinctions in the North
Atlantic can be attributed to increased deep-sea anoxia, which could be due to the slowdown of overturning
ocean currents, or the release and rapid oxidation of large amounts of methane.

In shallower waters, it's undeniable that increased CO2 levels result in a decreased oceanic pH, which has a
profound negative effect on corals.[83] Experiments suggest it is also very harmful to calcifying
plankton.[84] However, the strong acids used to simulate the natural increase in acidity which would result
from elevated CO2 concentrations may have given misleading results, and the most recent evidence is that
coccolithophores (E. huxleyi at least) become more, not less, calcified and abundant in acidic waters.[85] No
change in the distribution of calcareous nanoplankton such as the coccolithophores can be attributed to
acidification during the PETM.[85] Extinction rates among calcareous nannoplakton increased, but so did
origination rates.[86] Acidification did lead to an abundance of heavily calcified algae[78] and weakly
calcified forams.[87] The calcareous nannofossil species Neochiastozygus junctus thrived; its success is
attributable to enhanced surficial productivity caused by enhanced nutrient runoff.[88]

A study published in May 2021 concluded that fish thrived in at least some tropical areas during the PETM,
based on discovered fish fossils including Mene maculata at Ras Gharib, Egypt.[89]

Land
Humid conditions caused migration of modern Asian mammals northward, dependent on the climatic belts.
Uncertainty remains for the timing and tempo of migration.[32]

The increase in mammalian abundance is intriguing. Increased global temperatures may have promoted
dwarfing[90][91][92] – which may have encouraged speciation. Major dwarfing occurred early in the
PETM, with further dwarfing taking place during the middle of the hyperthermal.[9] The dwarfing of
various mammal lineages led to further dwarfing in other mammals whose reduction in body size was not
directly induced by the PETM.[93] Many major mammalian clades – including hyaenodontids, artiodactyls,
perissodactyls, and primates – appeared and spread around the globe 13,000 to 22,000 years after the
initiation of the PETM.[94][90]

The diversity of insect herbivory, as measured by the amount and diversity of damage to plants caused by
insects, increased during the PETM in correlation with global warming.[95] The ant genus Gesomyrmex
radiated across Eurasia during the PETM.[96] As with mammals, soil-dwelling invertebrates are observed to
have dwarfed during the PETM.[97]

A profound change in terrestrial vegetation across the globe is associated with the PETM. Across all
regions, floras from the latest Palaeocene are highly distinct from those of the PETM and the Early
Eocene.[98]

Geologic effects

Sediment deposition changed significantly at many outcrops and in many drill cores spanning this time
interval.[99] During the PETM, sediments are enriched with kaolinite from a detrital source due to
denudation (initial processes such as volcanoes, earthquakes, and plate tectonics). Increased precipitation
and enhanced erosion of older kaolinite-rich soils and sediments may have been responsible for this.
Increased weathering from the enhanced runoff formed thick paleosoil enriched with carbonate nodules
(Microcodium like), and this suggests a semi-arid climate.[32] Unlike during lesser, more gradual
hyperthermals, glauconite authigenesis was inhibited.[100]

In the Tremp-Graus Basin of northern Spain, fluvial systems grew and rates of deposition of alluvial
sediments increased with a lag time of around 3,800 years after the PETM.[101]

At some marine locations (mostly deep-marine), sedimentation rates must have decreased across the PETM,
presumably because of carbonate dissolution on the seafloor; at other locations (mostly shallow-marine),
sedimentation rates must have increased across the PETM, presumably because of enhanced delivery of
riverine material during the event.[102]

Possible causes
Discriminating between different possible causes of the PETM is difficult. Temperatures were rising
globally at a steady pace, and a mechanism must be invoked to produce an instantaneous spike which may
have been accentuated or catalyzed by positive feedback (or activation of "tipping or points"[103]). The
biggest aid in disentangling these factors comes from a consideration of the carbon isotope mass balance.
We know the entire exogenic carbon cycle (i.e. the carbon contained within the oceans and atmosphere,
which can change on short timescales) underwent a −0.2  % to −0.3  % perturbation in δ13 C, and by
considering the isotopic signatures of other carbon reserves, can consider what mass of the reserve would
be necessary to produce this effect. The assumption underpinning this approach is that the mass of exogenic
carbon was the same in the Paleogene as it is today – something which is very difficult to confirm.

Eruption of large kimberlite field

Although the cause of the initial warming has been attributed to a massive injection of carbon (CO2 and/or
CH4 ) into the atmosphere, the source of the carbon has yet to be found. The emplacement of a large cluster
of kimberlite pipes at ~56 Ma in the Lac de Gras region of northern Canada may have provided the carbon
that triggered early warming in the form of exsolved magmatic CO2 . Calculations indicate that the
estimated 900–1,100 Pg[104] of carbon required for the initial approximately 3 °C of ocean water warming
associated with the Paleocene-Eocene thermal maximum could have been released during the emplacement
of a large kimberlite cluster.[105] The transfer of warm surface ocean water to intermediate depths led to
thermal dissociation of seafloor methane hydrates, providing the isotopically depleted carbon that produced
the carbon isotopic excursion. The coeval ages of two other kimberlite clusters in the Lac de Gras field and
two other early Cenozoic hyperthermals indicate that CO2 degassing during kimberlite emplacement is a
plausible source of the CO2 responsible for these sudden global warming events.

Volcanic activity

To balance the mass of carbon and produce the observed


δ13 C value, at least 1,500 gigatons of carbon would have to
degas from the mantle via volcanoes over the course of the
two, 1,000 year, steps. To put this in perspective, this is about
200 times the background rate of degassing for the rest of the
Paleocene. There is no indication that such a burst of
volcanic activity has occurred at any point in Earth's history.
However, substantial volcanism had been active in East
Greenland for around the preceding million years or so, but
this struggles to explain the rapidity of the PETM. Even if
Satellite photo of Ardnamurchan – with
the bulk of the 1,500 gigatons of carbon was released in a
clearly visible circular shape, which is the
single pulse, further feedbacks would be necessary to
'plumbings of an ancient volcano'
produce the observed isotopic excursion.

On the other hand, there are suggestions that surges of


activity occurred in the later stages of the volcanism and associated continental rifting. Intrusions of hot
magma into carbon-rich sediments may have triggered the degassing of isotopically light methane in
sufficient volumes to cause global warming and the observed isotope anomaly. This hypothesis is
documented by the presence of extensive intrusive sill complexes and thousands of kilometer-sized
hydrothermal vent complexes in sedimentary basins on the mid-Norwegian margin and west of
Shetland.[106][107] Volcanic eruptions of a large magnitude can impact global climate, reducing the amount
of solar radiation reaching the Earth's surface, lowering temperatures in the troposphere, and changing
atmospheric circulation patterns. Large-scale volcanic activity may last only a few days, but the massive
outpouring of gases and ash can influence climate patterns for years. Sulfuric gases convert to sulfate
aerosols, sub-micron droplets containing about 75 percent sulfuric acid. Following eruptions, these aerosol
particles can linger as long as three to four years in the stratosphere.[108] Further phases of volcanic activity
could have triggered the release of more methane, and caused other early Eocene warm events such as the
ETM2.[36] It has also been suggested that volcanic activity around the Caribbean may have disrupted the
circulation of oceanic currents,[109] amplifying the magnitude of climate change.
Mercury isotope shifts show intense volcanism was concurrent with the beginning of the PETM.[110]
Osmium isotopic anomalies in Arctic Ocean sediments dating to the PETM have been interpreted as
evidence of a volcanic cause of this hyperthermal.[111]

A 2017 study noted strong evidence of a volcanic carbon source (greater than 10,000 gigatons of carbon),
associated with the North Atlantic Igneous Province.[5] A 2021 study found the PETM was directly
preceded by volcanism.[103]

Orbital forcing

The presence of later (smaller) warming events of a global scale, such as the Elmo horizon (aka ETM2),
has led to the hypothesis that the events repeat on a regular basis, driven by maxima in the 400,000 and
100,000 year eccentricity cycles in the Earth's orbit.[112] Cores from Howard's Tract, Maryland indicate the
PETM occurred as a result of an extreme in axial precession during an orbital eccentricity maximum.[113]
The current warming period is expected to last another 50,000 years due to a minimum in the eccentricity
of the Earth's orbit. Orbital increase in insolation (and thus temperature) would force the system over a
threshold and unleash positive feedbacks.[114] The orbital forcing hypothesis has been challenged by a
study finding the PETM to have coincided with a minimum in the ∼400 kyr eccentricity cycle, inconsistent
with a proposed orbital trigger for the hyperthermal.[115]

Comet impact

One theory holds that a 12 C-rich comet struck the earth and initiated the warming event. A cometary impact
coincident with the P/E boundary can also help explain some enigmatic features associated with this event,
such as the iridium anomaly at Zumaia, the abrupt appearance of a localized kaolinitic clay layer with
abundant magnetic nanoparticles, and especially the nearly simultaneous onset of the carbon isotope
excursion and the thermal maximum.

A key feature and testable prediction of a comet impact is that it should produce virtually instantaneous
environmental effects in the atmosphere and surface ocean with later repercussions in the deeper
ocean.[116] Even allowing for feedback processes, this would require at least 100 gigatons of extraterrestrial
carbon.[116] Such a catastrophic impact should have left its mark on the globe. A clay layer of 5-20m
thickness on the coastal shelf of New Jersey contained unusual amounts of magnetite, but it was found to
have formed 9-18 kyr too late for these magnetic particles to have been a result of a comet's impact, and the
particles had a crystal structure which was a signature of magnetotactic bacteria rather than an
extraterrestrial origin.[117] However, recent analyses have shown that isolated particles of non-biogenic
origin make up the majority of the magnetic particles in the clay sample.[118]

A 2016 report in Science describes the discovery of impact ejecta from three marine P-E boundary sections
from the Atlantic margin of the eastern U.S., indicating that an extraterrestrial impact occurred during the
carbon isotope excursion at the P-E boundary.[119][120] The silicate glass spherules found were identified
as microtektites and microkrystites.[119]

Burning of peat

The combustion of prodigious quantities of peat was once postulated, because there was probably a greater
mass of carbon stored as living terrestrial biomass during the Paleocene than there is today since plants in
fact grew more vigorously during the period of the PETM. This theory was refuted, because in order to
produce the δ13 C excursion observed, over 90 percent of the Earth's biomass would have to have been
combusted. However, the Paleocene is also recognized as a time of significant peat accumulation
worldwide. A comprehensive search failed to find evidence for the combustion of fossil organic matter, in
the form of soot or similar particulate carbon.[121]

Enhanced respiration

Respiration rates of organic matter increase when temperatures rise. One feedback mechanism proposed to
explain the rapid rise in carbon dioxide levels is a sudden, speedy rise in terrestrial respiration rates
concordant with global temperature rise initiated by any of the other causes of warming.[122]

Methane clathrate release

Methane hydrate dissolution has been invoked as a highly plausible causal mechanism for the carbon
isotope excursion and warming observed at the PETM.[123] The most obvious feedback mechanism that
could amplify the initial perturbation is that of methane clathrates. Under certain temperature and pressure
conditions, methane – which is being produced continually by decomposing microbes in sea bottom
sediments – is stable in a complex with water, which forms ice-like cages trapping the methane in solid
form. As temperature rises, the pressure required to keep this clathrate configuration stable increases, so
shallow clathrates dissociate, releasing methane gas to make its way into the atmosphere. Since biogenic
clathrates have a δ13 C signature of −60 ‰ (inorganic clathrates are the still rather large −40 ‰), relatively
small masses can produce large δ13 C excursions. Further, methane is a potent greenhouse gas as it is
released into the atmosphere, so it causes warming, and as the ocean transports this warmth to the bottom
sediments, it destabilizes more clathrates.[33]

In order for the clathrate hypothesis to be applicable to PETM, the oceans must show signs of having been
warmer slightly before the carbon isotope excursion, because it would take some time for the methane to
become mixed into the system and δ13 C-reduced carbon to be returned to the deep ocean sedimentary
record. Up until the 2000s, the evidence suggested that the two peaks were in fact simultaneous, weakening
the support for the methane theory. In 2002, a short gap between the initial warming and the δ13 C
excursion was detected.[124] In 2007, chemical markers of surface temperature (TEX86 ) had also indicated
that warming occurred around 3,000 years before the carbon isotope excursion, although this did not seem
to hold true for all cores.[42] However, research in 2005 found no evidence of this time gap in the deeper
(non-surface) waters.[125] Moreover, the small apparent change in TEX86 that precede the δ13 C anomaly
can easily (and more plausibly) be ascribed to local variability (especially on the Atlantic coastal plain, e.g.
Sluijs, et al., 2007) as the TEX86 paleo-thermometer is prone to significant biological effects. The δ18 O of
benthic or planktonic forams does not show any pre-warming in any of these localities, and in an ice-free
world, it is generally a much more reliable indicator of past ocean temperatures. Analysis of these records
reveals another interesting fact: planktonic (floating) forams record the shift to lighter isotope values earlier
than benthic (bottom dwelling) forams.[126] The lighter (lower δ13 C) methanogenic carbon can only be
incorporated into the forams' shells after it has been oxidised. A gradual release of the gas would allow it to
be oxidised in the deep ocean, which would make benthic forams show lighter values earlier. The fact that
the planktonic forams are the first to show the signal suggests that the methane was released so rapidly that
its oxidation used up all the oxygen at depth in the water column, allowing some methane to reach the
atmosphere unoxidised, where atmospheric oxygen would react with it. This observation also allows us to
constrain the duration of methane release to under around 10,000 years.[124]
However, there are several major problems with the methane hydrate dissociation hypothesis. The most
parsimonious interpretation for surface-water forams to show the δ13 C excursion before their benthic
counterparts (as in the Thomas et al. paper) is that the perturbation occurred from the top down, and not the
bottom up. If the anomalous δ13 C (in whatever form: CH4 or CO2 ) entered the atmospheric carbon
reservoir first, and then diffused into the surface ocean waters, which mix with the deeper ocean waters
over much longer time-scales, we would expect to observe the planktonics shifting toward lighter values
before the benthics. Moreover, careful examination of the Thomas et al. data set shows that there is not a
single intermediate planktonic foram value, implying that the perturbation and attendant δ13 C anomaly
happened over the lifespan of a single foram – much too fast for the nominal 10,000-year release needed
for the methane hypothesis to work.

An additional critique of the methane clathrate release hypothesis is that the warming effects of large-scale
methane release would not be sustainable for more than a millennium. Thus, exponents of this line of
criticism suggest that methane clathrate release could not have been the main driver of the PETM, which
lasted for 50,000 to 200,000 years.[127]

There has been some debate about whether there was a large enough amount of methane hydrate to be a
major carbon source; a 2011 paper proposed that was the case.[128] The present-day global methane
hydrate reserve was once considered to be between 2,000 and 10,000 Gt C (billions of tons of carbon), but
is now estimated between 1500 and 2000 Gt C.[129] However, because the global ocean bottom
temperatures were ~6 °C higher than today, which implies a much smaller volume of sediment hosting gas
hydrate than today, the global amount of hydrate before the PETM has been thought to be much less than
present-day estimates.[127] In a 2006 study, scientists regarded the source of carbon for the PETM to be a
mystery.[130] A 2011 study, using numerical simulations suggests that enhanced organic carbon
sedimentation and methanogenesis could have compensated for the smaller volume of hydrate stability.[128]
A 2016 study based on reconstructions of atmospheric CO2 content during the PETM's carbon isotope
excursions (CIE), using triple oxygen isotope analysis, suggests a massive release of seabed methane into
the atmosphere as the driver of climatic changes. The authors also state that a massive release of methane
hydrates through thermal dissociation of methane hydrate deposits has been the most convincing hypothesis
for explaining the CIE ever since it was first identified, according to them.[131] In 2019, a study suggested
that there was a global warming of around 2 degrees several millennia before PETM, and that this warming
had eventually destabilized methane hydrates and caused the increased carbon emission during PETM, as
evidenced by the large increase in barium ocean concentrations (since PETM-era hydrate deposits would
have been also been rich in barium, and would have released it upon their meltdown).[132] In 2022, a
foraminiferal records study had reinforced this conclusion, suggesting that the release of CO2 before PETM
was comparable to the current anthropogenic emissions in its rate and scope, to the point that that there was
enough time for a recovery to background levels of warming and ocean acidification in the centuries to
millennia between the so-called pre-onset excursion (POE) and the main event (carbon isotope excursion,
or CIE).[103] A 2021 paper had further indicated that while PETM began with a significant intensification
of volcanic activity and that lower-intensity volcanic activity sustained elevated carbon dioxide levels, "at
least one other carbon reservoir released significant greenhouse gases in response to initial warming".[133]

It was estimated in 2001 that it would take around 2,300 years for an increased temperature to diffuse
warmth into the sea bed to a depth sufficient to cause a release of clathrates, although the exact time-frame
is highly dependent on a number of poorly constrained assumptions.[134] Ocean warming due to flooding
and pressure changes due to a sea-level drop may have caused clathrates to become unstable and release
methane. This can take place over as short of a period as a few thousand years. The reverse process, that of
fixing methane in clathrates, occurs over a larger scale of tens of thousands of years.[135]

Ocean circulation

The large scale patterns of ocean circulation are important when considering how heat was transported
through the oceans. Our understanding of these patterns is still in a preliminary stage. Models show that
there are possible mechanisms to quickly transport heat to the shallow, clathrate-containing ocean shelves,
given the right bathymetric profile, but the models cannot yet match the distribution of data we observe.
"Warming accompanying a south-to-north switch in deepwater formation would produce sufficient
warming to destabilize seafloor gas hydrates over most of the world ocean to a water depth of at least 1900
m." This destabilization could have resulted in the release of more than 2000 gigatons of methane gas from
the clathrate zone of the ocean floor.[136] The timing of changes in ocean circulation with respect to the
shift in carbon isotope ratios has been argued to support the proposition that warmer deep water caused
methane hydrate release.[137] However, a different study found no evidence of a change in deep water
formation, instead suggesting that deepened subtropical subduction rather than subtropical deep water
formation occurred during the PETM.[138]

Arctic freshwater input into the North Pacific could serve as a catalyst for methane hydrate destabilization,
an event suggested as a precursor to the onset of the PETM.[139]

Recovery
Climate proxies, such as ocean sediments (depositional rates) indicate a duration of ∼83 ka, with ∼33 ka in
the early rapid phase and ∼50 ka in a subsequent gradual phase.[2]

The most likely method of recovery involves an increase in biological productivity, transporting carbon to
the deep ocean. This would be assisted by higher global temperatures and CO2 levels, as well as an
increased nutrient supply (which would result from higher continental weathering due to higher
temperatures and rainfall; volcanoes may have provided further nutrients). Evidence for higher biological
productivity comes in the form of bio-concentrated barium.[140] However, this proxy may instead reflect
the addition of barium dissolved in methane.[141] Diversifications suggest that productivity increased in
near-shore environments, which would have been warm and fertilized by run-off, outweighing the
reduction in productivity in the deep oceans.[87] Another pulse of NAIP volcanic activity may have also
played a role in terminating the hyperthermal via a volcanic winter.[142]

Comparison with today's climate change


Since at least 1997, the Paleocene–Eocene thermal maximum has been investigated in geoscience as an
analog to understand the effects of global warming and of massive carbon inputs to the ocean and
atmosphere,[143] including ocean acidification.[33] Humans today emit about 10 Gt of carbon (about 37 Gt
CO2e) per year, and will have released a comparable amount in about 1,000 years at that rate. A main
difference is that during the Paleocene–Eocene thermal maximum, the planet was ice-free, as the Drake
Passage had not yet opened and the Central American Seaway had not yet closed.[144] Although the PETM
is now commonly held to be a "case study" for global warming and massive carbon emission,[1][2][34] the
cause, details, and overall significance of the event remain uncertain.
Model simulations of peak carbon addition to the ocean–atmosphere system during the PETM give a
probable range of 0.3–1.7 petagrams of carbon per year (Pg C/yr), which is much slower than the currently
observed rate of carbon emissions. One petagram of carbon is equivalent to a gigaton of carbon (GtC); the
current rate of carbon injection into the atmosphere is over 10 GtC/yr, a rate much greater than the carbon
injection rate that occurred during the PETM.[145] It has been suggested that today's methane emission
regime from the ocean floor is potentially similar to that during the PETM.[146] Because the modern rate of
carbon release exceeds the PETM's, it is speculated the a PETM-like scenario is the best-case consequence
of anthropogenic global warming, with a mass extinction of a magnitude similar to the Cretaceous-
Palaeogene extinction event being a worst-case scenario.[147]

Professor of Earth and planetary sciences James Zachos notes that IPCC projections for 2300 in the
'business-as-usual' scenario could "potentially bring global temperature to a level the planet has not seen in
50 million years" – during the early Eocene.[148] Some have described the PETM as arguably the best
ancient analog of modern climate change.[149] Scientists have investigated effects of climate change on
chemistry of the oceans by exploring oceanic changes during the PETM.[150][151]

A study found that the PETM shows that substantial climate-shifting tipping points in the Earth system
exist, which "can trigger release of additional carbon reservoirs and drive Earth's climate into a hotter
state".[152][103]

Whether climate sensitivity was lower or higher during the PETM than today remains under debate. A 2022
study found that the Eurasian Epicontinental Sea acted as a major carbon sink during the PETM due to its
high biological productivity and helped to slow and mitigate the warming, and that the existence of many
large epicontinental seas at that time made the Earth's climate less sensitive to forcing by greenhouse gases
relative to today, when much fewer epicontinental seas exist.[153] Other research, however, suggests that
climate sensitivity was higher during the PETM than today, meaning that sensitivity to greenhouse gas
release increases the higher their concentration in the atmosphere.[154]

See also
Abrupt climate change Eocene
Azolla event Eocene Thermal Maximum 2
Canfield ocean Paleocene
Clathrate gun hypothesis Paleogene
Climate sensitivity Runaway climate change

References
1. Haynes, Laura L.; Hönisch, Bärbel (14 September 2020). "The seawater carbon inventory at
the Paleocene–Eocene Thermal Maximum" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC
7533689). Proceedings of the National Academy of Sciences of the United States of
America. 117 (39): 24088–24095. Bibcode:2020PNAS..11724088H (https://ui.adsabs.harvar
d.edu/abs/2020PNAS..11724088H). doi:10.1073/pnas.2003197117 (https://doi.org/10.107
3%2Fpnas.2003197117). PMC 7533689 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC75
33689). PMID 32929018 (https://pubmed.ncbi.nlm.nih.gov/32929018).
2. McInherney, F.A.; Wing, S. (2011). "A perturbation of carbon cycle, climate, and biosphere
with implications for the future" (http://www.whoi.edu/fileserver.do?id=136084&pt=2&p=1487
09). Annual Review of Earth and Planetary Sciences. 39: 489–516.
Bibcode:2011AREPS..39..489M (https://ui.adsabs.harvard.edu/abs/2011AREPS..39..489M).
doi:10.1146/annurev-earth-040610-133431 (https://doi.org/10.1146%2Fannurev-earth-0406
10-133431). Archived (https://web.archive.org/web/20160914003526/http://www.whoi.edu/fil
eserver.do?id=136084&pt=2&p=148709) from the original on 2016-09-14. Retrieved
2016-02-03.
3. Westerhold, T..; Röhl, U.; Raffi, I.; Fornaciari, E.; Monechi, S.; Reale, V.; Bowles, J.; Evans,
H. F. (2008). "Astronomical calibration of the Paleocene time" (https://www.geo.arizona.edu/
~reiners/fortransfer6/WesterholdEtAl_PPP2008.pdf) (PDF). Palaeogeography,
Palaeoclimatology, Palaeoecology. 257 (4): 377–403. Bibcode:2008PPP...257..377W (http
s://ui.adsabs.harvard.edu/abs/2008PPP...257..377W). doi:10.1016/j.palaeo.2007.09.016 (htt
ps://doi.org/10.1016%2Fj.palaeo.2007.09.016). Archived (https://web.archive.org/web/20170
809094938/http://www.geo.arizona.edu/~reiners/fortransfer6/WesterholdEtAl_PPP2008.pdf)
(PDF) from the original on 2017-08-09. Retrieved 2019-07-06.
4. Bowen; et al. (2015). "Two massive, rapid releases of carbon during the onset of the
Palaeocene–Eocene thermal maximum". Nature. 8 (1): 44–47. Bibcode:2015NatGe...8...44B
(https://ui.adsabs.harvard.edu/abs/2015NatGe...8...44B). doi:10.1038/ngeo2316 (https://doi.o
rg/10.1038%2Fngeo2316).
5. Gutjahr, Marcus; Ridgwell, Andy; Sexton, Philip F.; Anagnostou, Eleni; Pearson, Paul N.;
Pälike, Heiko; Norris, Richard D.; Thomas, Ellen; Foster, Gavin L. (August 2017). "Very large
release of mostly volcanic carbon during the Palaeocene–Eocene Thermal Maximum" (http
s://www.ncbi.nlm.nih.gov/pmc/articles/PMC5582631). Nature. 548 (7669): 573–577.
Bibcode:2017Natur.548..573G (https://ui.adsabs.harvard.edu/abs/2017Natur.548..573G).
doi:10.1038/nature23646 (https://doi.org/10.1038%2Fnature23646). ISSN 1476-4687 (http
s://www.worldcat.org/issn/1476-4687). PMC 5582631 (https://www.ncbi.nlm.nih.gov/pmc/arti
cles/PMC5582631). PMID 28858305 (https://pubmed.ncbi.nlm.nih.gov/28858305).
6. Jones, S.M.; Hoggett, M.; Greene, S.E.; Jones, T.D. (2019). "Large Igneous Province
thermogenic greenhouse gas flux could have initiated Paleocene-Eocene Thermal
Maximum climate change" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6895149). Nature
Communications. 10 (1): 5547. Bibcode:2019NatCo..10.5547J (https://ui.adsabs.harvard.ed
u/abs/2019NatCo..10.5547J). doi:10.1038/s41467-019-12957-1 (https://doi.org/10.1038%2F
s41467-019-12957-1). PMC 6895149 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC68951
49). PMID 31804460 (https://pubmed.ncbi.nlm.nih.gov/31804460).
7. Kennett, J.P.; Stott, L.D. (1991). "Abrupt deep-sea warming, palaeoceanographic changes
and benthic extinctions at the end of the Paleocene" (http://shadow.eas.gatech.edu/~kcobb/
warming_papers/kennett&stott91.pdf) (PDF). Nature. 353 (6341): 225–229.
Bibcode:1991Natur.353..225K (https://ui.adsabs.harvard.edu/abs/1991Natur.353..225K).
doi:10.1038/353225a0 (https://doi.org/10.1038%2F353225a0). S2CID 35071922 (https://api.
semanticscholar.org/CorpusID:35071922). Archived (https://web.archive.org/web/20160303
201412/http://shadow.eas.gatech.edu/~kcobb/warming_papers/kennett%26stott91.pdf)
(PDF) from the original on 2016-03-03. Retrieved 2020-01-08.
8. Koch, P.L.; Zachos, J.C.; Gingerich, P.D. (1992). "Correlation between isotope records in
marine and continental carbon reservoirs near the Palaeocene/Eocene boundary". Nature.
358 (6384): 319–322. Bibcode:1992Natur.358..319K (https://ui.adsabs.harvard.edu/abs/199
2Natur.358..319K). doi:10.1038/358319a0 (https://doi.org/10.1038%2F358319a0).
hdl:2027.42/62634 (https://hdl.handle.net/2027.42%2F62634). S2CID 4268991 (https://api.s
emanticscholar.org/CorpusID:4268991).
9. Van der Meulen, Bas; Gingerich, Philip D.; Lourens, Lucas J.; Meijer, Niels; Van
Broekhuizen, Sjors; Van Ginneken, Sverre; Abels, Hemmo A. (15 March 2020). "Carbon
isotope and mammal recovery from extreme greenhouse warming at the Paleocene–Eocene
boundary in astronomically-calibrated fluvial strata, Bighorn Basin, Wyoming, USA" (https://d
oi.org/10.1016%2Fj.epsl.2019.116044). Earth and Planetary Science Letters. 534: 116044.
Bibcode:2020E&PSL.53416044V (https://ui.adsabs.harvard.edu/abs/2020E&PSL.53416044
V). doi:10.1016/j.epsl.2019.116044 (https://doi.org/10.1016%2Fj.epsl.2019.116044).
S2CID 212852180 (https://api.semanticscholar.org/CorpusID:212852180).
10. Zachos, J. C.; Kump, L. R. (2005). "Carbon cycle feedbacks and the initiation of Antarctic
glaciation in the earliest Oligocene". Global and Planetary Change. 47 (1): 51–66.
Bibcode:2005GPC....47...51Z (https://ui.adsabs.harvard.edu/abs/2005GPC....47...51Z).
doi:10.1016/j.gloplacha.2005.01.001 (https://doi.org/10.1016%2Fj.gloplacha.2005.01.001).
11. "International Chronostratigraphic Chart" (https://stratigraphy.org/icschart/ChronostratChart2
020-01.pdf) (PDF). International Commission on Stratigraphy.
12. Zachos, J.C.; Dickens, G.R.; Zeebe, R.E. (2008). "An early Cenozoic perspective on
greenhouse warming and carbon-cycle dynamics" (http://es.ucsc.edu/%7Ejzachos/pubs/Zac
hos_Dickens_Zeebe_08.pdf) (PDF). Nature. 451 (7176): 279–283.
Bibcode:2008Natur.451..279Z (https://ui.adsabs.harvard.edu/abs/2008Natur.451..279Z).
doi:10.1038/nature06588 (https://doi.org/10.1038%2Fnature06588). PMID 18202643 (https://
pubmed.ncbi.nlm.nih.gov/18202643). S2CID 4360841 (https://api.semanticscholar.org/Corp
usID:4360841). Archived (https://web.archive.org/web/20080705030100/http://es.ucsc.ed
u/%7Ejzachos/pubs/Zachos_Dickens_Zeebe_08.pdf) (PDF) from the original on 2008-07-
05. Retrieved 2008-04-23.
13. Gilmour, Iain; Jolley, David; Kemp, David; Kelley, Simon; Gilmour, Mabs; Daly, Rob;
Widdowson, Mike (1 September 2014). "The early Danian hyperthermal event at Boltysh
(Ukraine): Relation to Cretaceous-Paleogene boundary events" (https://pubs.geosciencewor
ld.org/books/book/674/chapter-abstract/3807802/The-early-Danian-hyperthermal-event-at-B
oltysh?redirectedFrom=fulltext). In Keller, Gerta; Kerr, Andrew C. (eds.). Volcanism, Impacts,
and Mass Extinctions: Causes and Effects (https://pubs.geoscienceworld.org/gsa/books/boo
k/674/Volcanism-Impacts-and-Mass-Extinctions-Causes-and). Geological Society of
America. doi:10.1130/2014.2505(06) (https://doi.org/10.1130%2F2014.2505%2806%29).
ISBN 9780813725055.
14. Thomas, Ellen; Boscolo-Galazzo, Flavia; Balestra, Barbara; Monechi, Simonetta; Donner,
Barbara; Röhl, Ursula (1 July 2018). "Early Eocene Thermal Maximum 3: Biotic Response at
Walvis Ridge (SE Atlantic Ocean)" (https://agupubs.onlinelibrary.wiley.com/doi/full/10.1029/2
018PA003375). Paleoceanography and Paleoclimatology. 33 (8): 862–883.
Bibcode:2018PaPa...33..862T (https://ui.adsabs.harvard.edu/abs/2018PaPa...33..862T).
doi:10.1029/2018PA003375 (https://doi.org/10.1029%2F2018PA003375).
S2CID 133958051 (https://api.semanticscholar.org/CorpusID:133958051). Retrieved
22 November 2022.
15. Inglis, Gordon; et al. (2020). "Global mean surface temperature and climate sensitivity of the
early Eocene Climatic Optimum (EECO), Paleocene–Eocene Thermal Maximum (PETM),
and latest Paleocene" (https://doi.org/10.5194/cp-16-1953-2020). Climate of the Past. 16 (5):
1953–1968. Bibcode:2020CliPa..16.1953I (https://ui.adsabs.harvard.edu/abs/2020CliPa..16.
1953I). doi:10.5194/cp-16-1953-2020 (https://doi.org/10.5194%2Fcp-16-1953-2020).
S2CID 227178097 (https://api.semanticscholar.org/CorpusID:227178097).
16. Evans, David; Wade, Bridget S.; Henehan, Michael; Erez, Jonathan; Müller, Wolfgang (6
April 2016). "Revisiting carbonate chemistry controls on planktic foraminifera Mg / Ca:
implications for sea surface temperature and hydrology shifts over the Paleocene–Eocene
Thermal Maximum and Eocene–Oligocene transition" (https://cp.copernicus.org/articles/12/8
19/2016/). Climate of the Past. 12 (4): 819–835. Bibcode:2016CliPa..12..819E (https://ui.ads
abs.harvard.edu/abs/2016CliPa..12..819E). doi:10.5194/cp-12-819-2016 (https://doi.org/10.5
194%2Fcp-12-819-2016). Retrieved 5 April 2023.
17. Thomas, Ellen; Shackleton, Nicholas J. (1996). "The Paleocene-Eocene benthic
foraminiferal extinction and stable isotope anomalies" (http://wesscholar.wesleyan.edu/cgi/vi
ewcontent.cgi?article=1133&context=div3facpubs). Geological Society of London, Special
Publications. 101 (1): 401–441. Bibcode:1996GSLSP.101..401T (https://ui.adsabs.harvard.e
du/abs/1996GSLSP.101..401T). doi:10.1144/GSL.SP.1996.101.01.20 (https://doi.org/10.114
4%2FGSL.SP.1996.101.01.20). S2CID 130770597 (https://api.semanticscholar.org/CorpusI
D:130770597). Archived (https://web.archive.org/web/20130521003932/http://wesscholar.we
sleyan.edu/cgi/viewcontent.cgi?article=1133&context=div3facpubs) from the original on
2013-05-21. Retrieved 2013-04-21.
18. Speijer, Robert; Scheibner, Christian; Stassen, Peter; Morsi, Abdel-Mohsen M. (1 May 2012).
"Response of marine ecosystems to deep-time global warming: a synthesis of biotic patterns
across the Paleocene-Eocene thermal maximum (PETM)" (https://lirias.kuleuven.be/27866
6?limo=0). Austrian Journal of Earth Sciences. 105 (1): 6–16. Retrieved 6 April 2023.
19. Wing, Scott L.; Harrington, Guy J.; Smith, Francesca A.; Bloch, Jonathan I.; Boyer, Douglas
M.; Freeman, Katherine H. (11 November 2005). "Transient Floral Change and Rapid Global
Warming at the Paleocene-Eocene Boundary" (https://www.science.org/doi/abs/10.1126/sci
ence.1116913). Science. 310 (5750): 993–996. Bibcode:2005Sci...310..993W (https://ui.ads
abs.harvard.edu/abs/2005Sci...310..993W). doi:10.1126/science.1116913 (https://doi.org/10.
1126%2Fscience.1116913). PMID 16284173 (https://pubmed.ncbi.nlm.nih.gov/16284173).
S2CID 7069772 (https://api.semanticscholar.org/CorpusID:7069772). Retrieved 6 April
2023.
20. Frieling, Joost; Gebhardt, Holger; Huber, Matthew; Adekeye, Olabisi A.; Akande, Samuel O.;
Reichart, Gert-Jan; Middelburg, Jack J.; Schouten, Stefan; Sluijs, Appy (3 March 2017).
"Extreme warmth and heat-stressed plankton in the tropics during the Paleocene-Eocene
Thermal Maximum" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5336354). Science
Advances. 3 (3): e1600891. Bibcode:2017SciA....3E0891F (https://ui.adsabs.harvard.edu/ab
s/2017SciA....3E0891F). doi:10.1126/sciadv.1600891 (https://doi.org/10.1126%2Fsciadv.16
00891). PMC 5336354 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5336354).
PMID 28275727 (https://pubmed.ncbi.nlm.nih.gov/28275727).
21. Aze, T.; Pearson, P. N.; Dickson, A. J.; Badger, M. P. S.; Bown, P. R.; Pancost, Richard D.;
Gibbs, S. J.; Huber, Brian T.; Leng, M. J.; Coe, A. L.; Cohen, A. S.; Foster, G. L. (1 September
2014). "Extreme warming of tropical waters during the Paleocene–Eocene Thermal
Maximum" (https://pubs.geoscienceworld.org/gsa/geology/article/42/9/739/131603/Extreme-
warming-of-tropical-waters-during-the). Geology. 42 (9): 739–742.
Bibcode:2014Geo....42..739A (https://ui.adsabs.harvard.edu/abs/2014Geo....42..739A).
doi:10.1130/G35637.1 (https://doi.org/10.1130%2FG35637.1). S2CID 216051165 (https://ap
i.semanticscholar.org/CorpusID:216051165). Retrieved 6 April 2023.
22. Barnet, James S. K.; Harper, Dustin T.; LeVay, Leah J.; Edgar, Kirsty M.; Henehan, Michael
J.; Babila, Tali L.; Ullmann, Clemens V.; Leng, Melanie J.; Kroon, Dick; Zachos, James C.;
Littler, Kate (1 September 2020). "Coupled evolution of temperature and carbonate
chemistry during the Paleocene–Eocene; new trace element records from the low latitude
Indian Ocean" (https://www.sciencedirect.com/science/article/pii/S0012821X20303587).
Earth and Planetary Science Letters. 545: 1–15. doi:10.1016/j.epsl.2020.116414 (https://doi.
org/10.1016%2Fj.epsl.2020.116414). Retrieved 3 July 2023.
23. Zachos, James C.; Wara, Michael W.; Bohaty, Steven; Delaney, Margaret L.; Petrizzo, Maria
Rose; Brill, Amanda; Bralower, Timothy J.; Premoli-Silva, Isabella (28 November 2003). "A
Transient Rise in Tropical Sea Surface Temperature During the Paleocene-Eocene Thermal
Maximum" (https://www.science.org/doi/10.1126/science.1090110). Science. 302 (5650):
1551–1554. Bibcode:2003Sci...302.1551Z (https://ui.adsabs.harvard.edu/abs/2003Sci...302.
1551Z). doi:10.1126/science.1090110 (https://doi.org/10.1126%2Fscience.1090110).
PMID 14576441 (https://pubmed.ncbi.nlm.nih.gov/14576441). S2CID 6582869 (https://api.se
manticscholar.org/CorpusID:6582869). Retrieved 6 April 2023.
24. Hollis, Christopher J.; Taylor, Kyle W. R.; Handley, Luke; Pancost, Richard D.; Huber,
Matthew; Creech, John B.; Hines, Benjamin R.; Crouch, Erica M.; Morgans, Hugh E. G.;
Crampton, James S.; Gibbs, Samantha; Pearson, Paul N.; Zachos, James C. (15 July 2013).
"Erratum to "Early Paleogene temperature history of the Southwest Pacific Ocean:
Reconciling proxies and models" [Earth Planet. Sci. Lett. 349 (2012) 53–66]" (https://www.sc
iencedirect.com/science/article/pii/S0012821X13003282). Earth and Planetary Science
Letters. 374: 258–259. Bibcode:2013E&PSL.374..258H (https://ui.adsabs.harvard.edu/abs/2
013E&PSL.374..258H). doi:10.1016/j.epsl.2013.06.012 (https://doi.org/10.1016%2Fj.epsl.20
13.06.012). Retrieved 18 September 2022.
25. Hollis, Christopher J.; Taylor, Kyle W. R.; Handley, Luke; Pancost, Richard D.; Huber,
Matthew; Creech, John B.; Hines, Benjamin R.; Crouch, Erica M.; Morgans, Hugh E. G.;
Crampton, James S.; Gibbs, Samantha; Pearson, Paul N.; Zachos, James C. (1 October
2012). "Early Paleogene temperature history of the Southwest Pacific Ocean: Reconciling
proxies and models" (https://www.sciencedirect.com/science/article/abs/pii/S0012821X1200
3081). Earth and Planetary Science Letters. 349–350: 53–66.
Bibcode:2012E&PSL.349...53H (https://ui.adsabs.harvard.edu/abs/2012E&PSL.349...53H).
doi:10.1016/j.epsl.2012.06.024 (https://doi.org/10.1016%2Fj.epsl.2012.06.024). Retrieved
18 September 2022.
26. Sluijs, Appy; Bijl, P. K.; Schouten, Stefan; Reichart, G.-J.; Brinkhuis, H. (26 January 2011).
"Southern ocean warming, sea level and hydrological change during the Paleocene-Eocene
thermal maximum" (https://cp.copernicus.org/articles/7/47/2011/). Climate of the Past. 7 (1):
47–61. doi:10.5194/cp-7-47-2011 (https://doi.org/10.5194%2Fcp-7-47-2011). Retrieved
19 May 2023. {{cite journal}}: |first4= missing |last4= (help)
27. Stokke, Ella W.; Jones, Morgan T.; Tierney, Jessica E.; Svensen, Henrik H.; Whiteside,
Jessica H. (15 August 2020). "Temperature changes across the Paleocene-Eocene Thermal
Maximum – a new high-resolution TEX86 temperature record from the Eastern North Sea
Basin" (https://www.sciencedirect.com/science/article/pii/S0012821X20303320). Earth and
Planetary Science Letters. 544: 1–10. doi:10.1016/j.epsl.2020.116388 (https://doi.org/10.101
6%2Fj.epsl.2020.116388). Retrieved 3 July 2023.
28. Moran, K.; Backman, J.; Pagani, others (2006). "The Cenozoic palaeoenvironment of the
Arctic Ocean". Nature. 441 (7093): 601–605. Bibcode:2006Natur.441..601M (https://ui.adsab
s.harvard.edu/abs/2006Natur.441..601M). doi:10.1038/nature04800 (https://doi.org/10.103
8%2Fnature04800). hdl:11250/174276 (https://hdl.handle.net/11250%2F174276).
PMID 16738653 (https://pubmed.ncbi.nlm.nih.gov/16738653). S2CID 4424147 (https://api.se
manticscholar.org/CorpusID:4424147).
29. the dinoflagellates Apectodinium spp.
30. Sluijs, A.; Schouten, S.; Pagani, M.; Woltering, M.; Brinkhuis, H.; Damsté, J.S.S.; Dickens,
G.R.; Huber, M.; Reichart, G.J.; Stein, R.; et al. (2006). "Subtropical Arctic Ocean
temperatures during the Palaeocene/Eocene thermal maximum" (https://brage.npolar.no/npo
lar-xmlui/bitstream/11250/174280/1/SluijsNature2006.pdf) (PDF). Nature. 441 (7093): 610–
613. Bibcode:2006Natur.441..610S (https://ui.adsabs.harvard.edu/abs/2006Natur.441..610
S). doi:10.1038/nature04668 (https://doi.org/10.1038%2Fnature04668). hdl:11250/174280 (h
ttps://hdl.handle.net/11250%2F174280). PMID 16752441 (https://pubmed.ncbi.nlm.nih.gov/1
6752441). S2CID 4412522 (https://api.semanticscholar.org/CorpusID:4412522).
31. Shellito, Cindy J.; Sloan, Lisa C.; Huber, Matthew (2003). "Climate model sensitivity to
atmospheric CO2 levels in the Early-Middle Paleogene". Palaeogeography,
Palaeoclimatology, Palaeoecology. 193 (1): 113–123. Bibcode:2003PPP...193..113S (http
s://ui.adsabs.harvard.edu/abs/2003PPP...193..113S). doi:10.1016/S0031-0182(02)00718-6
(https://doi.org/10.1016%2FS0031-0182%2802%2900718-6).
32. Thierry Adatte; Hassan Khozyem; Jorge E. Spangenberg; Bandana Samant & Gerta Keller
(2014). "Response of terrestrial environment to the Paleocene-Eocene Thermal Maximum
(PETM), new insights from India and NE Spain" (https://www.researchgate.net/publication/2
63430375). Rendiconti della Società Geologica Italiana. 31: 5–6. doi:10.3301/ROL.2014.17
(https://doi.org/10.3301%2FROL.2014.17).
33. Dickens, G.R.; Castillo, M.M.; Walker, J.C.G. (1997). "A blast of gas in the latest Paleocene;
simulating first-order effects of massive dissociation of oceanic methane hydrate" (https://se
manticscholar.org/paper/1d9497804d68d41cde21bbe07b15a204e4a93fc2). Geology. 25
(3): 259–262. Bibcode:1997Geo....25..259D (https://ui.adsabs.harvard.edu/abs/1997Geo....2
5..259D). doi:10.1130/0091-7613(1997)025<0259:abogit>2.3.co;2 (https://doi.org/10.1130%
2F0091-7613%281997%29025%3C0259%3Aabogit%3E2.3.co%3B2). PMID 11541226 (htt
ps://pubmed.ncbi.nlm.nih.gov/11541226). S2CID 24020720 (https://api.semanticscholar.org/
CorpusID:24020720).
34. Zeebe, R.; Zachos, J.C.; Dickens, G.R. (2009). "Carbon dioxide forcing alone insufficient to
explain Palaeocene–Eocene Thermal Maximum warming". Nature Geoscience. 2 (8): 576–
580. Bibcode:2009NatGe...2..576Z (https://ui.adsabs.harvard.edu/abs/2009NatGe...2..576Z).
CiteSeerX 10.1.1.704.7960 (https://citeseerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.704.
7960). doi:10.1038/ngeo578 (https://doi.org/10.1038%2Fngeo578).
35. Norris, R.D.; Röhl, U. (1999). "Carbon cycling and chronology of climate warming during the
Palaeocene/Eocene transition". Nature. 401 (6755): 775–778.
Bibcode:1999Natur.401..775N (https://ui.adsabs.harvard.edu/abs/1999Natur.401..775N).
doi:10.1038/44545 (https://doi.org/10.1038%2F44545). S2CID 4421998 (https://api.semantic
scholar.org/CorpusID:4421998).
36. Panchuk, K.; Ridgwell, A.; Kump, L.R. (2008). "Sedimentary response to Paleocene-Eocene
Thermal Maximum carbon release: A model-data comparison". Geology. 36 (4): 315–318.
Bibcode:2008Geo....36..315P (https://ui.adsabs.harvard.edu/abs/2008Geo....36..315P).
doi:10.1130/G24474A.1 (https://doi.org/10.1130%2FG24474A.1).
37. Cui, Y.; Kump, L.R.; Ridgwell, A.J.; Charles, A.J.; Junium, C.K.; Diefendorf, A.F.; Freeman,
K.H.; Urban, N.M.; Harding, I.C. (2011). "Slow release of fossil carbon during the
Palaeocene-Eocene thermal maximum". Nature Geoscience. 4 (7): 481–485.
Bibcode:2011NatGe...4..481C (https://ui.adsabs.harvard.edu/abs/2011NatGe...4..481C).
doi:10.1038/ngeo1179 (https://doi.org/10.1038%2Fngeo1179).
38. Röhl, U.; Bralower, T.J.; Norris, R.D.; Wefer, G. (2000). "New chronology for the late
Paleocene thermal maximum and its environmental implications". Geology. 28 (10): 927–
930. Bibcode:2000Geo....28..927R (https://ui.adsabs.harvard.edu/abs/2000Geo....28..927R).
doi:10.1130/0091-7613(2000)28<927:NCFTLP>2.0.CO;2 (https://doi.org/10.1130%2F0091-
7613%282000%2928%3C927%3ANCFTLP%3E2.0.CO%3B2).
39. Röhl, Ursula; Westerhold, Thomas; Bralower, Timothy J.; Zachos, James C. (11 December
2007). "On the duration of the Paleocene-Eocene thermal maximum (PETM)" (https://agupu
bs.onlinelibrary.wiley.com/doi/10.1029/2007GC001784). Geochemistry, Geophysics,
Geosystems. 8 (12): 1–13. Bibcode:2007GGG.....812002R (https://ui.adsabs.harvard.edu/ab
s/2007GGG.....812002R). doi:10.1029/2007GC001784 (https://doi.org/10.1029%2F2007GC
001784). S2CID 53349725 (https://api.semanticscholar.org/CorpusID:53349725). Retrieved
8 April 2023.
40. Dickens, G.R. (2000). "Methane oxidation during the late Palaeocene thermal maximum".
Bulletin de la Société Géologique de France. 171: 37–49.
41. Farley, K.A.; Eltgroth, S.F. (2003). "An alternative age model for the Paleocene—Eocene
thermal maximum using extraterrestrial 3He" (https://authors.library.caltech.edu/35478/2/mm
c1.xls). Earth and Planetary Science Letters. 208 (3–4): 135–148.
Bibcode:2003E&PSL.208..135F (https://ui.adsabs.harvard.edu/abs/2003E&PSL.208..135F).
doi:10.1016/S0012-821X(03)00017-7 (https://doi.org/10.1016%2FS0012-821X%2803%290
0017-7).
42. Sluijs, A.; Brinkhuis, H.; Schouten, S.; Bohaty, S.M.; John, C.M.; Zachos, J.C.; Reichart, G.J.;
Sinninghe Damste, J.S.; Crouch, E.M.; Dickens, G.R. (2007). "Environmental precursors to
rapid light carbon injection at the Palaeocene/Eocene boundary". Nature. 450 (7173): 1218–
21. Bibcode:2007Natur.450.1218S (https://ui.adsabs.harvard.edu/abs/2007Natur.450.1218
S). doi:10.1038/nature06400 (https://doi.org/10.1038%2Fnature06400). hdl:1874/31621 (http
s://hdl.handle.net/1874%2F31621). PMID 18097406 (https://pubmed.ncbi.nlm.nih.gov/18097
406). S2CID 4359625 (https://api.semanticscholar.org/CorpusID:4359625).
43. Baczynski, Allison A.; McInerney, Francesca A.; Wing, Scott L.; Kraus, Mary J.; Bloch,
Jonathan I.; Boyer, Doug M.; Secord, Ross; Morse, Paul E.; Fricke, Henry J. (6 September
2013). "Chemostratigraphic implications of spatial variation in the Paleocene-Eocene
Thermal Maximum carbon isotope excursion, SE Bighorn Basin, Wyoming" (https://agupubs.
onlinelibrary.wiley.com/doi/full/10.1002/ggge.20265). Geochemistry, Geophysics,
Geosystems. 14 (10): 4133–4152. doi:10.1002/ggge.20265 (https://doi.org/10.1002%2Fggg
e.20265). Retrieved 19 May 2023.
44. Baczynski, Allison A.; McInerney, Francesca A.; Wing, Scott L.; Kraus, Mary J.; Morse, Paul
E.; Bloch, Jonathan I.; Chung, Angela H.; Freeman, Katherine H. (1 September 2016).
"Distortion of carbon isotope excursion in bulk soil organic matter during the Paleocene-
Eocene thermal maximum" (https://pubs.geoscienceworld.org/gsa/gsabulletin/article-abstrac
t/128/9-10/1352/185391/Distortion-of-carbon-isotope-excursion-in-bulk?redirectedFrom=fullt
ext). Geological Society of America Bulletin. 128 (9–10): 1352–1366. doi:10.1130/B31389.1
(https://doi.org/10.1130%2FB31389.1). Retrieved 19 May 2023.
45. Pagani, M.; Pedentchouk, N.; Huber, M.; Sluijs, A.; Schouten, S.; Brinkhuis, H.; Sinninghe
Damsté, J.S.; Dickens, G.R.; Others (2006). "Arctic hydrology during global warming at the
Palaeocene/Eocene thermal maximum". Nature. 442 (7103): 671–675.
Bibcode:2006Natur.442..671P (https://ui.adsabs.harvard.edu/abs/2006Natur.442..671P).
doi:10.1038/nature05043 (https://doi.org/10.1038%2Fnature05043). hdl:1874/22388 (https://
hdl.handle.net/1874%2F22388). PMID 16906647 (https://pubmed.ncbi.nlm.nih.gov/1690664
7). S2CID 96915252 (https://api.semanticscholar.org/CorpusID:96915252).
46. Speelman, E. N.; van Kempen, M. M. L.; Barke, J.; Brinkhuis, H.; Reichart, G. J.; Smolders,
A. J. P.; Roelofs, J. G. M.; Sangeorgi, F.; de Leeuw, J. W.; Lotter, A. F.; Sinninghe Damest, J.
S. (March 2009). "The Eocene Arctic Azolla bloom: environmental conditions, productivity
and carbon drawdown" (https://www.researchgate.net/publication/24236486). Geobiology. 7
(2): 155–170. Bibcode:2009Gbio....7..155S (https://ui.adsabs.harvard.edu/abs/2009Gbio....
7..155S). doi:10.1111/j.1472-4669.2009.00195.x (https://doi.org/10.1111%2Fj.1472-4669.20
09.00195.x). PMID 19323694 (https://pubmed.ncbi.nlm.nih.gov/19323694).
S2CID 13206343 (https://api.semanticscholar.org/CorpusID:13206343). Retrieved 12 July
2019.
47. Xie, Yulong; Wu, Fuli; Fang, Xiaomin (January 2022). "A transient south subtropical forest
ecosystem in central China driven by rapid global warming during the Paleocene-Eocene
Thermal Maximum" (https://www.sciencedirect.com/science/article/abs/pii/S1342937X21002
471). Gondwana Research. 101: 192–202. Bibcode:2022GondR.101..192X (https://ui.adsab
s.harvard.edu/abs/2022GondR.101..192X). doi:10.1016/j.gr.2021.08.005 (https://doi.org/10.1
016%2Fj.gr.2021.08.005). Retrieved 28 September 2022.
48. Samanta, A.; Bera, M. K.; Ghosh, Ruby; Bera, Subir; Filley, Timothy; Prade, Kanchan;
Rathore, S. S.; Rai, Jyotsana; Sarkar, A. (1 October 2013). "Do the large carbon isotopic
excursions in terrestrial organic matter across Paleocene–Eocene boundary in India
indicate intensification of tropical precipitation?" (https://www.sciencedirect.com/science/arti
cle/abs/pii/S0031018213003295). Palaeogeography, Palaeoclimatology, Palaeoecology.
387: 91–103. Bibcode:2013PPP...387...91S (https://ui.adsabs.harvard.edu/abs/2013PPP...38
7...91S). doi:10.1016/j.palaeo.2013.07.008 (https://doi.org/10.1016%2Fj.palaeo.2013.07.00
8). Retrieved 15 November 2022.
49. Walters, Gregory L.; Kemp, Simon J.; Hemingway, Jordon D.; Johnston, David T.; Hodell,
David A. (22 December 2022). "Clay hydroxyl isotopes show an enhanced hydrologic cycle
during the Paleocene-Eocene Thermal Maximum" (https://www.nature.com/articles/s41467-
022-35545-2). Nature Communications. 13 (1): 1–10. doi:10.1038/s41467-022-35545-2 (http
s://doi.org/10.1038%2Fs41467-022-35545-2). Retrieved 3 July 2023.
50. Garel, Sylvain; Schnyder, Johann; Jacob, Jérémy; Dupuis, Christian; Boussafir, Mohammed;
Le Milbeau, Claude; Storme, Jean-Yves; Iakovleva, Alina I.; Yans, Johan; Baudin, François;
Fléhoc, Christine; Quesnel, Florence (15 April 2013). "Paleohydrological and
paleoenvironmental changes recorded in terrestrial sediments of the Paleocene–Eocene
boundary (Normandy, France)" (https://www.sciencedirect.com/science/article/abs/pii/S0031
018213001223). Palaeogeography, Palaeoclimatology, Palaeoecology. 376: 184–199.
doi:10.1016/j.palaeo.2013.02.035 (https://doi.org/10.1016%2Fj.palaeo.2013.02.035).
Retrieved 11 June 2023.
51. Beard, K. Christopher (11 March 2008). "The oldest North American primate and mammalian
biogeography during the Paleocene–Eocene Thermal Maximum" (https://www.ncbi.nlm.nih.
gov/pmc/articles/PMC2268774). Proceedings of the National Academy of Sciences of the
United States of America. 105 (10): 3815–3818. Bibcode:2008PNAS..105.3815B (https://ui.a
dsabs.harvard.edu/abs/2008PNAS..105.3815B). doi:10.1073/pnas.0710180105 (https://doi.
org/10.1073%2Fpnas.0710180105). PMC 2268774 (https://www.ncbi.nlm.nih.gov/pmc/articl
es/PMC2268774). PMID 18316721 (https://pubmed.ncbi.nlm.nih.gov/18316721).
52. Baczynski, Allison A.; McInerney, Francesca A.; Wing, Scott L.; Kraus, Mary J.; Bloch,
Jonathan I.; Secord, Ross (1 January 2017). "Constraining paleohydrologic change during
the Paleocene-Eocene Thermal Maximum in the continental interior of North America" (http
s://www.sciencedirect.com/science/article/abs/pii/S0031018216306435). Palaeogeography,
Palaeoclimatology, Palaeoecology. 465: 237–246. doi:10.1016/j.palaeo.2016.10.030 (http
s://doi.org/10.1016%2Fj.palaeo.2016.10.030). Retrieved 19 May 2023.
53. Handley, Luke; O'Halloran, Aiofe; Pearson, Paul N.; Hawkins, Elizabeth; Nicholas,
Christopher J.; Schouten, Stefan; McMillan, Ian K.; Pancost, Richard D. (15 April 2012).
"Changes in the hydrological cycle in tropical East Africa during the Paleocene–Eocene
Thermal Maximum" (https://www.sciencedirect.com/science/article/abs/pii/S0031018212000
752). Palaeogeography, Palaeoclimatology, Palaeoecology. 329–330: 10–21.
Bibcode:2012PPP...329...10H (https://ui.adsabs.harvard.edu/abs/2012PPP...329...10H).
doi:10.1016/j.palaeo.2012.02.002 (https://doi.org/10.1016%2Fj.palaeo.2012.02.002).
Retrieved 22 April 2023.
54. Giusberti, L.; Boscolo Galazzo, F.; Thomas, E. (9 February 2016). "Variability in climate and
productivity during the Paleocene–Eocene Thermal Maximum in the western Tethys (Forada
section)" (https://cp.copernicus.org/articles/12/213/2016/). Climate of the Past. 12 (2): 213–
240. doi:10.5194/cp-12-213-2016 (https://doi.org/10.5194%2Fcp-12-213-2016). Retrieved
11 June 2023.
55. Bornemann, André; Norris, Richard D.; Lyman, Johnnie A.; D'haenens, Simon; Groeneveld,
Jeroen; Röhl, Ursula; Farley, Kenneth A.; Speijer, Robert P. (15 May 2014). "Persistent
environmental change after the Paleocene–Eocene Thermal Maximum in the eastern North
Atlantic" (https://www.sciencedirect.com/science/article/abs/pii/S0012821X14001617). Earth
and Planetary Science Letters. 394: 70–81. doi:10.1016/j.epsl.2014.03.017 (https://doi.org/1
0.1016%2Fj.epsl.2014.03.017). Retrieved 11 June 2023.
56. Zachos, J.C.; Röhl, U.; Schellenberg, S.A.; Sluijs, A.; Hodell, D.A.; Kelly, D.C.; Thomas, E.;
Nicolo, M.; Raffi, I.; Lourens, L.J.; et al. (2005). "Rapid Acidification of the Ocean During the
Paleocene-Eocene Thermal Maximum" (http://es.ucsc.edu/%7Ejzachos/pubs/Zachos_etal_
2005A.pdf) (PDF). Science. 308 (5728): 1611–1615. Bibcode:2005Sci...308.1611Z (https://u
i.adsabs.harvard.edu/abs/2005Sci...308.1611Z). doi:10.1126/science.1109004 (https://doi.or
g/10.1126%2Fscience.1109004). hdl:1874/385806 (https://hdl.handle.net/1874%2F385806).
PMID 15947184 (https://pubmed.ncbi.nlm.nih.gov/15947184). S2CID 26909706 (https://api.s
emanticscholar.org/CorpusID:26909706). Archived (https://web.archive.org/web/200809102
14531/http://es.ucsc.edu/%7Ejzachos/pubs/Zachos_etal_2005A.pdf) (PDF) from the original
on 2008-09-10. Retrieved 2008-04-23.
57. Zhou, X.; Thomas, E.; Rickaby, R. E. M.; Winguth, A. M. E.; Lu, Z. (2014). "I/Ca evidence for
global upper ocean deoxygenation during the Paleocene-Eocene Thermal Maximum
(PETM)". Paleoceanography and Paleoclimatology. 29 (10): 964–975.
Bibcode:2014PalOc..29..964Z (https://ui.adsabs.harvard.edu/abs/2014PalOc..29..964Z).
doi:10.1002/2014PA002702 (https://doi.org/10.1002%2F2014PA002702).
58. Carmichael, Matthew J.; Inglis, Gordon N.; Badger, Marcus P. S.; Naafs, B. David A.;
Behrooz, Leila; Remmelzwaal, Serginio; Monteiro, Fanny M.; Rohrssen, Megan; Farnsworth,
Alexander; Buss, Heather L.; Dickson, Alexander J.; Valdes, Paul J.; Lunt, Daniel J.;
Pancost, Richard D. (October 2017). "Hydrological and associated biogeochemical
consequences of rapid global warming during the Paleocene-Eocene Thermal Maximum" (h
ttps://www.sciencedirect.com/science/article/pii/S0921818117300723). Global and
Planetary Change. 157: 114–138. doi:10.1016/j.gloplacha.2017.07.014 (https://doi.org/10.10
16%2Fj.gloplacha.2017.07.014). Retrieved 24 April 2023.
59. Schoon, Petra L.; Heilmann-Clausen, Claus; Schultz, Bo Pagh; Sinninghe Damsté, Jaap S.;
Schouten, Stefan (January 2015). "Warming and environmental changes in the eastern
North Sea Basin during the Palaeocene–Eocene Thermal Maximum as revealed by
biomarker lipids" (https://www.sciencedirect.com/science/article/abs/pii/S014663801400275
7). Organic Geochemistry. 78: 79–88. doi:10.1016/j.orggeochem.2014.11.003 (https://doi.or
g/10.1016%2Fj.orggeochem.2014.11.003). Retrieved 24 April 2023.
60. Sluijs, Appy; Van Roij, L.; Harrington, G. J.; Schouten, Stefan; Sessa, J. A.; LeVay, L. J.;
Reichart, G.-J.; Slomp, C. P. (25 July 2014). "Warming, euxinia and sea level rise during the
Paleocene–Eocene Thermal Maximum on the Gulf Coastal Plain: implications for ocean
oxygenation and nutrient cycling" (https://cp.copernicus.org/articles/10/1421/2014/). Climate
of the Past. 10 (4): 1421–1439. doi:10.5194/cp-10-1421-2014 (https://doi.org/10.5194%2Fcp
-10-1421-2014). Retrieved 3 July 2023.
61. Dickson, Alexander J.; Rees-Owen, Rhian L.; März, Christian; Coe, Angela L.; Cohen,
Anthony S.; Pancost, Richard D.; Taylor, Kyle; Shcherbinina, Ekaterina (30 April 2014). "The
spread of marine anoxia on the northern Tethys margin during the Paleocene-Eocene
Thermal Maximum" (https://agupubs.onlinelibrary.wiley.com/doi/full/10.1002/2014PA00262
9). Paleoceanography and Paleoclimatology. 29 (6): 471–488. doi:10.1002/2014PA002629
(https://doi.org/10.1002%2F2014PA002629). Retrieved 3 July 2023.
62. Von Strandmann, Philip A. E. Pogge; Jones, Morgan T.; West, A. Joshua; Murphy, Melissa J.;
Stokke, Ella W.; Tarbuck, Gary; Wilson, David J.; Pearce, Christopher R.; Schmidt, Daniela
N. (15 October 2021). "Lithium isotope evidence for enhanced weathering and erosion
during the Paleocene-Eocene Thermal Maximum" (https://www.ncbi.nlm.nih.gov/pmc/article
s/PMC8519576). Science Advances. 7 (42): eabh4224. Bibcode:2021SciA....7.4224P (http
s://ui.adsabs.harvard.edu/abs/2021SciA....7.4224P). doi:10.1126/sciadv.abh4224 (https://do
i.org/10.1126%2Fsciadv.abh4224). PMC 8519576 (https://www.ncbi.nlm.nih.gov/pmc/article
s/PMC8519576). PMID 34652934 (https://pubmed.ncbi.nlm.nih.gov/34652934).
63. Nunes, F.; Norris, R.D. (2006). "Abrupt reversal in ocean overturning during the
Palaeocene/Eocene warm period". Nature. 439 (7072): 60–3. Bibcode:2006Natur.439...60N
(https://ui.adsabs.harvard.edu/abs/2006Natur.439...60N). doi:10.1038/nature04386 (https://d
oi.org/10.1038%2Fnature04386). PMID 16397495 (https://pubmed.ncbi.nlm.nih.gov/163974
95). S2CID 4301227 (https://api.semanticscholar.org/CorpusID:4301227).
64. Pak, Dorothy K.; Miller, Kenneth J. (August 1992). "Paleocene to Eocene benthic
foraminiferal isotopes and assemblages: Implications for deepwater circulation" (https://agup
ubs.onlinelibrary.wiley.com/doi/10.1029/92PA01234). Paleoceanography and
Paleoclimatology. 7 (4): 405–422. Bibcode:1992PalOc...7..405P (https://ui.adsabs.harvard.e
du/abs/1992PalOc...7..405P). doi:10.1029/92PA01234 (https://doi.org/10.1029%2F92PA012
34). Retrieved 7 April 2023.
65. Fantle, Matthew S.; Ridgwell, Andy (5 August 2020). "Towards an understanding of the Ca
isotopic signal related to ocean acidification and alkalinity overshoots in the rock record" (htt
ps://www.sciencedirect.com/science/article/abs/pii/S0009254120302114). Chemical
Geology. 547: 119672. doi:10.1016/j.chemgeo.2020.119672 (https://doi.org/10.1016%2Fj.ch
emgeo.2020.119672). S2CID 219461270 (https://api.semanticscholar.org/CorpusID:219461
270). Retrieved 23 April 2023.
66. Kaitlin Alexander; Katrin J. Meissner & Timothy J. Bralower (11 May 2015). "Sudden
spreading of corrosive bottom water during the Palaeocene–Eocene Thermal Maximum".
Nature Geoscience. 8 (6): 458–461. Bibcode:2015NatGe...8..458A (https://ui.adsabs.harvar
d.edu/abs/2015NatGe...8..458A). doi:10.1038/ngeo2430 (https://doi.org/10.1038%2Fngeo24
30).
67. Colosimo, A. B.; Bralower, T. J.; Zachos, James C. (June 2006). "Evidence for Lysocline
Shoaling at the Paleocene/Eocene Thermal Maximum on Shatsky Rise, Northwest Pacific"
(https://www.researchgate.net/publication/279407589). In Bralower, T. J.; Silva, I. Premoli;
Malone, M. J. (eds.). Proceedings of the Ocean Drilling Program, 198 Scientific Results (http
s://www.scienceopen.com/book?vid=200170de-fed7-4b17-bb24-e588d20eb9de).
Proceedings of the Ocean Drilling Program. Vol. 198. Ocean Drilling Program.
doi:10.2973/odp.proc.sr.198.112.2006 (https://doi.org/10.2973%2Fodp.proc.sr.198.112.200
6).
68. Ma, Zhongwu; Gray, Ellen; Thomas, Ellen; Murphy, Brandon; Zachos, James C.; Paytan,
Adina (13 April 2014). "Carbon sequestration during the Palaeocene–Eocene Thermal
Maximum by an efficient biological pump" (https://www.nature.com/articles/ngeo2139).
Nature Geoscience. 7 (1): 382–388. doi:10.1038/ngeo2139 (https://doi.org/10.1038%2Fngeo
2139). Retrieved 27 April 2023.
69. Kelly, D. Clay; Zachos, James C.; Bralower, Timothy J.; Schellenberg, Stephen A. (17
December 2005). "Enhanced terrestrial weathering/runoff and surface ocean carbonate
production during the recovery stages of the Paleocene-Eocene thermal maximum" (https://a
gupubs.onlinelibrary.wiley.com/doi/10.1029/2005PA001163). Paleoceanography and
Paleoclimatology. 20 (4): 1–11. Bibcode:2005PalOc..20.4023K (https://ui.adsabs.harvard.ed
u/abs/2005PalOc..20.4023K). doi:10.1029/2005PA001163 (https://doi.org/10.1029%2F2005
PA001163). Retrieved 10 April 2023.
70. Peter C. Lippert (2008). "Big discovery for biogenic magnetite" (https://www.ncbi.nlm.nih.go
v/pmc/articles/PMC2584755). Proceedings of the National Academy of Sciences of the
United States of America. 105 (46): 17595–17596. Bibcode:2008PNAS..10517595L (https://
ui.adsabs.harvard.edu/abs/2008PNAS..10517595L). doi:10.1073/pnas.0809839105 (https://
doi.org/10.1073%2Fpnas.0809839105). PMC 2584755 (https://www.ncbi.nlm.nih.gov/pmc/ar
ticles/PMC2584755). PMID 19008352 (https://pubmed.ncbi.nlm.nih.gov/19008352).
71. Schumann; et al. (2008). "Gigantism in unique biogenic magnetite at the Paleocene–
Eocene Thermal Maximum" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2584680).
Proceedings of the National Academy of Sciences of the United States of America. 105 (46):
17648–17653. Bibcode:2008PNAS..10517648S (https://ui.adsabs.harvard.edu/abs/2008PN
AS..10517648S). doi:10.1073/pnas.0803634105 (https://doi.org/10.1073%2Fpnas.0803634
105). PMC 2584680 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2584680).
PMID 18936486 (https://pubmed.ncbi.nlm.nih.gov/18936486).
72. O. Strbak; P. Kopcansky; I. Frollo (2011). "Biogenic Magnetite in Humans and New Magnetic
Resonance Hazard Questions" (http://www.measurement.sk/2011/Strbak.pdf) (PDF).
Measurement Science Review. 11 (3): 85. Bibcode:2011MeScR..11...85S (https://ui.adsabs.
harvard.edu/abs/2011MeScR..11...85S). doi:10.2478/v10048-011-0014-1 (https://doi.org/10.
2478%2Fv10048-011-0014-1). S2CID 36212768 (https://api.semanticscholar.org/CorpusID:
36212768). Archived (https://web.archive.org/web/20160304093626/http://www.measureme
nt.sk/2011/Strbak.pdf) (PDF) from the original on 2016-03-04. Retrieved 2015-05-28.
73. Thomas E (1989). "Development of Cenozoic deep-sea benthic foraminiferal faunas in
Antarctic waters". Geological Society of London, Special Publications. 47 (1): 283–296.
Bibcode:1989GSLSP..47..283T (https://ui.adsabs.harvard.edu/abs/1989GSLSP..47..283T).
doi:10.1144/GSL.SP.1989.047.01.21 (https://doi.org/10.1144%2FGSL.SP.1989.047.01.21).
S2CID 37660762 (https://api.semanticscholar.org/CorpusID:37660762).
74. Thomas E (1990). "Late Cretaceous–early Eocene mass extinctions in the deep sea".
Global Catastrophes in Earth History; an Interdisciplinary Conference on Impacts,
Volcanism, and Mass Mortality. GSA Special Papers. Vol. 247. pp. 481–495.
doi:10.1130/SPE247-p481 (https://doi.org/10.1130%2FSPE247-p481). ISBN 0-8137-2247-
0.
75. Thomas, E. (1998). "The biogeography of the late Paleocene benthic foraminiferal
extinction". In Aubry, M.-P.; Lucas, S.; Berggren, W. A. (eds.). Late Paleocene-early Eocene
Biotic and Climatic Events in the Marine and Terrestrial Records. Columbia University
Press. pp. 214–243.
76. Takeda, Kotaro; Kaiho, Kunio (3 August 2007). "Faunal turnovers in central Pacific benthic
foraminifera during the Paleocene–Eocene thermal maximum" (https://www.sciencedirect.co
m/science/article/abs/pii/S0031018207000983). Palaeogeography, Palaeoclimatology,
Palaeoecology. 251 (2): 175–197. doi:10.1016/j.palaeo.2007.02.026 (https://doi.org/10.101
6%2Fj.palaeo.2007.02.026). Retrieved 3 July 2023.
77. Li, Juan; Hu, Xiumian; Zachos, James C.; Garzanti, Eduardo; BouDagher-Fadel, Marcelle
(November 2020). "Sea level, biotic and carbon-isotope response to the Paleocene–Eocene
thermal maximum in Tibetan Himalayan platform carbonates" (https://www.sciencedirect.co
m/science/article/abs/pii/S0921818120302071). Global and Planetary Change. 194:
103316. Bibcode:2020GPC...19403316L (https://ui.adsabs.harvard.edu/abs/2020GPC...194
03316L). doi:10.1016/j.gloplacha.2020.103316 (https://doi.org/10.1016%2Fj.gloplacha.202
0.103316). S2CID 222117770 (https://api.semanticscholar.org/CorpusID:222117770).
Retrieved 17 April 2023.
78. Bralower, Timothy J. (31 May 2002). "Evidence of surface water oligotrophy during the
Paleocene-Eocene thermal maximum: Nanofossil assemblage data from Ocean Drilling
Program Site 690, Maud Rise, Weddell Sea" (https://agupubs.onlinelibrary.wiley.com/doi/10.
1029/2001PA000662). Paleoceanography and Paleoclimatology. 17 (2): 13-1–13-12.
Bibcode:2002PalOc..17.1023B (https://ui.adsabs.harvard.edu/abs/2002PalOc..17.1023B).
doi:10.1029/2001PA000662 (https://doi.org/10.1029%2F2001PA000662). Retrieved 11 April
2023.
79. Hupp, Brittany N.; Kelly, D. Clay; Williams, John W. (22 February 2022). "Isotopic filtering
reveals high sensitivity of planktic calcifiers to Paleocene–Eocene thermal maximum
warming and acidification" (https://www.pnas.org/doi/abs/10.1073/pnas.2115561119).
Proceedings of the National Academy of Sciences of the United States of America. 119 (9):
1–7. doi:10.1073/pnas.2115561119 (https://doi.org/10.1073%2Fpnas.2115561119).
Retrieved 11 June 2023.
80. Thomas, E. (2007). "Cenozoic mass extinctions in the deep sea: What perturbs the largest
habitat on Earth?". In Monechi, S.; Coccioni, R.; Rampino, M. (eds.). Large Ecosystem
Perturbations: Causes and Consequences. GSA Special Papers. Vol. 424. pp. 1–24.
doi:10.1130/2007.2424(01) (https://doi.org/10.1130%2F2007.2424%2801%29).
ISBN 9780813724249.
81. Winguth A, Thomas E, Winguth C (2012). "Global decline in ocean ventilation, oxygenation
and productivity during the Paleocene-Eocene Thermal Maximum – Implications for the
benthic extinction". Geology. 40 (3): 263–266. Bibcode:2012Geo....40..263W (https://ui.adsa
bs.harvard.edu/abs/2012Geo....40..263W). doi:10.1130/G32529.1 (https://doi.org/10.1130%2
FG32529.1).
82. Ma Z, Gray E, Thomas E, Murphy B, Zachos JC, Paytan A (2014). "Carbon sequestration
during the Paleocene-Eocene Thermal maximum by an efficient biological pump". Nature
Geoscience. 7 (5): 382–388. Bibcode:2014NatGe...7..382M (https://ui.adsabs.harvard.edu/a
bs/2014NatGe...7..382M). doi:10.1038/NGEO2139 (https://doi.org/10.1038%2FNGEO2139).
83. Langdon, C.; Takahashi, T.; Sweeney, C.; Chipman, D.; Goddard, J.; Marubini, F.; Aceves, H.;
Barnett, H.; Atkinson, M.J. (2000). "Effect of calcium carbonate saturation state on the
calcification rate of an experimental coral reef". Global Biogeochemical Cycles. 14 (2): 639–
654. Bibcode:2000GBioC..14..639L (https://ui.adsabs.harvard.edu/abs/2000GBioC..14..639
L). doi:10.1029/1999GB001195 (https://doi.org/10.1029%2F1999GB001195).
S2CID 128987509 (https://api.semanticscholar.org/CorpusID:128987509).
84. Riebesell, U.; Zondervan, I.; Rost, B.; Tortell, P.D.; Zeebe, R.E.; Morel, F.M.M. (2000).
"Reduced calcification of marine plankton in response to increased atmospheric CO2" (http
s://epic.awi.de/id/eprint/3784/1/Rie2000a.pdf) (PDF). Nature. 407 (6802): 364–367.
Bibcode:2000Natur.407..364R (https://ui.adsabs.harvard.edu/abs/2000Natur.407..364R).
doi:10.1038/35030078 (https://doi.org/10.1038%2F35030078). PMID 11014189 (https://pub
med.ncbi.nlm.nih.gov/11014189). S2CID 4426501 (https://api.semanticscholar.org/CorpusI
D:4426501).
85. Iglesias-Rodriguez, M. Debora; Halloran, Paul R.; Rickaby, Rosalind E. M.; Hall, Ian R.;
Colmenero-Hidalgo, Elena; Gittins, John R.; Green, Darryl R. H.; Tyrrell, Toby; Gibbs,
Samantha J.; von Dassow, Peter; Rehm, Eric; Armbrust, E. Virginia; Boessenkool, Karin P.
(April 2008). "Phytoplankton Calcification in a High-CO2 World" (https://semanticscholar.org/
paper/a2154e884200c2f46e1850bdc3e6b4fcc8265b21). Science. 320 (5874): 336–40.
Bibcode:2008Sci...320..336I (https://ui.adsabs.harvard.edu/abs/2008Sci...320..336I).
doi:10.1126/science.1154122 (https://doi.org/10.1126%2Fscience.1154122).
PMID 18420926 (https://pubmed.ncbi.nlm.nih.gov/18420926). S2CID 206511068 (https://ap
i.semanticscholar.org/CorpusID:206511068).
86. Gibbs, Samantha J.; Bown, Paul R.; Sessa, Jocelyn A.; Bralower, Timothy J.; Wilson, Paul A.
(15 December 2006). "Nannoplankton Extinction and Origination Across the Paleocene-
Eocene Thermal Maximum" (https://www.science.org/doi/full/10.1126/science.1133902).
Science. 314 (5806): 1770–1173. Bibcode:2006Sci...314.1770G (https://ui.adsabs.harvard.e
du/abs/2006Sci...314.1770G). doi:10.1126/science.1133902 (https://doi.org/10.1126%2Fsci
ence.1133902). PMID 17170303 (https://pubmed.ncbi.nlm.nih.gov/17170303).
S2CID 41286627 (https://api.semanticscholar.org/CorpusID:41286627). Retrieved 16 April
2023.
87. Kelly, D.C.; Bralower, T.J.; Zachos, J.C. (1998). "Evolutionary consequences of the latest
Paleocene thermal maximum for tropical planktonic foraminifera". Palaeogeography,
Palaeoclimatology, Palaeoecology. 141 (1): 139–161. Bibcode:1998PPP...141..139K (http
s://ui.adsabs.harvard.edu/abs/1998PPP...141..139K). doi:10.1016/S0031-0182(98)00017-0
(https://doi.org/10.1016%2FS0031-0182%2898%2900017-0).
88. He, Tianchen; Kemp, David B.; Li, Juan; Ruhl, Micha (March 2023). "Paleoenvironmental
changes across the Mesozoic–Paleogene hyperthermal events" (https://www.sciencedirect.c
om/science/article/pii/S0921818123000310). Global and Planetary Change. 222: 1–8.
doi:10.1016/j.gloplacha.2023.104058 (https://doi.org/10.1016%2Fj.gloplacha.2023.104058).
Retrieved 3 July 2023.
89. Sanaa El-Sayed; et al. (2021). "Diverse marine fish assemblages inhabited the paleotropics
during the Paleocene-Eocene thermal maximum". Geology. 49 (8): 993–998.
Bibcode:2021Geo....49..993E (https://ui.adsabs.harvard.edu/abs/2021Geo....49..993E).
doi:10.1130/G48549.1 (https://doi.org/10.1130%2FG48549.1). S2CID 236585231 (https://ap
i.semanticscholar.org/CorpusID:236585231).
90. Gingerich, P.D. (2003). "Mammalian responses to climate change at the Paleocene-Eocene
boundary: Polecat Bench record in the northern Bighorn Basin, Wyoming" (http://www-perso
nal.umich.edu/~gingeric/PDFfiles/PDG402_Mammresppebound.pdf) (PDF). In Wing, Scott
L. (ed.). Causes and Consequences of Globally Warm Climates in the Early Paleogene.
GSA Special Papers. Vol. 369. Geological Society of America. pp. 463–78. doi:10.1130/0-
8137-2369-8.463 (https://doi.org/10.1130%2F0-8137-2369-8.463). ISBN 978-0-8137-2369-
3.
91. D'Ambrosia, Abigail R.; Clyde, William C.; Fricke, Henry C.; Gingerich, Philip D.; Abels,
Hemmo A. (15 March 2017). "Repetitive mammalian dwarfing during ancient greenhouse
warming events" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5351980). Science
Advances. 3 (3): e1601430. Bibcode:2017SciA....3E1430D (https://ui.adsabs.harvard.edu/ab
s/2017SciA....3E1430D). doi:10.1126/sciadv.1601430 (https://doi.org/10.1126%2Fsciadv.16
01430). PMC 5351980 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5351980).
PMID 28345031 (https://pubmed.ncbi.nlm.nih.gov/28345031).
92. Secord, R.; Bloch, J. I.; Chester, S. G. B.; Boyer, D. M.; Wood, A. R.; Wing, S. L.; Kraus, M. J.;
McInerney, F. A.; Krigbaum, J. (2012). "Evolution of the Earliest Horses Driven by Climate
Change in the Paleocene-Eocene Thermal Maximum" (http://digitalcommons.unl.edu/cgi/vie
wcontent.cgi?article=1314&context=geosciencefacpub). Science. 335 (6071): 959–962.
Bibcode:2012Sci...335..959S (https://ui.adsabs.harvard.edu/abs/2012Sci...335..959S).
doi:10.1126/science.1213859 (https://doi.org/10.1126%2Fscience.1213859).
PMID 22363006 (https://pubmed.ncbi.nlm.nih.gov/22363006). S2CID 4603597 (https://api.se
manticscholar.org/CorpusID:4603597). Archived (https://web.archive.org/web/20190205182
314/http://digitalcommons.unl.edu/cgi/viewcontent.cgi?article=1314&context=geosciencefac
pub) from the original on 2019-02-05. Retrieved 2018-12-23.
93. Solé, Floréal; Morse, Paul E.; Bloch, Jonathan I.; Gingerich, Philip D.; Smith, Thierry (July
2021). "New specimens of the mesonychid Dissacus praenuntius from the early Eocene of
Wyoming and evaluation of body size through the PETM in North America" (https://www.scie
ncedirect.com/science/article/abs/pii/S0016699521000255). Geobios. 66–67: 103–118.
Bibcode:2021Geobi..66..103S (https://ui.adsabs.harvard.edu/abs/2021Geobi..66..103S).
doi:10.1016/j.geobios.2021.02.005 (https://doi.org/10.1016%2Fj.geobios.2021.02.005).
S2CID 234877826 (https://api.semanticscholar.org/CorpusID:234877826). Retrieved
3 January 2023.
94. Bowen, Gabriel J.; Clyde, William C.; Koch, Paul L.; Ting, Suyin; Alroy, John; Tsubamoto,
Takehisa; Wang, Yuanqing; Wang, Yuan (15 March 2002). "Mammalian Dispersal at the
Paleocene/Eocene Boundary" (https://www.science.org/doi/full/10.1126/science.1068700).
Science. 295 (5562): 2062–2065. Bibcode:2002Sci...295.2062B (https://ui.adsabs.harvard.e
du/abs/2002Sci...295.2062B). doi:10.1126/science.1068700 (https://doi.org/10.1126%2Fsci
ence.1068700). PMID 11896275 (https://pubmed.ncbi.nlm.nih.gov/11896275).
S2CID 10729711 (https://api.semanticscholar.org/CorpusID:10729711). Retrieved 15 April
2023.
95. Currano, Ellen C.; Wilf, Peter; Wild, Scott L.; Labandeira, Conrad C.; Lovecock, Elizabeth C.;
Royer, Dana L. (12 February 2008). "Sharply increased insect herbivory during the
Paleocene–Eocene Thermal Maximum" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC253
8865). Proceedings of the National Academy of Sciences of the United States of America.
105 (6): 1960–1964. doi:10.1073/pnas.0708646105 (https://doi.org/10.1073%2Fpnas.07086
46105). PMC 2538865 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2538865).
PMID 18268338 (https://pubmed.ncbi.nlm.nih.gov/18268338).
96. Aria, Cédric; Jouault, Corentin; Perrichot, Vincent; Nel, André (2 February 2023). "The
megathermal ant genus Gesomyrmex (Formicidae: Formicinae), palaeoindicator of wide
latitudinal biome homogeneity during the PETM" (https://pubs.geoscienceworld.org/geolma
g/article-abstract/160/1/187/621504/The-megathermal-ant-genus-Gesomyrmex-Formicidae).
Geological Magazine. 160 (1): 187–197. Bibcode:2023GeoM..160..187A (https://ui.adsabs.h
arvard.edu/abs/2023GeoM..160..187A). doi:10.1017/S0016756822001248 (https://doi.org/1
0.1017%2FS0016756822001248). S2CID 256564242 (https://api.semanticscholar.org/Corp
usID:256564242). Retrieved 11 April 2023.
97. Smith, Jon J.; Hasiotis, Stephen T.; Kraus, Mary J.; Woody, Daniel T. (20 October 2009).
"Transient dwarfism of soil fauna during the Paleocene–Eocene Thermal Maximum" (https://
www.pnas.org/doi/abs/10.1073/pnas.0909674106). Proceedings of the National Academy of
Sciences of the United States of America. 106 (42): 17655–17660.
doi:10.1073/pnas.0909674106 (https://doi.org/10.1073%2Fpnas.0909674106). Retrieved
11 June 2023.
98. Korasidis, Vera A.; Wing, Scott L.; Shields, Christine A.; Kiehl, Jeffrey T. (9 April 2022).
"Global Changes in Terrestrial Vegetation and Continental Climate During the Paleocene-
Eocene Thermal Maximum" (https://agupubs.onlinelibrary.wiley.com/doi/10.1029/2021PA00
4325). Paleoceanography and Paleoclimatology. 37 (4): 1–21. doi:10.1029/2021PA004325
(https://doi.org/10.1029%2F2021PA004325). Retrieved 28 April 2023.
99. Clyde, William C.; Gingerich, Philip D.; Wing, S. L.; Röhl, Ursula; Westerhold, T.; Bowen, G.;
Johnson, K.; Baczynski, A. A.; Diefendorf, A.; McInerney, F.; Schnurrenberger, D.; Noren, A.;
Brady, K. (5 November 2013). "Bighorn Basin Coring Project (BBCP): a continental
perspective on early Paleogene hyperthermals" (https://sd.copernicus.org/articles/16/21/201
3/). Scientific Drilling. 16: 21–31. Bibcode:2013SciDr..16...21C (https://ui.adsabs.harvard.ed
u/abs/2013SciDr..16...21C). doi:10.5194/sd-16-21-2013 (https://doi.org/10.5194%2Fsd-16-2
1-2013). Retrieved 30 December 2022.
100. Choudhury, Tathagata Roy; Khanolkar, Sonal; Banerjee, Santanu (July 2022). "Glauconite
authigenesis during the warm climatic events of Paleogene: Case studies from shallow
marine sections of Western India" (https://www.sciencedirect.com/science/article/abs/pii/S09
21818122001242?via%3Dihub). Global and Planetary Change. 214.
doi:10.1016/j.gloplacha.2022.103857 (https://doi.org/10.1016%2Fj.gloplacha.2022.103857).
Retrieved 3 July 2023.
101. Pujalte, Victoriano; Schmitz, Birger; Payros, Aitor (1 March 2022). "A rapid sedimentary
response to the Paleocene-Eocene Thermal Maximum hydrological change: New data from
alluvial units of the Tremp-Graus Basin (Spanish Pyrenees)" (https://www.sciencedirect.com/
science/article/pii/S0031018221006039). Palaeogeography, Palaeoclimatology,
Palaeoecology. 589: 110818. Bibcode:2022PPP...589k0818P (https://ui.adsabs.harvard.ed
u/abs/2022PPP...589k0818P). doi:10.1016/j.palaeo.2021.110818 (https://doi.org/10.1016%2
Fj.palaeo.2021.110818). Retrieved 16 January 2023.
102. Giusberti, L.; Rio, D.; Agnini, C.; Backman, J.; Fornaciari, E.; Tateo, F.; Oddone, M. (2007).
"Mode and tempo of the Paleocene-Eocene thermal maximum in an expanded section from
the Venetian pre-Alps". Geological Society of America Bulletin. 119 (3–4): 391–412.
Bibcode:2007GSAB..119..391G (https://ui.adsabs.harvard.edu/abs/2007GSAB..119..391G).
doi:10.1130/B25994.1 (https://doi.org/10.1130%2FB25994.1).
103. Kender, Sev; Bogus, Kara; Pedersen, Gunver K.; Dybkjær, Karen; Mather, Tamsin A.;
Mariani, Erica; Ridgwell, Andy; Riding, James B.; Wagner, Thomas; Hesselbo, Stephen P.;
Leng, Melanie J. (31 August 2021). "Paleocene/Eocene carbon feedbacks triggered by
volcanic activity" (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8408262). Nature
Communications. 12 (1): 5186. Bibcode:2021NatCo..12.5186K (https://ui.adsabs.harvard.ed
u/abs/2021NatCo..12.5186K). doi:10.1038/s41467-021-25536-0 (https://doi.org/10.1038%2F
s41467-021-25536-0). hdl:10871/126942 (https://hdl.handle.net/10871%2F126942).
ISSN 2041-1723 (https://www.worldcat.org/issn/2041-1723). PMC 8408262 (https://www.ncb
i.nlm.nih.gov/pmc/articles/PMC8408262). PMID 34465785 (https://pubmed.ncbi.nlm.nih.gov/
34465785).
104. Carozza, D. A.; Mysak, L. A.; Schmidt, G. A. (2011). "Methane and environmental change
during the Paleocene-Eocene thermal maximum (PETM): Modeling the PETM onset as a
two-stage event". Geophysical Research Letters. 38 (5): L05702.
Bibcode:2011GeoRL..38.5702C (https://ui.adsabs.harvard.edu/abs/2011GeoRL..38.5702C).
doi:10.1029/2010GL046038 (https://doi.org/10.1029%2F2010GL046038).
S2CID 129460348 (https://api.semanticscholar.org/CorpusID:129460348).
105. Patterson, M. V.; Francis, D. (2013). "Kimberlite eruptions as triggers for early Cenozoic
hyperthermals" (https://doi.org/10.1002%2Fggge.20054). Geochemistry, Geophysics,
Geosystems. 14 (2): 448–456. Bibcode:2013GGG....14..448P (https://ui.adsabs.harvard.edu/
abs/2013GGG....14..448P). doi:10.1002/ggge.20054 (https://doi.org/10.1002%2Fggge.2005
4).
106. Svensen, H.; Planke, S.; Malthe-Sørenssen, A.; Jamtveit, B.; Myklebust, R.; Eidem, T.; Rey,
S. S. (2004). "Release of methane from a volcanic basin as a mechanism for initial Eocene
global warming". Nature. 429 (6991): 542–545. Bibcode:2004Natur.429..542S (https://ui.ads
abs.harvard.edu/abs/2004Natur.429..542S). doi:10.1038/nature02566 (https://doi.org/10.103
8%2Fnature02566). PMID 15175747 (https://pubmed.ncbi.nlm.nih.gov/15175747).
S2CID 4419088 (https://api.semanticscholar.org/CorpusID:4419088).
107. Storey, M.; Duncan, R.A.; Swisher III, C.C. (2007). "Paleocene-Eocene Thermal Maximum
and the Opening of the Northeast Atlantic" (https://semanticscholar.org/paper/591958e58752
82bb62373d03fcea48e08c00215d). Science. 316 (5824): 587–9.
Bibcode:2007Sci...316..587S (https://ui.adsabs.harvard.edu/abs/2007Sci...316..587S).
doi:10.1126/science.1135274 (https://doi.org/10.1126%2Fscience.1135274).
PMID 17463286 (https://pubmed.ncbi.nlm.nih.gov/17463286). S2CID 6145117 (https://api.se
manticscholar.org/CorpusID:6145117).
108. Jason Wolfe (5 September 2000). "Volcanoes and Climate Change" (http://earthobservatory.
nasa.gov/Features/Volcano/). Earth Observatory. NASA. Archived (https://web.archive.org/w
eb/20170711231320/https://earthobservatory.nasa.gov/Features/Volcano/) from the original
on 11 July 2017. Retrieved 19 February 2009.
109. Bralower, T.J.; Thomas, D.J.; Zachos, J.C.; Hirschmann, M.M.; Röhl, U.; Sigurdsson, H.;
Thomas, E.; Whitney, D.L. (1997). "High-resolution records of the late Paleocene thermal
maximum and circum-Caribbean volcanism: Is there a causal link?". Geology. 25 (11): 963–
966. Bibcode:1997Geo....25..963B (https://ui.adsabs.harvard.edu/abs/1997Geo....25..963B).
doi:10.1130/0091-7613(1997)025<0963:HRROTL>2.3.CO;2 (https://doi.org/10.1130%2F00
91-7613%281997%29025%3C0963%3AHRROTL%3E2.3.CO%3B2).
110. Jin, Simin; Kemp, David B.; Yin, Runsheng; Sun, Ruyang; Shen, Jun; Jolley, David W.;
Vieira, Manuel; Huang, Chunju (15 January 2023). "Mercury isotope evidence for protracted
North Atlantic magmatism during the Paleocene-Eocene Thermal Maximum" (https://www.sc
iencedirect.com/science/article/abs/pii/S0012821X22005623). Earth and Planetary Science
Letters. 602. doi:10.1016/j.epsl.2022.117926 (https://doi.org/10.1016%2Fj.epsl.2022.11792
6). Retrieved 3 July 2023.
111. Dickson, Alexander J.; Cohen, Anthony S.; Coe, Angela L.; Davies, Marc; Shcherbinina,
Ekaterina A.; Gavrilov, Yuri O. (15 November 2015). "Evidence for weathering and volcanism
during the PETM from Arctic Ocean and Peri-Tethys osmium isotope records" (https://www.s
ciencedirect.com/science/article/abs/pii/S0031018215004551). Palaeogeography,
Palaeoclimatology, Palaeoecology. 438: 300–307. doi:10.1016/j.palaeo.2015.08.019 (http
s://doi.org/10.1016%2Fj.palaeo.2015.08.019). Retrieved 11 June 2023.
112. Piedrahita, Victor A.; Galeotti, Simone; Zhao, Xiang; Roberts, Andrew P.; Rohling, Eelco J.;
Heslop, David; Florindo, Fabio; Grant, Katharine M.; Rodríguez-Sanz, Laura; Reghellin,
Daniele; Zeebe, Richard E. (15 November 2022). "Orbital phasing of the Paleocene-Eocene
Thermal Maximum" (https://www.sciencedirect.com/science/article/abs/pii/S0012821X22004
757). Earth and Planetary Science Letters. 598: 117839. Bibcode:2022E&PSL.59817839P
(https://ui.adsabs.harvard.edu/abs/2022E&PSL.59817839P).
doi:10.1016/j.epsl.2022.117839 (https://doi.org/10.1016%2Fj.epsl.2022.117839).
S2CID 252730173 (https://api.semanticscholar.org/CorpusID:252730173). Retrieved
22 November 2022.
113. Lee, Mingsong; Bralower, Timothy J.; Kump, Lee R.; Self-Trail, Jean M.; Zachos, James C.;
Rush, William D.; Robinson, Marci M. (24 September 2022). "Astrochronology of the
Paleocene-Eocene Thermal Maximum on the Atlantic Coastal Plain" (https://www.ncbi.nlm.n
ih.gov/pmc/articles/PMC9509358). Nature Communications. 13 (1): 5618.
Bibcode:2022NatCo..13.5618L (https://ui.adsabs.harvard.edu/abs/2022NatCo..13.5618L).
doi:10.1038/s41467-022-33390-x (https://doi.org/10.1038%2Fs41467-022-33390-x).
PMC 9509358 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC9509358). PMID 36153313
(https://pubmed.ncbi.nlm.nih.gov/36153313).
114. Lourens, L.J.; Sluijs, A.; Kroon, D.; Zachos, J.C.; Thomas, E.; Röhl, U.; Bowles, J.; Raffi, I.
(2005). "Astronomical pacing of late Palaeocene to early Eocene global warming events".
Nature. 435 (7045): 1083–1087. Bibcode:2005Natur.435.1083L (https://ui.adsabs.harvard.ed
u/abs/2005Natur.435.1083L). doi:10.1038/nature03814 (https://doi.org/10.1038%2Fnature03
814). hdl:1874/11299 (https://hdl.handle.net/1874%2F11299). PMID 15944716 (https://pubm
ed.ncbi.nlm.nih.gov/15944716). S2CID 2139892 (https://api.semanticscholar.org/CorpusID:2
139892).
115. Cramer, Benjamin S.; Wright, James D.; Kent, Dennis V.; Aubry, Marie-Pierre (18 December
2003). "Orbital climate forcing of δ13C excursions in the late Paleocene–early Eocene
(chrons C24n–C25n)" (https://www.researchgate.net/publication/252649720).
Paleoceanography and Paleoclimatology. 18 (4): 1097. Bibcode:2003PalOc..18.1097C (http
s://ui.adsabs.harvard.edu/abs/2003PalOc..18.1097C). doi:10.1029/2003PA000909 (https://d
oi.org/10.1029%2F2003PA000909). Retrieved 16 April 2023.
116. Kent, D.V.; Cramer, B.S.; Lanci, L.; Wang, D.; Wright, J.D.; Van Der Voo, R. (2003). "A case
for a comet impact trigger for the Paleocene/Eocene thermal maximum and carbon isotope
excursion". Earth and Planetary Science Letters. 211 (1–2): 13–26.
Bibcode:2003E&PSL.211...13K (https://ui.adsabs.harvard.edu/abs/2003E&PSL.211...13K).
doi:10.1016/S0012-821X(03)00188-2 (https://doi.org/10.1016%2FS0012-821X%2803%290
0188-2).
117. Kopp, R.E.; Raub, T.; Schumann, D.; Vali, H.; Smirnov, A.V.; Kirschvink, J.L. (2007).
"Magnetofossil spike during the Paleocene-Eocene thermal maximum: Ferromagnetic
resonance, rock magnetic, and electron microscopy evidence from Ancora, New Jersey,
United States" (https://doi.org/10.1029%2F2007PA001473). Paleoceanography and
Paleoclimatology. 22 (4): PA4103. Bibcode:2007PalOc..22.4103K (https://ui.adsabs.harvard.
edu/abs/2007PalOc..22.4103K). doi:10.1029/2007PA001473 (https://doi.org/10.1029%2F20
07PA001473).
118. Wang, H.; Kent, Dennis V.; Jackson, Michael J. (2012). "Evidence for abundant isolated
magnetic nanoparticles at the Paleocene–Eocene boundary" (https://www.ncbi.nlm.nih.gov/
pmc/articles/PMC3545797). Proceedings of the National Academy of Sciences of the United
States of America. 110 (2): 425–430. Bibcode:2013PNAS..110..425W (https://ui.adsabs.harv
ard.edu/abs/2013PNAS..110..425W). doi:10.1073/pnas.1205308110 (https://doi.org/10.107
3%2Fpnas.1205308110). PMC 3545797 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC35
45797). PMID 23267095 (https://pubmed.ncbi.nlm.nih.gov/23267095).
119. Schaller, M. F.; Fung, M. K.; Wright, J. D.; Katz, M. E.; Kent, D. V. (2016). "Impact ejecta at the
Paleocene-Eocene boundary". Science. 354 (6309): 225–229. Bibcode:2016Sci...354..225S
(https://ui.adsabs.harvard.edu/abs/2016Sci...354..225S). doi:10.1126/science.aaf5466 (http
s://doi.org/10.1126%2Fscience.aaf5466). ISSN 0036-8075 (https://www.worldcat.org/issn/00
36-8075). PMID 27738171 (https://pubmed.ncbi.nlm.nih.gov/27738171). S2CID 30852592
(https://api.semanticscholar.org/CorpusID:30852592).
120. Timmer, John (2016-10-13). "Researchers push argument that comet caused ancient climate
change" (https://arstechnica.com/science/2016/10/researchers-push-argument-that-comet-c
aused-ancient-climate-change/). Ars Technica. Archived (https://web.archive.org/web/20161
013220339/http://arstechnica.com/science/2016/10/researchers-push-argument-that-comet-
caused-ancient-climate-change/) from the original on 2016-10-13. Retrieved 2016-10-13.
121. Moore, E; Kurtz, Andrew C. (2008). "Black carbon in Paleocene-Eocene boundary
sediments: A test of biomass combustion as the PETM trigger". Palaeogeography,
Palaeoclimatology, Palaeoecology. 267 (1–2): 147–152. Bibcode:2008PPP...267..147M (htt
ps://ui.adsabs.harvard.edu/abs/2008PPP...267..147M). doi:10.1016/j.palaeo.2008.06.010 (ht
tps://doi.org/10.1016%2Fj.palaeo.2008.06.010).
122. Bowen, Gabriel J. (October 2013). "Up in smoke: A role for organic carbon feedbacks in
Paleogene hyperthermals" (https://www.sciencedirect.com/science/article/abs/pii/S0921818
113001550). Global and Planetary Change. 109: 18–29.
doi:10.1016/j.gloplacha.2013.07.001 (https://doi.org/10.1016%2Fj.gloplacha.2013.07.001).
Retrieved 19 May 2023.
123. Dickens, G. R. (5 August 2011). "Down the Rabbit Hole: toward appropriate discussion of
methane release from gas hydrate systems during the Paleocene-Eocene thermal maximum
and other past hyperthermal events" (https://cp.copernicus.org/articles/7/831/2011/). Climate
of the Past. 7 (3): 831–846. Bibcode:2011CliPa...7..831D (https://ui.adsabs.harvard.edu/abs/
2011CliPa...7..831D). doi:10.5194/cp-7-831-2011 (https://doi.org/10.5194%2Fcp-7-831-201
1). S2CID 55252499 (https://api.semanticscholar.org/CorpusID:55252499). Retrieved
11 April 2023.
124. Thomas, D.J.; Zachos, J.C.; Bralower, T.J.; Thomas, E.; Bohaty, S. (2002). "Warming the fuel
for the fire: Evidence for the thermal dissociation of methane hydrate during the Paleocene-
Eocene thermal maximum" (https://web.archive.org/web/20190108010303/https://wesschola
r.wesleyan.edu/div3facpubs/106/). Geology. 30 (12): 1067–1070.
Bibcode:2002Geo....30.1067T (https://ui.adsabs.harvard.edu/abs/2002Geo....30.1067T).
doi:10.1130/0091-7613(2002)030<1067:WTFFTF>2.0.CO;2 (https://doi.org/10.1130%2F009
1-7613%282002%29030%3C1067%3AWTFFTF%3E2.0.CO%3B2). Archived from the
original (http://wesscholar.wesleyan.edu/div3facpubs/106) on 2019-01-08. Retrieved
2018-12-23.
125. Tripati, A.; Elderfield, H. (2005). "Deep-Sea Temperature and Circulation Changes at the
Paleocene-Eocene Thermal Maximum". Science. 308 (5730): 1894–1898.
Bibcode:2005Sci...308.1894T (https://ui.adsabs.harvard.edu/abs/2005Sci...308.1894T).
doi:10.1126/science.1109202 (https://doi.org/10.1126%2Fscience.1109202).
PMID 15976299 (https://pubmed.ncbi.nlm.nih.gov/15976299). S2CID 38935414 (https://api.s
emanticscholar.org/CorpusID:38935414).
126. Kelly, D. Clay (28 December 2002). "Response of Antarctic (ODP Site 690) planktonic
foraminifera to the Paleocene–Eocene thermal maximum: Faunal evidence for
ocean/climate change" (https://agupubs.onlinelibrary.wiley.com/doi/full/10.1029/2002PA000
761). Paleoceanography and Paleoclimatology. 17 (4): 23-1–23-13.
Bibcode:2002PalOc..17.1071K (https://ui.adsabs.harvard.edu/abs/2002PalOc..17.1071K).
doi:10.1029/2002PA000761 (https://doi.org/10.1029%2F2002PA000761). Retrieved 6 April
2023.
127. Higgins, John A.; Schrag, Daniel P. (30 May 2006). "Beyond methane: Towards a theory for
the Paleocene–Eocene Thermal Maximum" (https://www.sciencedirect.com/science/article/a
bs/pii/S0012821X06002147). Earth and Planetary Science Letters. 245 (3–4): 523–537.
Bibcode:2006E&PSL.245..523H (https://ui.adsabs.harvard.edu/abs/2006E&PSL.245..523
H). doi:10.1016/j.epsl.2006.03.009 (https://doi.org/10.1016%2Fj.epsl.2006.03.009).
Retrieved 6 April 2023.
128. Gu, Guangsheng; Dickens, G.R.; Bhatnagar, G.; Colwell, F.S.; Hirasaki, G.J.; Chapman,
W.G. (2011). "Abundant Early Palaeogene marine gas hydrates despite warm deep-ocean
temperatures". Nature Geoscience. 4 (12): 848–851. Bibcode:2011NatGe...4..848G (https://u
i.adsabs.harvard.edu/abs/2011NatGe...4..848G). doi:10.1038/ngeo1301 (https://doi.org/10.1
038%2Fngeo1301).
129. Fox-Kemper, B.; Hewitt, H.T.; Xiao, C.; Aðalgeirsdóttir, G.; Drijfhout, S.S.; Edwards, T.L.;
Golledge, N.R.; Hemer, M.; Kopp, R.E.; Krinner, G.; Mix, A. (2021). Masson-Delmotte, V.;
Zhai, P.; Pirani, A.; Connors, S.L.; Péan, C.; Berger, S.; Caud, N.; Chen, Y.; Goldfarb, L.
(eds.). "Chapter 5: Global Carbon and other Biogeochemical Cycles and Feedbacks" (http
s://www.ipcc.ch/report/ar6/wg1/downloads/report/IPCC_AR6_WGI_Full_Report.pdf) (PDF).
Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the
Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge
University Press, Cambridge, UK and New York, NY, USA: 80.
doi:10.1017/9781009157896.011 (https://doi.org/10.1017%2F9781009157896.011)
(inactive 31 December 2022).
130. Pagani, Mark; Caldeira, K.; Archer, D.; Zachos, J.C. (8 December 2006). "An Ancient Carbon
Mystery" (https://semanticscholar.org/paper/34b355256a88a9f492290efdbd2f4c73532b647
1). Science. 314 (5805): 1556–7. doi:10.1126/science.1136110 (https://doi.org/10.1126%2F
science.1136110). PMID 17158314 (https://pubmed.ncbi.nlm.nih.gov/17158314).
S2CID 128375931 (https://api.semanticscholar.org/CorpusID:128375931).
131. Gehler; et al. (2015). "Temperature and atmospheric CO2 concentration estimates through
the PETM using triple oxygen isotope analysis of mammalian bioapatite" (https://www.ncbi.n
lm.nih.gov/pmc/articles/PMC4948332). Proceedings of the National Academy of Sciences of
the United States of America. 113 (8): 7739–7744. Bibcode:2016PNAS..113.7739G (https://u
i.adsabs.harvard.edu/abs/2016PNAS..113.7739G). doi:10.1073/pnas.1518116113 (https://d
oi.org/10.1073%2Fpnas.1518116113). PMC 4948332 (https://www.ncbi.nlm.nih.gov/pmc/arti
cles/PMC4948332). PMID 27354522 (https://pubmed.ncbi.nlm.nih.gov/27354522).
132. Frieling, J.; Peterse, F.; Lunt, D. J.; Bohaty, S. M.; Sinninghe Damsté, J. S.; Reichart, G. -J.;
Sluijs, A. (18 March 2019). "Widespread Warming Before and Elevated Barium Burial During
the Paleocene-Eocene Thermal Maximum: Evidence for Methane Hydrate Release?" (http
s://www.ncbi.nlm.nih.gov/pmc/articles/PMC6582550). Paleoceanography and
Paleoclimatology. 34 (4): 546–566. Bibcode:2019PaPa...34..546F (https://ui.adsabs.harvard.
edu/abs/2019PaPa...34..546F). doi:10.1029/2018PA003425 (https://doi.org/10.1029%2F201
8PA003425). PMC 6582550 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6582550).
PMID 31245790 (https://pubmed.ncbi.nlm.nih.gov/31245790).
133. Babila, Tali L.; Penman, Donald E.; Standish, Christopher D.; Doubrawa, Monika; Bralower,
Timothy J.; Robinson, Marci M.; Self-Trail, Jean M.; Speijer, Robert P.; Stassen, Peter;
Foster, Gavin L.; Zachos, James C. (16 March 2022). "Surface ocean warming and
acidification driven by rapid carbon release precedes Paleocene-Eocene Thermal
Maximum" (https://lirias.kuleuven.be/handle/20.500.12942/694229). Science Advances. 8
(11): eabg1025. Bibcode:2022SciA....8G1025B (https://ui.adsabs.harvard.edu/abs/2022Sci
A....8G1025B). doi:10.1126/sciadv.abg1025 (https://doi.org/10.1126%2Fsciadv.abg1025).
PMC 8926327 (https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8926327). PMID 35294237
(https://pubmed.ncbi.nlm.nih.gov/35294237). S2CID 247498325 (https://api.semanticscholar.
org/CorpusID:247498325).
134. Katz, M.E.; Cramer, B.S.; Mountain, G.S.; Katz, S.; Miller, K.G. (2001). "Uncorking the bottle:
What triggered the Paleocene/Eocene thermal maximum methane release" (https://web.arch
ive.org/web/20080513053145/http://geology.rutgers.edu/pdf/Katz.etal.2001.pdf) (PDF).
Paleoceanography and Paleoclimatology. 16 (6): 667. Bibcode:2001PalOc..16..549K (http
s://ui.adsabs.harvard.edu/abs/2001PalOc..16..549K). CiteSeerX 10.1.1.173.2201 (https://cite
seerx.ist.psu.edu/viewdoc/summary?doi=10.1.1.173.2201). doi:10.1029/2000PA000615 (htt
ps://doi.org/10.1029%2F2000PA000615). Archived from the original (http://geology.rutgers.e
du/pdf/Katz.etal.2001.pdf) (PDF) on 2008-05-13. Retrieved 2008-02-28.
135. MacDonald, Gordon J. (1990). "Role of methane clathrates in past and future climates".
Climatic Change. 16 (3): 247–281. Bibcode:1990ClCh...16..247M (https://ui.adsabs.harvard.
edu/abs/1990ClCh...16..247M). doi:10.1007/BF00144504 (https://doi.org/10.1007%2FBF00
144504). S2CID 153361540 (https://api.semanticscholar.org/CorpusID:153361540).
136. Bice, K.L.; Marotzke, J. (2002). "Could changing ocean circulation have destabilized
methane hydrate at the Paleocene/Eocene boundary" (https://eprints.soton.ac.uk/231/1/BIC
E_%2526_MAROTZKE_paper_paleoce_figures.pdf) (PDF). Paleoceanography and
Paleoclimatology. 17 (2): 1018. Bibcode:2002PalOc..17.1018B (https://ui.adsabs.harvard.ed
u/abs/2002PalOc..17.1018B). doi:10.1029/2001PA000678 (https://doi.org/10.1029%2F2001
PA000678). hdl:11858/00-001M-0000-0014-3AC0-A (https://hdl.handle.net/11858%2F00-00
1M-0000-0014-3AC0-A). Archived (https://web.archive.org/web/20120419090536/http://eprin
ts.soton.ac.uk/231/1/BICE_%26_MAROTZKE_paper_paleoce_figures.pdf) (PDF) from the
original on 2012-04-19. Retrieved 2019-09-01.
137. Tripati, Aradhna K.; Elderfield, Henry (14 February 2004). "Abrupt hydrographic changes in
the equatorial Pacific and subtropical Atlantic from foraminiferal Mg/Ca indicate greenhouse
origin for the thermal maximum at the Paleocene-Eocene Boundary" (https://agupubs.onlinel
ibrary.wiley.com/doi/10.1029/2003GC000631). Geochemistry, Geophysics, Geosystems. 5
(2): 1–11. Bibcode:2004GGG.....5.2006T (https://ui.adsabs.harvard.edu/abs/2004GGG.....5.2
006T). doi:10.1029/2003GC000631 (https://doi.org/10.1029%2F2003GC000631).
S2CID 129878181 (https://api.semanticscholar.org/CorpusID:129878181). Retrieved 8 April
2023.
138. Bice, Karen L.; Marotzke, Jochem (15 June 2001). "Numerical evidence against reversed
thermohaline circulation in the warm Paleocene/Eocene ocean" (https://agupubs.onlinelibrar
y.wiley.com/doi/10.1029/2000JC000561). Journal of Geophysical Research. 106 (C6):
11529–11542. Bibcode:2001JGR...10611529B (https://ui.adsabs.harvard.edu/abs/2001JG
R...10611529B). doi:10.1029/2000JC000561 (https://doi.org/10.1029%2F2000JC000561).
hdl:11858/00-001M-0000-0014-3AC6-D (https://hdl.handle.net/11858%2F00-001M-0000-00
14-3AC6-D). Retrieved 7 April 2023.
139. Cope, Jesse Tiner (2009). On The Sensitivity Of Ocean Circulation To Arctic Freshwater
Pulses During The Paleocene/Eocene Thermal Maximum (https://rc.library.uta.edu/uta-ir/ha
ndle/10106/2004) (Masters thesis). University of Texas Arlington. hdl:10106/2004 (https://hdl.
handle.net/10106%2F2004). Retrieved 2013-08-07.
140. Bains, S.; Norris, R.D.; Corfield, R.M.; Faul, K.L. (2000). "Termination of global warmth at the
Palaeocene/Eocene boundary through productivity feedback" (https://www.nature.com/articl
es/35025035). Nature. 407 (6801): 171–4. Bibcode:2000Natur.407..171B (https://ui.adsabs.h
arvard.edu/abs/2000Natur.407..171B). doi:10.1038/35025035 (https://doi.org/10.1038%2F35
025035). PMID 11001051 (https://pubmed.ncbi.nlm.nih.gov/11001051). S2CID 4419536 (htt
ps://api.semanticscholar.org/CorpusID:4419536). Retrieved 6 April 2023.
141. Dickens, G. R.; Fewless, T.; Thomas, E.; Bralower, T. J. (2003). "Excess barite accumulation
during the Paleocene-Eocene thermal Maximum: Massive input of dissolved barium from
seafloor gas hydrate reservoirs" (https://semanticscholar.org/paper/cdae139ad9bbd2111143
b423e9afc0d276870ea5). Special Paper 369: Causes and consequences of globally warm
climates in the early Paleogene. Vol. 369. p. 11. doi:10.1130/0-8137-2369-8.11 (https://doi.or
g/10.1130%2F0-8137-2369-8.11). ISBN 978-0-8137-2369-3. S2CID 132420227 (https://api.
semanticscholar.org/CorpusID:132420227).
142. Stokke, Ella W.; Jones, Morgan T.; Tierney, Jessica E.; Svensen, Henrik H.; Whiteside,
Jessica H. (15 August 2020). "Temperature changes across the Paleocene-Eocene Thermal
Maximum – a new high-resolution TEX86 temperature record from the Eastern North Sea
Basin" (https://www.sciencedirect.com/science/article/pii/S0012821X20303320). Earth and
Planetary Science Letters. 544: 116388. Bibcode:2020E&PSL.54416388S (https://ui.adsab
s.harvard.edu/abs/2020E&PSL.54416388S). doi:10.1016/j.epsl.2020.116388 (https://doi.or
g/10.1016%2Fj.epsl.2020.116388). S2CID 225387296 (https://api.semanticscholar.org/Corp
usID:225387296). Retrieved 6 April 2023.
143. Kiehl, Jeffrey T.; Shields, Christine A.; Snyder, Mark A.; Zachos, James C.; Rothstein,
Mathew (3 September 2018). "Greenhouse- and orbital-forced climate extremes during the
early Eocene" (https://royalsocietypublishing.org/doi/10.1098/rsta.2017.0085). Philosophical
Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences.
376: 1–24. doi:10.1098/rsta.2017.0085 (https://doi.org/10.1098%2Frsta.2017.0085).
Retrieved 1 May 2023.
144. "PETM Weirdness" (http://www.realclimate.org/index.php/archives/2009/08/petm-weirdnes
s/). RealClimate. 2009. Archived (https://web.archive.org/web/20160212041029/http://www.r
ealclimate.org/index.php/archives/2009/08/petm-weirdness) from the original on 2016-02-
12. Retrieved 2016-02-03.
145. Ying Cui; Lee R. Kump; Andy J. Ridgwell; Adam J. Charles; Christopher K. Junium; Aaron F.
Diefendorf; Katherine H. Freeman; Nathan M. Urban & Ian C. Harding (2011). "Slow release
of fossil carbon during the Palaeocene–Eocene Thermal Maximum". Nature Geoscience. 4
(7): 481–485. Bibcode:2011NatGe...4..481C (https://ui.adsabs.harvard.edu/abs/2011NatG
e...4..481C). doi:10.1038/ngeo1179 (https://doi.org/10.1038%2Fngeo1179).
146. Ruppel and Kessler (2017). "The interaction of climate change and methane hydrates" (http
s://doi.org/10.1002%2F2016RG000534). Reviews of Geophysics. 55 (1): 126–168.
Bibcode:2017RvGeo..55..126R (https://ui.adsabs.harvard.edu/abs/2017RvGeo..55..126R).
doi:10.1002/2016RG000534 (https://doi.org/10.1002%2F2016RG000534).
147. Keller, Gerta; Mateo, Paula; Punekar, Jahnavi; Khozyem, Hassan; Gertsch, Brian;
Spangenberg, Jorge E.; Bitchong, Andre Mbabi; Adatte, Thierry (April 2018). "Environmental
changes during the Cretaceous-Paleogene mass extinction and Paleocene-Eocene
Thermal Maximum: Implications for the Anthropocene" (https://www.sciencedirect.com/scien
ce/article/abs/pii/S1342937X17303702). Gondwana Research. 56: 69–89.
doi:10.1016/j.gr.2017.12.002 (https://doi.org/10.1016%2Fj.gr.2017.12.002). Retrieved
11 June 2023.
148. "High-fidelity record of Earth's climate history puts current changes in context" (https://phys.o
rg/news/2020-09-high-fidelity-earth-climate-history-current.html). phys.org. Retrieved
26 September 2021.
149. "Ancient Climate Events: Paleocene Eocene Thermal Maximum | EARTH 103: Earth in the
Future" (https://www.e-education.psu.edu/earth103/node/639). www.e-education.psu.edu.
Retrieved 26 September 2021.
150. "Scientists draw new connections between climate change and warming oceans" (https://ph
ys.org/news/2018-08-scientists-climate-oceans.html). phys.org. University of Toronto.
Retrieved 26 September 2021.
151. Yao, Weiqi; Paytan, Adina; Wortmann, Ulrich G. (24 August 2018). "Large-scale ocean
deoxygenation during the Paleocene-Eocene Thermal Maximum" (https://doi.org/10.1126%
2Fscience.aar8658). Science. 361 (6404): 804–806. Bibcode:2018Sci...361..804Y (https://ui.
adsabs.harvard.edu/abs/2018Sci...361..804Y). doi:10.1126/science.aar8658 (https://doi.org/
10.1126%2Fscience.aar8658). PMID 30026315 (https://pubmed.ncbi.nlm.nih.gov/3002631
5).
152. " 'Tipping points' in Earth's system triggered rapid climate change 55 million years ago,
research shows" (https://phys.org/news/2021-08-earth-triggered-rapid-climate-million.html).
phys.org. Retrieved 21 September 2021.
153. Kaya, Mustafa Y.; Dupont-Nivet, Guillaume; Frieling, Joost; Fioroni, Chiara; Rohrmann,
Alexander; Altıner, Sevinç Özkan; Vardar, Ezgi; Tanyaş, Hakan; Mamtimin, Mehmut; Zhaojie,
Guo (31 May 2022). "The Eurasian epicontinental sea was an important carbon sink during
the Palaeocene-Eocene thermal maximum" (https://doi.org/10.1038%2Fs43247-022-00451-
4). Communications Earth & Environment. 3 (124): 124. Bibcode:2022ComEE...3..124K (http
s://ui.adsabs.harvard.edu/abs/2022ComEE...3..124K). doi:10.1038/s43247-022-00451-4 (htt
ps://doi.org/10.1038%2Fs43247-022-00451-4). S2CID 249184616 (https://api.semanticschol
ar.org/CorpusID:249184616).
154. Tierney, Jessica E.; Zhu, Jiang; Li, Mingsong; Ridgwell, Andy; Hakim, Gregory J.; Poulsen,
Christopher J.; Whiteford, Ross D. M.; Rae, James W. B.; Kump, Lee R. (10 October 2022).
"Spatial patterns of climate change across the Paleocene–Eocene Thermal Maximum" (http
s://www.pnas.org/doi/10.1073/pnas.2205326119). Proceedings of the National Academy of
Sciences of the United States of America. 119 (42): 1–7. doi:10.1073/pnas.2205326119 (http
s://doi.org/10.1073%2Fpnas.2205326119). Retrieved 26 April 2023.

Further reading
Jardine, Phil (2011). "Paleocene–Eocene Thermal Maximum" (http://www.palaeontologyonli
ne.com/articles/2011/the-paleocene-eocene-thermal-maximum/). Palaeontology Online. 1
(5): 1–7.

External links
BBC Radio 4, In Our Time, The Paleocene–Eocene Thermal Maximum, 16 March 2017 (http
s://www.bbc.co.uk/programmes/b08hpmmf)
Global Warming 56 Million Years Ago: What it Means for Us (https://www.youtube.com/watc
h?v=81Zb0pJa3Hg) (Video)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Paleocene–


Eocene_Thermal_Maximum&oldid=1163242889"

You might also like