ArticlePDF Available

Coupling between Grand cycles and Events in Earth’s climate during the past 115 million years

Springer Nature
Scientific Reports
Authors:

Abstract and Figures

Geological sediment archives document a rich periodic series of astronomically driven climate, but record also abrupt, severe climatic changes called events, the multi-Myr boundary conditions of which have generally been ascribed to acyclic processes from Earth’s interior dynamics. These events have rarely been considered together within extended time series for potential correlation with long-term (multi-million year, Myr) cycling. Here I show a coupling between events and multi-Myr cycles in a temperature and ice-volume climatic proxy of the geological past 115 Myr. I use Cenozoic through middle Cretaceous climatic variations, as recorded in benthic foraminifera δ18O, to highlight prominent ~9 and ~36 Myr cyclicities. These cyclicities were previously attributed either to astronomical or tectonic variations. In particular, I point out that most of the well-known events during the past 115 Myr geological interval occur during extremes in the ~9 and ~36 Myr cycling. One exception is the early Cenozoic hyperthermal events including the salient Paleocene-Eocene Thermal Maximum (~56 Ma), which do not match extremes in long-period cyclicities, but to inflection point of these cycles. Specific focus on climatic events, as inferred from δ18O proxy, suggest that some “events”, marked by gradual trends within the ~9 and ~36 Myr cycle extremes, would principally be paced by long-term cycling, while “events”, recorded as abrupt δ18O changes nearby cycle extremes, would be rather induced by acyclic processes. The connection between cyclic and acyclic processes, as triggers or feedbacks, is very likely. Such link between cycling and events in Earth’s past climate provides insight into celestial dynamics governing perturbations in Earth’s surface systems, but also the potential connection between external and Earth’s interior processes.
Time-series analysis of benthic foraminifera δ 18 O data of the past 115 Ma (see Methods). (a) The raw data with two smoothing fits: a third-order polynomial fit and a 14% weighted-average fit. (b) Detrended δ 18 O data (third-order polynomial fit shown in 'a' is removed) with two smoothing fits: a 14% weightedaverage fit to highlight the 36 Myr cycles, and a 4% weighted-average fit to highlight both the ~9 and ~36 Myr cycles. The well known climatic events are shown: IA: Pliocene-Pleistocenbe Ice Age, MMCO: Mid-Miocene Climatic Optimum, Mi-1 and Oi-1: Oligocene-Miocene and Eocene-Oligocene major glacial events, MECO: Mid-Eocene Climatic Optimum, EECO: Early Eocene Climatic Optimum, PETM: Paleocene-Eocene Thermal Maximum, K/T: Cretaceous/Tertiary boundary, EMaC: Early Maastrichtian Cooling episode, OAE2 and OAE1b: Oceanic Anoxic Events. (c) Bandpass filtering of δ 18 O data: 36 Myr cycle band (0.03 ±0.01 cycles/ Myr) and 9 Myr cycle band (0.11 ±0.03 cycles/Myr). Ages, P: Pleistocene, Pl.: Pliocene, Oligoc.: Oligocene, Paleoc.: Paleocene, Ma.: Maastrichtian, Campan.: Campanian, S: Santonian, Co: Coniacian, Tu.: Turonian, Cen.: Cenomanian, A.: Aptian. Vertical shaded bars are extrema (minima in red and maxima in blue) of the ~36 Myr δ 18 O cycles, which coincide with the most known thermal and cooling episodes. Also indicated climatic events (vertical grey lines), some of them matching extrema in the ~9 Myr cycles. Ox1 to Ox12 are the interpreted ~9 Myr δ 18 O cycles. Ox1, Ox2, Ox3,… to designate ~9 Myr δ 18 O oscillations 1, 2, 3,… Ox for oxygen (δ 18 O), and increasing numbers indicate the older oscillations. This is used by correlation with Cb1, Cb2, Cb3,… that designate the ~9 Myr δ 13 C oscillations 1, 2, 3,… (Supplementary Figure S1) as in Boulila et al. 6 , Cb. for carbon (δ 13 C), and increasing numbers indicate older cycles. Ox1 to Ox4 indicated by asterisks, correlate with their equivalents Cb1 to Cb4 in δ 13 C record (ref. 6 , Supplementary Figure S1). Ox7 and Ox8 indicated by question marks include one oscillation, equivalent to two ~9 Myr oscillations in the δ 13 C record (see text for discussion, and Supplementary Figure S2), Ox12, shown with a question mark, is a poorly constrained cycle, due to lowresolution data within this time interval.
… 
This content is subject to copyright. Terms and conditions apply.
1
SCIEntIfIC REpoRTS | (2019) 9:327 | DOI:10.1038/s41598-018-36509-7
www.nature.com/scientificreports
Coupling between Grand cycles and
Events in Earth’s climate during the
past 115 million years
Slah Boulila1,2
Geological sediment archives document a rich periodic series of astronomically driven climate, but
record also abrupt, severe climatic changes called events, the multi-Myr boundary conditions of which
have generally been ascribed to acyclic processes from Earth’s interior dynamics. These events have
rarely been considered together within extended time series for potential correlation with long-term
(multi-million year, Myr) cycling. Here I show a coupling between events and multi-Myr cycles in a
temperature and ice-volume climatic proxy of the geological past 115 Myr. I use Cenozoic through
middle Cretaceous climatic variations, as recorded in benthic foraminifera δ18O, to highlight prominent
~9 and ~36 Myr cyclicities. These cyclicities were previously attributed either to astronomical or tectonic
variations. In particular, I point out that most of the well-known events during the past 115 Myr
geological interval occur during extremes in the ~9 and ~36 Myr cycling. One exception is the early
Cenozoic hyperthermal events including the salient Paleocene-Eocene Thermal Maximum (~56 Ma),
which do not match extremes in long-period cyclicities, but to inection point of these cycles. Specic
focus on climatic events, as inferred from δ18O proxy, suggest that some “events”, marked by gradual
trends within the ~9 and ~36 Myr cycle extremes, would principally be paced by long-term cycling, while
“events”, recorded as abrupt δ18O changes nearby cycle extremes, would be rather induced by acyclic
processes. The connection between cyclic and acyclic processes, as triggers or feedbacks, is very likely.
Such link between cycling and events in Earth’s past climate provides insight into celestial dynamics
governing perturbations in Earth’s surface systems, but also the potential connection between external
and Earth’s interior processes.
Earth’s past climate has varied quasi-periodically from hundred to billion years, and these variations were driven
by celestial dynamics over a large frequency band13. In particular, Earth’s climate during the Cenozoic-Late
Cretaceous (past 115 million years, Ma) experienced long-term (multi-million year, multi-Myr) changes punc-
tuated by severe, short-lived events48. e multi-Myr climatic variations have oen been considered irregular
(acyclic) in nature on the very long term9. is has led to the hypothesis that these changes were mainly driven
by Earth’s interior dynamics and plate tectonic motions assuming that these processes are acyclic in nature9,
although more recent studies point to the cyclic behavior of large plate tectonic motions1016. e climatic events
have been intensively investigated in terms of their heavy impacts and consequences on Earth’s supercial sys-
tems, in particular the resulting profound modications in ocean and terrestrial environments, expressed as per-
turbations in the sedimentary biogeochemical cycles4,5. Delineating the cyclic versus acyclic nature of multi-Myr
climatic variations and their potential link with the climatic events is of paramount importance in deciphering the
origin of events and thus their causal mechanisms. However, the study of multi-Myr climatic variations requires
the acquisition of high-resolution data covering several million years, hence hampering the characterization of
long-term cycling in previous studies1719. New advancements in laboratory analytical techniques together with
large collaborative projects allow today the availability of a highly resolved 115-Myr-long climate record from
sedimentary benthic foraminifera δ18O (see Methods). e δ18O record indicates variations in deep-sea temper-
ature as well as changes in ice volume4,7,8.
1Sorbonne Université, CNRS, Institut des Sciences de la Terre Paris, ISTeP, F-75005, Paris, France. 2ASD/IMCCE,
CNRS-UMR8028, Observatoire de Paris, PSL University, Sorbonne Université, 77 Avenue Denfert-Rochereau,
75014, Paris, France. Correspondence and requests for materials should be addressed to S.B. (email: slah.boulila@
sorbonne-universite.fr)
Received: 23 July 2018
Accepted: 20 November 2018
Published: xx xx xxxx
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
2
SCIEntIfIC REpoRTS | (2019) 9:327 | DOI:10.1038/s41598-018-36509-7
Results
Time-series analysis of δ18O data shows two prominent cyclicities of ~9 and ~36 Myr (Figs1 and 2). Other
low-frequency cyclicities are also present, in particular, 1.3 Myr, 1.6 Myr, a triplet peaks of 2.3, 2.5 and 2.8 Myr,
and a 4.7 Myr peak (Fig.2a), which are close to Milankovitch astronomical periods (Fig.2b). e 1.3 Myr corre-
sponds to the 1.2 Myr obliquity modulation term. e 1.6 Myr matches the eccentricity term. e triplet corre-
sponds to combined eccentricity and obliquity terms of 2, 2.3 and 2.6 Myr (Fig.3 and Supplementary FigureS3).
Finally, the 4.7 Myr is close to the 4.6 Myr eccentricity peak. Even though spectrally well detected and strikingly
close to Milankovitch frequency band (Fig.3) the periods from 1 to 5 Myr could be altered by the stacking pro-
cess of the data. Data from single sites document with high delity some of these periods2032. In contrast, longer
periods (>5 Myr) are notably preserved in the compiled δ18O data. In the present study, I focus on the prominent
cyclicities of ~9 and ~36 Myr, and their possible link with the geologically well known climatic events. Another
potential low-frequency (16 to 18 Myr) cycle could also be seen in the δ18O power spectra (Fig.3), but will not be
discussed below.
Figure 1. Time-series analysis of benthic foraminifera δ18O data of the past 115 Ma (see Methods). (a) e
raw data with two smoothing ts: a third-order polynomial t and a 14% weighted-average t. (b) Detrended
δ18O data (third-order polynomial t shown in ‘a’ is removed) with two smoothing ts: a 14% weighted-
average t to highlight the 36 Myr cycles, and a 4% weighted-average t to highlight both the ~9 and ~36 Myr
cycles. e well known climatic events are shown: IA: Pliocene-Pleistocenbe Ice Age, MMCO: Mid-Miocene
Climatic Optimum, Mi-1 and Oi-1: Oligocene-Miocene and Eocene-Oligocene major glacial events, MECO:
Mid-Eocene Climatic Optimum, EECO: Early Eocene Climatic Optimum, PETM: Paleocene-Eocene ermal
Maximum, K/T: Cretaceous/Tertiary boundary, EMaC: Early Maastrichtian Cooling episode, OAE2 and
OAE1b: Oceanic Anoxic Events. (c) Bandpass ltering of δ18O data: 36 Myr cycle band (0.03 ±0.01 cycles/
Myr) and 9 Myr cycle band (0.11 ±0.03 cycles/Myr). Ages, P: Pleistocene, Pl.: Pliocene, Oligoc.: Oligocene,
Paleoc.: Paleocene, Ma.: Maastrichtian, Campan.: Campanian, S: Santonian, Co: Coniacian, Tu.: Turonian, Cen.:
Cenomanian, A.: Aptian. Vertical shaded bars are extrema (minima in red and maxima in blue) of the ~36 Myr
δ18O cycles, which coincide with the most known thermal and cooling episodes. Also indicated climatic events
(vertical grey lines), some of them matching extrema in the ~9 Myr cycles. Ox1 to Ox12 are the interpreted
~9 Myr δ18O cycles. Ox1, Ox2, Ox3, to designate ~9 Myr δ18O oscillations 1, 2, 3, Ox for oxygen (δ18O),
and increasing numbers indicate the older oscillations. is is used by correlation with Cb1, Cb2, Cb3, that
designate the ~9 Myr δ13C oscillations 1, 2, 3, (Supplementary FigureS1) as in Boulila et al.6, Cb. for carbon
(δ13C), and increasing numbers indicate older cycles. Ox1 to Ox4 indicated by asterisks, correlate with their
equivalents Cb1 to Cb4 in δ13C record (ref.6, Supplementary FigureS1). Ox7 and Ox8 indicated by question
marks include one oscillation, equivalent to two ~9 Myr oscillations in the δ13C record (see text for discussion,
and Supplementary FigureS2), Ox12, shown with a question mark, is a poorly constrained cycle, due to low-
resolution data within this time interval.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
3
SCIEntIfIC REpoRTS | (2019) 9:327 | DOI:10.1038/s41598-018-36509-7
Discussion
e ~9 Myr cyclicity was previously detected in carbon-cycle variations6. Comparison of δ18O and δ13C data
shows a coupling between climate and carbon cycle at the ~9 Myr cycle band, especially during icehouse, i.e.
the past 34 Ma (Supplementary FigureS1). A strong decoupling between them is remarkably noted within the
interval from 50 to 65 Ma as follows. While the δ13C document the two strongest ~9 Myr cycles Cb7 and Cb8 of
the early Cenozoic6, the δ18O has exceptionally only one oscillation (Fig.3d, see also Supplementary FigureS2).
e origin of ~9 Myr cyclicity has previously been attributed to the modulation of 2.4 Myr eccentricity cycle
band6. Here I investigate additional statistical tests to support the modulation at the 2.4 Myr cycle band, but with
possible contribution from both eccentricity and obliquity terms (Fig.2b, and Supplementary FiguresS3 and S4).
Prominent ~9 Myr geological oscillations have been detected in carbon-cycle and sedimentological proxies of
Cenozoic and Mesozoic strata6,3335.
Interestingly, I note four ~9 Myr oscillations grouped in each ~36 Myr cycle (Fig.1b,c), suggesting a relation-
ship between these two climatic cyclicities. Previous studies ascribed the ~36 Myr climatic variability to another
dimension of astronomical forcing36,37. In particular, the ~36 Myr period has been attributed to the vertical pas-
sage of the Solar System through the Galactic midplane that modulates the ux of galactic cosmic rays (GCR) on
Earth36. Despite our limited knowledge on the gravitational potential of the Galaxy, there is some consensus that
the Solar System vertically oscillates across the Galactic midplane, with a half-period of about 36 Myr16,3840. is
eective 36 Myr half-period corresponds to the motion of the Solar System when it moves down- and upwards
the galactic midplane. is would induce signicant change in GCR ux on Earth, especially when crossing the
galactic midplane. is would favor the formation of cloud layer from the incidental GCR ux on Earth, hence
resulting in variations of Earth’s albedo and the consequent climate change16,36,41,42. Yet, the impact of cosmic ray
on climate remains a controversial subject43. Another possible hypothesis is that Earth’s interior dynamics may
resonate every 36 Myr inducing global climate change, although recent studies pointed to dierent periodicities of
25 to 50 Myr13,14 and 56 Myr15. A combined eect from the aforcited mechanisms on climate is presumably44. e
most intriguing result in time-series analysis of δ18O climatic proxy is the record of both ~9 and ~36 Myr oscilla-
tions, sharing very likely the same forcing process. While the ~9 Myr periodicity may represent a Milankovitch
origin, the ~36 periodicity could be forced by solar system vertical motion, that could in turn modulate Earth’s
incident (Milankovitch) insolation, or may correspond to a not resolved Milankovitch band. Although I draw
Figure 2. 2π-MTM power spectra of benthic foraminifera δ18O data of the interval from 0 to 115 Ma (see
Methods) and of its time-equivalent astronomical variations. (a) Upper panel: Spectrum of the detrended δ18O
data (50% weighted average of the series is removed), lower panel: Spectrum of 1x zero padded δ18O aer the
detrend (50% weighted average of the series is removed). Signicant spectral periods are shown across vertical
grey bars. (b) Upper panel: Spectrum of lowpass ltered (>1 Myr), composite astronomical time series of the
interval 0 to 115 Ma (see Methods), the power axis is a logarithmic scale for comparison with spectra in ‘a’,
lower panel: Spectrum of the same astronomical time series (power in linear scale) but over the interval from
0 to 146 Ma (Cenozoic and Cretaceous time). e most prominent astronomical periods and origins of some
of them are shown across vertical grey bars, periods indicated by asterisk may represent minor components.
Note that the ~9 Myr δ18O cyclicity originates from the interfering astronomical terms 2 and 2.6 Myr (see
Supplementary FigureS3). e 16.3 Myr δ18O period may correspond to a harmonic or may originate from the
couple interfering terms 2 Myr vs 2.3 Myr and 2.3 Myr vs 2.6 Myr.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
4
SCIEntIfIC REpoRTS | (2019) 9:327 | DOI:10.1038/s41598-018-36509-7
attention to climate and astronomy linkage at the scale of ~9 and ~36 Myr periods, we must caution that the eect
of tectonics on climate at this timescale remains well plausible16.
Along climate-proxy variations in the Cenozoic and Mesozoic eras (past 0–250 Ma) were recognized and
widely studied a number of climatic events4,5,8. e studied interval includes some of them (Fig.1b). Features
and nature of Earth’s system responses to these perturbations differ from one event to another5, however,
tectonic-volcanic mechanisms were thought to be the principal, common cause for most of the events4.
Here, I show that most of Cenozoic–Cretaceous climatic events and phases (see Fig.3 for denition of ‘event’
versus ‘phase’) during the past 115 Ma fall into extremes in amplitudes of the ~9 and ~36 Myr cyclicities (Fig.1).
For instance, the so-called Oi-1 and Mi-1 glacial episodes are close to the ~9 Myr cycle extremes (Ox4,
Fig.1c). e Mid-Miocene Climatic Optimum (MMCO), which represents a warming phase, bounds ~9 Myr
Ox2 and Ox3 δ18O oscillations (see Fig.1 caption for ‘Ox’ cycle numbering). Importantly, some of the events
coincide with extremes in the ~36 Myr cycling, thus, they would be with greater magnitudes (Fig.1b). e three
cooling/glacial episodes at Pliocene-Pleistocene Ice Age (IA), Eocene-Oligocene glacial event (Oi-1) and Early
Maastrichtian Cooling phase (EMaC) correspond to maxima of the ~36 Myr δ18O cycles. e two warming phases
of Mid-Miocene Climatic Optimum (MMCO) and Early Eocene Climatic Optimum (EECO) coincide with the
~36 Myr cycle minima. Finally, the globally recognized Oceanic Anoxic Event OAE-2 occurs around a minimum
of the ~36 Myr δ18O cyclicity. In contrast, the early Cenozoic hyperthermals including the most pronounced
Paleocene-Eocene ermal Maximum (PETM) event do not match extremes in the long-term cycling. e rela-
tive variance of the ~36 Myr δ18O cyclicity is more than three times higher than that of the ~9 Myr cyclicity. is
implies that events (and phases) matching only extremes in the ~9 Myr cycles are with lesser intensities compared
to events (and phases) in relation with the ~36 Myr oscillations. However, Earth’s climate response to the forcing
processes at the ~9 and ~36 Myr cycle bands from astronomy, for example, could not be direct, because insolation
(or GCR) change at these very longer timescales should be weak (see below).
The most striking result is the match between ~9 and ~36 Myr cycle extremes and most of the
Cenozoic-Cretaceous events. is result hints at a coupling between Earth’s climate cycling and perturbations.
e ~9 and ~36 Myr cyclicities could have an astronomical and/or a tectonic origin (see above). e events have
been postulated to be caused by tectonics, volcanism, pCO2 trend, etc.
I suggest that among the events in relation with long-period cyclicities some of them may have been mainly
related to cyclic process, others may have been the result of combined eects of acyclic and cyclic processes. e
Figure 3. Expanded views of climatic events and phases within the ~36 Myr cycle extrema. e term ‘phase’
opposes to the term ‘event’: ‘event’ indicates abrupt, severe change, whereas ‘phase’ indicates gradual/progressive
trend in climate change that could reach an optimum (see below). I also used the term ‘episode’ to designate
either event or phase. (a) e gradual entry of the Earth into the Pliocene-Pleistocene ice age (IA) as expressed
in the trend (40% weighted average). Note that there is no trend during the 100-kyr-cycle dominated climate of
the past ca. 800 ka. (b) Gradual variations within the Mid-Miocene Climatic Optimum, MMCO (35% weighted
average). (c) Abrupt change within the Eocene-Oligocene transition (EOT) including the Oi-1 glacial event
(15% weighted average). (d) Gradual variations within the Early Eocene Climatic Optimum, EECO (35%
weighted average). Note that the Paleocene-Eocene ermal Maximum, PETM, corresponds to a deviation
from the long-term cycling within the EECO, and is the most abrupt, severe event in the Cenozoic era.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
5
SCIEntIfIC REpoRTS | (2019) 9:327 | DOI:10.1038/s41598-018-36509-7
IA, MMCO, EECO and possibly EMaC4,8 show gradual variations (Figs1 and 3), thus they may mainly be the
result of a cyclic forcing. e Oi-1 and OAE-24,5 show abrupt changes (Figs1 and 3), thus they may occur acy-
clically, but could also be triggered by a cyclic mechanism. Large-scale (multi-Myr) Milankovitch astronomical
forcing would induce negligible changes in insolation (or GCR) budget on Earth6, hence a nonlinear mechanism
is required to explain the match between the long-period astroclimatic cycles and events. Energy transfer process
from higher to lower frequency driving forces has previously been suggested6, but this process cannot argue the
match between the cycles and events. An astronomically forced gravitational distortion of Earth’s viscose mantle
in boundary conditions may explain the potential link between cycle extremes and events at the multi-Myr times-
cales. Astronomical control of gravitational deformation (e.g., dynamical ellipticity)45 of the mantle at boundary
conditions would result in intensication of mantle convection, and the occurrence of climate events, which have
been argued to be paced by processes from Earth’s interior dynamics5,44. Also, feedback responses of Earth’s inte-
rior dynamics to astronomically driven climate and Earth’s surface processes have equally been suggested even
at the shorter Milankovitch timescales46,47. For instance, glacial cycles have been correlated to oceanic crust pro-
duction47,48. Astroclimatically paced deglaciations would promote mantle decompression, hence increase magma
production and thus volcanic eruptions44,46.
Conclusions and Perspectives
In summary, the detection of a potential link between long-period climatic cyclicities (driven by astronomy and/
or tectonics) and events provides insights into multi-disciplinary studies involving acyclic versus cyclic nature
of Earth’s interior dynamics, astronomical forcing at multi-Myr timescales, and the possible interaction between
astronomical and tectonic forcings. e correspondence between very long climate cycle extremes and events
during the mid-Cretaceous to Cenozoic time interval suggests a highly nonlinear mechanism response, if the
events were responding to insolation, such as astronomical pacing, at boundary conditions, of gravitational per-
turbation of the Earths spin axis and shape. is may, in turn, aect the gravitational deformation of the mantle,
stimulate mantle convection and the dynamic topography, hence the occurrence of geodynamically paced cli-
matic events. Cyclic tectonic forcing at the 9 and 36 Myr cycle bands, without evoking the nonlinear astronomical
driving force, remains equally plausible.
Future studies should focus on the cyclic vs acyclic nature of Earth’s interior dynamics (tectonics, volcan-
ism,), at ten to tens of Myr, and likely on how could Earth’s interior dynamics vary or interact with climate4451,
and its potential link with the recognized climatic events. Other potential questions remain open, e.g., what would
be the paleoenvironmental implications of the strong decoupling, at the 9 Myr oscillations, between climate and
carbon cycles during the Paleocene-Eocene extreme greenhouse? Future studies should also focus on the origin
of the ~36 Myr climatic cyclicity, by exploring the hypothesis of the possible link between GCR driven climates
from vertical motion of the solar system, and insolation (Milankovitch) forced climates.
Methods
I used compiled benthic foraminifera δ13C and δ18O data from more than 40 deep-sea cores of the Deep Sea
Drilling Project and Ocean Drilling Program4,7,8. These data spanning the Cenozoic to the middle-Late
Cretaceous, i.e. the past ~115 Ma (Fig.1a), indicate global variations in climate and carbon cycle, including the
well known climatic events4. e Cenozoic compilation is from Cramer et al.7, and the Cretaceous compilation
is from Friedrich et al.8. Although the resolution of the stacked data would distort the record of high-frequency
cyclicities, the low-frequency, multi-Myr, if present would emerge in the compiled records6. Here I performed
time-series analysis on the compiled δ18O record to seek for multi-Myr cycling in climate; comparision with
long-term carbon cycling6 was also carried out (Supplementary FigureS1).
I applied spectral analysis to the δ18O data (Fig.2a) and to long-period (>1 Myr) astronomical variations
(Fig.2b). I generated long-term astronomical variations by extracting amplitude maxima in short eccentricity
and amplitude maxima in high-frequency obliquity time series52,53, then summing them aer standardization,
nally I lowpass ltered periods longer than 1 Myr. e astronomical variations from amplitude maxima of the
signals are susceptible to detect lower frequencies, as the conventional amplitude modulation technique. I applied
this procedure to the interval from 0 to 115 Ma (Fig.2b, upper spectrum), and to the interval from 0 to 146 Ma
(Cenozoic and Cretaceous time, Fig.2b, lower spectrum). For spectral analysis, I used the multi-taper method
(MTM) and the robust red noise as implemented in the ‘astrochron’ R packages54. MTM spectral analysis was
conducted, with three 2π tapers (Fig.2), on the linearly resampled (5 kyr) δ18O data. To precisely determine the
long periods of 9 and 36 Myr I additionally applied 1x zero padding to the δ18O time series prior spectral analysis
(Fig.2a). I carried out smoothing and tting long-term δ18O variations using both polynomial method and the
Lowess weighted average method (Fig.1). For the detailed study of δ18O record (expanded views in Fig.3), I have
rather performed only the weighted average method on the uneven spaced raw data.
Bandpass ltering was conducted using conjointly the Taner lter, which is characterized by a steep stop-
band, and the Gaussian lter which has a gentle stopband. Lowpass ltering was conducted using the Taner lter.
Filtering δ18O signal was applied to the 5 kyr linearly resampled (5 kyr) data, while Lowess weighted average
smoothing was applied to the raw data. Finally, in order to highlight the amplitude modulation of the 2.4 Myr
cycle band by the ~9 Myr cycles in the astronomical variations, I used evolutive harmonic analysis (EHA) with a
function computing a running periodogram of a uniformly sampled time series using FFTs of zero-padded seg-
ments, and normalized to the highest amplitudes (Supplementary Figure 3).
References
1. Mitchel, J. J. An overview of climatic variability and its causal mechanisms. Quat. Res. 6, 481–493 (1976).
2. Hays, J. D., Imbrie, J. & Schacleton, N. J. Variations in the Earth’s orbit: Pacemaer of the ice ages. Science 194, 1121–1132 (1976).
3. Pelletier, J. D. Analysis and Modeling of the Natural Variability of Climate. Journal of Climate 10, 1331–1342 (1997).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
6
SCIEntIfIC REpoRTS | (2019) 9:327 | DOI:10.1038/s41598-018-36509-7
4. Zachos, J. C., Pagani, M., Sloan, L., omas, E. & Billups, . Trends, hythms, and Aberrations in Global Climate 65 Ma to Present.
Science 292, 686–693 (2001).
5. Jenyns, H. C. Geochemistry of oceanic anoxic events. Geochem. Geophys. Geosyst. 11, Q03004, https://doi.
org/10.1029/2009GC002788 (2010).
6. Boulila, S., Galbrun, B., Lasar, J. & Pälie, H. A ~9 myr cycle in Cenozoic δ13C record and long-term orbital eccentricity modulation:
Is there a lin? Earth Planet. Sci. Lett. 317–318, 273–281 (2012).
7. Cramer, B. S., Toggweiler, J. ., Wright, J. D., atz, M. E. & Miller, . G. Ocean overturning since the Late Cretaceous: Inferences
from a new benthic foraminiferal isotope compilation. Paleoceanography 24, PA4216, https://doi.org/10.1029/2008PA001683
(2009).
8. Friedrich, O., Norris, . D. & Erbacher, J. Evolution of middle to Late Cretaceous oceans—A 55 m.y. record of Earth’s temperature
and carbon cycle. Geology 40, 107–110 (2012).
9. Vail, P. . et al. Seismic stratigraphy and global changes of sea level. In: C.E. Payton, C.E. Ed., Seismic Stratigraphy – Applications to
Hydrocarbon Exploration. Mem. AAPG 26, 49–212 (1977).
10. ampino, M. . & Caldeira, . Periodic impact cratering and extinction events over the last 260 million years. MNRAS 454,
3480–3484 (2015).
11.  ampino, M. . & Caldeira, . Major episodes of geologic change: correlations, time structure and possible causes. Earth Planet. Sci.
Lett. 114, 215–227 (1993).
12. aiho, . & Saito, S. Oceanic crust production and climate during the last 100 Ma. Terra Nova 6, 376–384 (1994).
13. Cogné, J.-P. & Humler, E. Trends and rhythms in global seafloor generation rate. Geochem. Geophys. Geosyst. 7, https://doi.
org/10.1029/2005GC001148 (2006).
14. DeC elles, P. G., Ducea, M. N., app, P. & Zandt, G. Cyclicity in Cordilleran orogenic systems. Nature Geosciences 2, 251–257 (2009).
15. Meyers, S. . & Peters, S. E. A 56 million year rhythm in North American sedimentation during the Phanerozoic. Earth Planet. Sci.
Lett. 303, 174–180 (2011).
16. B oulila, S., Lasar, J., Haq, B. U., Galbrun, B. & Hara, N. Long-term cyclicities in Phanerozoic sea-level sedimentary record and their
potential drivers. Global and Planetary Change 165, 128–136 (2018).
17. Shacleton, N. J. & Imbrie e δ18O spectrum of oceanic deep water over a ve-decade band. Climatic Change 16, 217–230 (1990).
18. aup, D. M. & Seposi, J. J. Jr. 1984. Periodicity of extinctions in the geologic past. Proc. Natl. Acad. Sci. USA 81, 801–805 (1984).
19. ampino, M. . & Stothers, . B. Geological rhythms and cometary impacts. Science 226, 142–143 (1984).
20. Beaufort, L. Climatic importance of the modulation of the 100 yr cycle inferred from 16 m.y. long Miocene records.
Paleoceanography 9, 821–834 (1994).
21. Lourens, L. J. & Hilgen, F. J. Long-periodic variations in the earth’s obliquity and their relation to third-order eustatic cycles and Late
Neogene glaciations. Quat. Int. 40, 43–52 (1997).
22. Olsen, P. E. & ent, D. V. Long-term Milanovitch cycles from the Late Triassic and Early Jurassic of eastern North America and
their implications for the calibration of the Early Mesozoic timescale and the long term behavior of the planets. Royal Soc. (London).
Phil. Trans. Ser. A. 357, 1761–1788 (1999).
23. Wade, B. S. & Pälie, H. Oligocene climate dynamics. Paleoceanography 19, PA4019, https://doi.org/10.1029/2004PA001042 (2004).
24. Pälie, H. et al. e Heartbeat of the Oligocene Climate System. Science 314, 1894–1898 (2006).
25. van Dam, J. A. et al. Long-period astronomical forcing of mammal turnover. Nature 443, 687–691 (2006).
26. Lirer, F. et al. Astronomically forced teleconnection between Paratethyan and Mediterranean sediments during the Middle and Late
Miocene. Palaeogeogr. Palaeoclimatol. Palaeoecol. 275, 1–13 (2009).
27. Abels, H. A., Abdul Aziz, H., rijgsman, W., Smeets, S. J. B. & Hilgen, F. J. Long-period eccentricity control on sedimentary
sequences in the continental Madrid Basin (middle Miocene, Spain). Earth Planet. Sci. Lett. 289, 220–231 (2010).
28. B oulila, S. et al. On the origin of Cenozoic and Mesozoic third-order eustatic sequences. Earth-Sci. Rev. 109, 94–112 (2011).
29. Liebrand, D. et al . Cyclostratigraphy and eccentricity tuning of the early Oligocene through early Miocene (30.1–17.1Ma): Cibicides
mundulusstable oxygen and carbon isotope records from Walvis idge Site 12. Earth Planet. Sci. Lett. 450, 392–405 (2016).
30. Liebrand, D. et al. Evolution of the early Antarctic ice ages. Proc Natl Acad Sci USA 114, 3867–3872 (2017).
31. Valero, L., Garcés, M., Cabrera, L., Costa, E. & Sáez, A. 20 Myr of eccentricity paced lacustrine cycles in the Cenozoic Ebro Basin.
Earth Planet. Sci. Lett. 408, 183–193 (2014).
32. Valero, L., Cabrera, L., Sáez, A. & Garcés, M. Long-period astronomically-forced terrestrial carbon sins. Earth Planet. Sci. Lett. 444,
131–138 (2016).
33. Ieda, M. & Tada, . Long period astronomical cycles from the Triassic to Jurassic bedded chert sequence (Inuyama, Japan),
Geologic evidences for the chaotic behavior of solar planets. Earth Planets Space 65, 1–10 (2013).
34. Sprovieri, M. et al. Late Cretaceous orbitally-paced carbon isotope stratigraphy from the Bottaccione Gorge (Italy). Palaeogeo.
Palaeoclim. Palaeoecol. 379–380, 81–94 (2013).
35. Martinez, M. & Dera, G. Orbital pacing of carbon uxes by a 9-My eccentricity cycle during the Mesozoic. Proc. Natl. Acad. Sci.
112, 12604–12609 (2015).
36. Svensmar, H. Imprint of Galactic dynamics on Earth’s climate. Astron. Nachr./AN 327(9), 866–870 (2006).
37. ampino, M. . Disc dar matter in the Galaxy and potential cycles of extraterrestrial impacts, mass extinctions and geological
events. MNRAS 448, 1816–1820 (2015).
38. Bahcall, J. N. & Bahcall, S. e Sun’s motion perpendicular to the galactic plane. Nature 316, 706–708 (1985).
39. Stothers, . B. 1998. Galactic dis dar matter, terrestrial impact cratering and the law of large numbers. Mon. Not. R. Astron. Soc.
300, 1098–1104 (1998).
40. andal, L. & eece, M. Dar matter as a trigger for periodic comet impacts. Phys. Rev. Lett. 112(161301), 1–161301.5 (2014).
41. Svensmar, H. Cosmoclimatology: a new theory emerges. Astron. Geophys. 48(1), 18–1.24 (2007).
42. Gies, D. . & Helsel, J. W. Ice age epochs and the sun’s path through the galaxy. Astrophys. J. 626, 844–847 (2005).
43. irby, J. Cosmic rays and climate. Surveys in Geophysics 28, 333–375, https://doi.org/10.1007/s10712-008-9030-6 (2007).
44. Huybers, P. & Langmuir, C. Feedbac between deglaciation, volcanism and atmospheric CO2. Earth Planet. Sci. Lett. 286, 479–491
(2009).
45. Daradich, A., Huybers, P., Mitovica, J. X., Chan, N.-H. & Austermann, J. e inuence of true polar wander on glacial inception in
North America. Earth Planet. Sci. Lett. 461, 96–104 (2017).
46. utterolf, S. et al. A detection of Milanovitch frequencies in global volcanic activity. Geology 41, 227–230, https://doi.org/10.1130/
G33419.1 (2012).
47. Crowley, J. W., atz, . F., Huybers, P., Langmuir, H. C. & Sung-Hyun, P. Glacial cycles drive variations in the production of oceanic
crust. Science 347(6227), 1237–1240 (2015).
48. Huybers, P. & Langmuir, C. H. Delayed CO2 emissions from mid-ocean ridge volcanism as a possible cause of late-Pleistocene glacial
cycles. Earth Planet. Sci. Lett. 457, 238–249 (2017).
49. Lamb, S. & Davis, P. M. C enozoic climate change as a possible cause for the rise of the Andes. Nature 425, 792–297 (2003).
50. Lagabrielle, Y., Goddéris, Y., Donnadieu, Y., Malavieille, J. & Suarez, M. e tectonic history of Drae Passage and its possible
impacts on global climate. Earth Planet. Sci. Lett. 279, 197–211 (2009).
51. Whipple, . X. e inuence of climate on the tectonic evolution of mountain belts. Nature Geosciences 2, 97–104 (2009).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
7
SCIEntIfIC REpoRTS | (2019) 9:327 | DOI:10.1038/s41598-018-36509-7
52. L asar, J. et al. A long-term numerical solution for the insolution quantities of the Earth. Astronomy and Astrophysics 428, 261–285
(2004).
53. Lasar, J., Fienga, A., Gastineau, M. & Manche, H. La2010: A new orbital solution for the long term motion of the Earth. Astron.
Astrophys. 532, A89, https://doi.org/10.1051/0004-6361/201116836 (2011).
54. Meyers, S. ., 2014. Astrochon: An  pacage for astrochronology. Available at: https://cran.rproject.org/web/pacages/astrochron/
index.html.
Acknowledgements
I thank O. Friedrich (Heidelberg University), who kindly provided the compiled benthic foraminifera isotope
data. is work was supported by INSU-SYSTER grant. I acknowledge reviews of two anonymous colleagues and
the associate editor that led to important revisions of my manuscript.
Additional Information
Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-018-36509-7.
Competing Interests: e author declares no competing interests.
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2019
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com

Supplementary resource (1)

... For the low-frequency orbital terms, energy could be transferred from higher to lower frequency terms via amplitude and/or frequency modulation (e.g., Boulila et al., 2012). Indeed, time-series analysis of astronomical models La2004 (Laskar et al., 2004) and La2010d indicate that a disturbance to the 405-kyr eccentricity cycle leads to 4.7-and 9.5-Myr grand eccentricity cycles (Boulila et al., 2012(Boulila et al., , 2020(Boulila et al., , 2021Sprovieri et al., 2013;Martinez and Dera, 2015;Boulila, 2019). The frequency decomposition of the proper modes of the secular frequencies in the precession of the perihelia and nodes of the inner planets allows for the extraction of the 4.7-Myr cyclicity (Laskar, 1990;Boulila et al., 2012Boulila et al., , 2020Wang et al., 2020). ...
... The frequency decomposition of the proper modes of the secular frequencies in the precession of the perihelia and nodes of the inner planets allows for the extraction of the 4.7-Myr cyclicity (Laskar, 1990;Boulila et al., 2012Boulila et al., , 2020Wang et al., 2020). Perihelial or nodal precession has been suggested as the cause of the 9.5-Myr cyclicity, which was derived from the modulation of the 2.4-Myr eccentricity cycle band (Boulila et al., 2012(Boulila et al., , 2020Boulila, 2019). Amplitude modulation analysis of the ~10-Myr eccentricity band further revealed 35-Myr cyclicity (Boulila et al., 2021) in both the La2004 and La2010d astronomical solutions. ...
... Key evidence from global sea-level data compilations indicates that the 35-Myr cycle is caused by amplitude modulation of the 10-Myr Milankovitch cycle (Boulila et al., 2021). Prominent ~5-and ~ 9-Myr geological cycles have been detected successively in the planktonic foraminifera diversity records (Prokoph et al., 2004), the benthic foraminifera δ 13 C record (Boulila et al., 2012), the belemnite δ 13 C and δ 18 O records (Martinez and Dera, 2015), and the benthic foraminifera δ 18 O record (Boulila, 2019) from the Mesozoic and Cenozoic. Furthermore, abrupt climatic perturbations, such as the Paleocene-Eocene Thermal Maximum, could coincide with a critical interval in the 9-Myr cycling (Boulila et al., 2012;Boulila, 2019). ...
... Apart from the well-known astronomical cycles with periods of 19 kyr, 23 kyr, 41 kyr, 100 kyr, and 400 kyr that pace Earth's climate 3 , the geological record also contains signals of much longerperiod "grand cycles" 4 , which are predicted by astronomical theory 5 . These "grand cycles" include orbitally-forced periodicities of millions and even tens of millions of years that are similarly linked to changes in incoming solar radiation and paleoclimate [6][7][8][9] . The 2.4 Myr (g4-g3) eccentricity cycle related to the precession of the perihelions of Earth (g3) and Mars (g4), and the 1.2 Myr (s4-s3) obliquity cycle associated with the precession of the nodes of the two planets, are of particular interest 5 . ...
... Most studies of grand orbital cycles are focused on relatively short (<10 Myr), high-resolution continuous stratigraphic records that yield several eccentricity cycles in the stable 405 kyr band 10,11 , but very few longer-period (2.4 Myr) modulating cycles e.g., ref. 12. The~2.4 Myr eccentricity cycle, however, has been found embedded in the Cenozoic δ 18 O and δ 13 C isotope record [7][8][9]12,13 , in geophysical signals of stratigraphic sequences 4,[14][15][16] , and in fossil assemblages 17,18 . The cycles are variably attributed to changes in temperature and ice-volume 7 , fluctuations in sediment and organic carbon accumulation linked to river runoff 8 , variations in seawater temperature 12 , and changes in ocean circulation and structure 17 all of which are driven by astronomically forced changes in insolation and climate. ...
Article
Full-text available
Astronomical forcing of Earth’s climate is embedded in the rhythms of stratigraphic records, most famously as short-period (10⁴–10⁵ year) Milankovitch cycles. Astronomical grand cycles with periods of millions of years also modulate climate variability but have been detected in relatively few proxy records. Here, we apply spectral analysis to a dataset of Cenozoic deep-sea hiatuses to reveal a ~2.4 Myr eccentricity signal, disrupted by episodes of major tectonic forcing. We propose that maxima in the hiatus cycles correspond to orbitally-forced intensification of deep-water circulation and erosive bottom current activity, linked to eccentricity maxima and peaks in insolation and seasonality. A prominent episode of cyclicity disturbance coincides with the Paleocene-Eocene Thermal Maximum (PETM) at ~56 Myr ago, and correlates with a chaotic orbital transition in the Solar System evident in several astronomical solutions. This hints at a potential intriguing coupling between the PETM and Solar System chaos.
... Even more important are the astral factors that contribute to modifying the climatic situation of the earth, they play a decisive role, they include: the earth's orbit and the quantity of solar rays that reach its hemispheres, the alternation of the equinoxes , the orbital eccentricity and the inclination of the Earth's axis. They are all temporal events that have a cyclical trend [14]. These cycles explain the global climate changes that occurred over millions of years corresponding to the periods of glaciation/ deglaciation described by Paleoclimatology. ...
... , are more frequent (Figure S2; see alsoBoulila, 2019;Boulila et al., 2018Boulila et al., , 2021 for more detailed spectral analyses) than the usually proposed alternate between greenhouse and icehouse conditions(Figure 1).In the present-day world(from 1992 to 2016), if all continental ice would melt, a large volume of water would be added to the ocean, which would correspond to a sea-level rise of 65.5 ± 1.0 m (IPCC, 2013; Bamber et al., 2018; ...
Article
Full-text available
Throughout the Phanerozoic and more, the Earth has experienced cold and hot periods, which are typically associated with long‐lasting (hundreds of million years, Ma) greenhouse and icehouse climate regimes. Now, most published sea‐level curves report two main maxima in the Cretaceous and Ordovician superimposed on a multitude of short‐term fluctuations. The big humps are shown to be predominantly the results of the plate tectonic configuration, not icehouse and greenhouse regimes, suggesting that the small oscillations are related to continental ice variations. From this point of view, it can be inferred that polar ice caps are present almost all the time, and climate regime changes appear much more frequent and shorter than usually considered and are not well‐documented from glaciogenic deposits. Relying on short‐term oscillations, the volume of continental ice can be retrieved over the Phanerozoic.
Article
The Cenomanian to Turonian (C-T) transition marks a period of significant environmental disturbance characterized by shifts in climate, biogeochemical cycles, and ecosystems. During this time, marine ecosystems struggled to adapt to ocean oxygenation and nutrient availability changes. This study hypothesizes that Oceanic Anoxic Event 2 (OAE2) was driven by orbital influences from the 2.4 and 1.1 million-year g4-g3 eccentricity and obliquity cycles, respectively, during the C-T transition. These orbital factors probably contributed to contrasting climatic conditions, which played a crucial role in the impact on biogeochemical cycles and ecosystems. The astronomical calibration of the Oued Ettalla section reveals a period of 2.15 million years, from 94.73 to 92.58 million years ago. This newly established, astronomically-calibrated timeline, based on previously published data from the same section and samples, includes six calcareous nannoplankton bioevents, ten foraminiferal bioevents, and seven carbon isotope excursions, providing a refined temporal framework for understanding the paleoenvironmental changes. The different phases of OAE2 (a-d), distinguished by chemostratigraphic data occurred during a 0.8 million-year portion of the obliquity antinode modulation. The Lulworth carbon isotope event marks the onset of the Middle Turonian, while two 1.1 million-year obliquity modulations frame the Early Turonian. The Filament Event coincided with an eccentricity and obliquity node, which led to intensified summer insolation and increased ocean temperatures, supported by warm-water nannofossils.
Article
As one of the most important greenhouse gases, CO2 is considered a major controlling factor of Earth’s climate over geological timescales. However, the origins of quasi-periodic fluctuations in pCO2 on a million-year time- scale remain unclear. Here, we used published datasets of atmospheric pCO2, oxygen isotopes of benthic fora- minifera (δ18Obenthic) and global mean sea-level (GMSL) from 23 Ma to the present to explore the pacing of pCO2 changes and concomitant climatic effects using multiple time series analysis approaches. Our results indicate that the evolution of late Cenozoic pCO2 and climate was paced by the grand orbital cycles, in particular the ~4.5 Myr and ~2.4 Myr eccentricity cycles, and ~1.3 Myr obliquity cycle. Periodic occurrence of cold conditions was associated with low climate seasonality during the minima of ~4.5 Myr and ~2.4 Myr eccentricity cycles. We suggest that cooler conditions are associated with decreased atmospheric pCO2 as a result of higher organic carbon burial due to lower metabolic rate of heterotrophic bacteria and more organic carbon export to the deep ocean. Furthermore, the buildup of glaciers during the minima of grand eccentricity cycles might lower pCO2 via increased ice cover and enhanced dust fluxes. In contrast, high seasonal climate may lead to an opposite effect on atmospheric pCO2 during the maxima of the grand eccentricity cycles. Moreover, we found a distinct shift in the dominant signal from eccentricity to obliquity cycles recorded in the pCO2, δ18Obenthic and GMSL datasets at ~13 Ma, a time when perennial sea ice occurred in the Arctic and significant ice growth shown in Antarctica. We suggest that the change in the type and distribution of the ice sheets would shift glacial response to orbital forcing and hence mediated global climate and pCO2. Our analysis reveals a clear synchrony among atmospheric pCO2, climate change, and the grand orbital cycles in the late Cenozoic.
Preprint
Full-text available
The Upper Triassic Dachstein Limestone is an archetypal example of platform carbonate formation that consists of peritidal-subtidal Lofer cycles. However, despite a long history of study, their allocyclic vs autocyclic origin remains controversial. Using modern cyclostratigraphic methodology, here we revisit the archive data of the Dachstein Limestone in core Po-89 from the Transdanubian Range, Hungary. Analysis of high-resolution time series of the Lofer facies types, colour, greyscale, and lithology revealed several spectral peaks that are identified with the Milankovitch cycles. The Lofer cycles correspond to a robust peak with a period of 3.6 m that is the expression of precession cycles with a period 21.65 kyr. The sedimentary rate is 16.5 cm/kyr, thus the studied core section represents 2.5 Myr. Our model of cyclic carbonate deposition suggests that aquifer- and limno-eustasy were the most likely drivers of cyclic sea level changes. Differences in the depositional processes and preservation of the Lofer cycles led to variations in their completeness that is also orbitally controlled and responds primarily to long eccentricity forcing. We detected several sub-Milankovitch cycles, of which the 13.5, 7, 5, 3.4, 2.4, and 1.48 kyr cycles are the most prominent. Our results are in good agreement with other studies of Upper Triassic Lofer cyclic successions. The similarities in the thickness of individual Lofer cycles, the periods of all identified cycles, and the sedimentary rates suggest common allocyclic processes. Our findings provide definitive evidence for the orbital forcing of carbonate sedimentation in the Late Triassic Dachstein platform system.
Article
Full-text available
Cyclic sedimentation has varied at several timescales and this variability has been geologically well documented at Milankovitch timescales, controlled in part by climatically (insolation) driven sea-level changes. At the longer (tens of Myr) timescales connection between astronomical parameters and sedimentation via cyclic solar-system motions within the Milky Way has also been proposed, but this hypothesis remains controversial because of the lack of long geological records. In addition, the absence of a meaningful physical mechanism that could explain the connection between climate and astronomy at these longer timescales led to the more plausible explanation of plate motions as the main driver of climate and sedimentation through changes in ocean and continent mass distribution on Earth. Here we statistically show a prominent and persistent ~36 Myr sedimentary cyclicity superimposed on two megacycles (~250 Myr) in a relatively well-constrained sea-level (SL) record of the past 542 Myr (Phanerozoic eon). We also show two other significant ~9.3 and ~91 Myr periodicities, but with lower amplitudes. The ~9.3 Myr cyclicity was previously attributed to long-period Milankovitch band based on the Cenozoic record. However, the ~91 Myr cyclicity has never been observed before in the geologic record. The ~250 Myr cyclicity was attributed to the Wilson tectonic (supercontinent) cycle. The ~36 Myr periodicity, also detected for the first time in SL record, has previously been ascribed either to tectonics or to astronomical cyclicity. Given the possible link between amplitudes of the ~36 and ~250 Myr cyclicities in SL record and the potential that these periodicities fall into the frequency band of solar system motions, we suggest an astronomical origin, and model these periodicities as originating from the path of the solar system in the Milky Way as vertical and radial periods that modulate the flux of cosmic rays on Earth. Our finding of the ~36 Myr SL cyclicity lends credibility to the existing hypothesis about the imprint of solar-system vertical period on the geological record. The ~250 Myr megacycles are tentatively attributed to a radial period. However, tectonic causal mechanisms remain equally plausible. The potential existence of a correlation between the modeled astronomical signal and the geological record may offer an indirect proxy to understand the structure and history of the Milky Way by providing a 542 Myr long record of the path of the Sun in our Galaxy.
Article
Full-text available
Significance The Antarctic ice cap waxed and waned on astronomical time scales throughout the Oligo-Miocene time interval. We quantify geometries of Antarctic ice age cycles, as expressed in a new climate record from the South Atlantic Ocean, to track changing dynamics of the unipolar icehouse climate state. We document numerous ∼110-thousand-year-long oscillations between a near-fully glaciated and deglaciated Antarctica that transitioned from being symmetric in the Oligocene to asymmetric in the Miocene. We infer that distinctly asymmetric ice age cycles are not unique to the Late Pleistocene or to extremely large continental ice sheets. The patterns of long-term change in Antarctic climate interpreted from this record are not readily reconciled with existing CO 2 records.
Article
Full-text available
The impact that long-term changes in Earth's rotation axis relative to the surface geography, or true polar wander (TPW), and continental drift have had in driving cooling of high-latitude North America since the Eocene is explored. Recent reanalyses of paleomagnetic pole positions suggest a secular drift in Earth's rotation axis of ∼8° over the last 40 Myr, in a direction that has brought North America to increasingly higher latitudes. Using modern temperature data in tandem with a simple model, a reduction in the annual sum of positive degree days (PDDs) driven by this polar and plate motion over the last 20 Myr is quantified. At sites in Baffin Island, the TPW- and continental drift-driven decrease in insolation forcing over the last 20 Myr rivals changes in insolation forcing caused by variations in Earth's obliquity and precession. Using conservative PDD scaling factors and an annual snowfall equal to modern station observations, the snowiest location in Baffin Island 20 Myr ago had a mass balance deficit of ∼0.75–2 m yr⁻¹ (water equivalent thickness) relative to its projected mass balance at 2.7 Ma. This mass balance deficit would have continued to increase as one goes back in time until ∼40 Myr ago based on adopted paleopole locations. TPW and continental drift that moved Arctic North America poleward would have strongly promoted glacial inception in Baffin Island at ∼3 Ma, a location where the proto-Laurentide Ice Sheet is thought to have originated.
Article
Full-text available
The coupled 100,000 year variations in ice volume, temperature, and atmospheric CO2 during the late Pleistocene are generally considered to arise from a combination of orbital forcing, ice dynamics, and ocean circulation. Also previously argued is that changes in glaciation influence atmospheric CO2 concentrations through modifying subaerial volcanic eruptions and CO2 emissions. Building on recent evidence that ocean ridge volcanism responds to changes in sea level, here it is suggested that ocean ridges may play an important role in generating late-Pleistocene 100 ky glacial cycles. If all volcanic CO2 emissions responded immediately to changes in pressure, subaerial and ocean-ridge volcanic emissions anomalies would oppose one another. At ocean ridges, however, the egress of CO2 from the mantle is likely to be delayed by tens-of-thousands of years, or longer, owing to ascent time. A simple model involving temperature, ice, and CO2 is presented that oscillates at ∼100 ky time scales when incorporating a delayed CO2 contribution from ocean ridge volcanism, even if the feedback accounts for only a small fraction of total changes in CO2. Oscillations readily become phase-locked with insolation forcing associated with changes in Earth's orbit. Under certain parameterizations, a transition from ∼40 ky to larger ∼100 ky oscillations occurs during the middle Pleistocene in response to modulations in orbital forcing. This novel description of Pleistocene glaciation should be testable through ongoing advances in understanding the circulation of carbon through the solid earth.
Article
Full-text available
Few astronomically calibrated high-resolution (≤5 kyr) climate records exist that span the Oligocene–Miocene time interval. Notably, available proxy records show responses varying in amplitude at frequencies related to astronomical forcing, and the main pacemakers of global change on astronomical time-scales remain debated. Here we present newly generated X-ray fluorescence core scanning and benthic foraminiferal stable oxygen and carbon isotope records from Ocean Drilling Program Site 1264 (Walvis Ridge, southeastern Atlantic Ocean). Complemented by data from nearby Site 1265, the Site 1264 benthic stable isotope records span a continuous ∼13-Myr interval of the Oligo-Miocene (30.1–17.1 Ma) at high resolution (∼3.0 kyr). Spectral analyses in the stratigraphic depth domain indicate that the largest amplitude variability of all proxy records is associated with periods of and , which correspond to 405- and ∼110-kyr eccentricity, using a magnetobiostratigraphic age model. Maxima in CaCO3 content, δ18O and δ13C are interpreted to coincide with ∼110 kyr eccentricity minima. The strong expression of these cycles in combination with the weakness of the precession- and obliquity-related signals allow construction of an astronomical age model that is solely based on tuning the CaCO3 content to the nominal (La2011_ecc3L) eccentricity solution. Very long-period eccentricity maxima (∼2.4-Myr) are marked by recurrent episodes of high-amplitude ∼110-kyr δ18O cycles at Walvis Ridge, indicating greater sensitivity of the climate/cryosphere system to short eccentricity modulation of climatic precession. In contrast, the responses of the global (high-latitude) climate system, cryosphere, and carbon cycle to the 405-kyr cycle, as expressed in benthic δ18O and especially δ13C signals, are more pronounced during ∼2.4-Myr minima. The relationship between the recurrent episodes of high-amplitude ∼110-kyr δ18O cycles and the ∼1.2-Myr amplitude modulation of obliquity is not consistent through the Oligo-Miocene. Identification of these recurrent episodes at Walvis Ridge, and their pacing by the ∼2.4-Myr eccentricity cycle, revises the current understanding of the main climate events of the Oligo-Miocene.
Article
Full-text available
The claims of periodicity in impact cratering and biological extinction events are controversial. A newly revised record of dated impact craters has been analyzed for periodicity, and compared with the record of extinctions over the past 260 Myr. A digital circular spectral analysis of 37 crater ages (ranging in age from 15 to 254 Myr ago) yielded evidence for a significant 25.8 ± 0.6 Myr cycle. Using the same method, we found a significant 27.0 ± 0.7 Myr cycle in the dates of the eight recognized marine extinction events over the same period. The cycles detected in impacts and extinctions have a similar phase. The impact crater dataset shows 11 apparent peaks in the last 260 Myr, at least 5 of which correlate closely with significant extinction peaks. These results suggest that the hypothesis of periodic impacts and extinction events is still viable.
Article
Full-text available
Significance The Milankovitch cycles are orbitally paced variations in insolation that drove periodic climate changes on Earth at the scale of tens to hundreds of thousands years. Longer ‟grand orbital cycles” also exist, but their impacts on paleoclimate dynamics are not well documented for pre-Cenozoic times. Here we tackle this issue by analyzing the stable isotope fluctuations recorded by fossil cephalopods throughout the Jurassic–Early Cretaceous interval. We document a periodicity of ∼9 My in the carbon cycle, except from 190 to 180 Ma when disturbances occurred. This orbital forcing affected carbon transfers by modulating the hydrological processes and sea-level changes. In summary, this ∼9-My orbital cycle is an important metronome of the greenhouse climate dynamics.
Article
Astronomical theory predicts that ∼2 Myr eccentricity cycle have changed its periodicity and amplitude through time because of the chaotic behavior of solar planets, especially Earth-Mars secular resonance. Although the ∼2 Myr eccentricity cycle has been occasionally recognized in geological records, their frequency transitions have never been reported. To explore the frequency evolution of ∼2 Myr eccentricity cycle, we used the bedded chert sequence in Inuyama, Japan, of which rhythms were proven to be of astronomical origin, covering the ∼30 Myr long spanning from the Triassic to Jurassic. The frequency modulation of ∼2 Myr cycle between ∼1.6 and ∼1.8 Myr periodicity detected from wavelet analysis of chert bed thickness variation are the first geologic record of chaotic transition of Earth-Mars secular resonance. The frequency modulation of ∼2 Myr cycle will provide new constraints for the orbital models. Additionally, ∼8 Myr cycle detected as chert bed thickness variation and its amplitude modulation of ∼2 Myr cycle may be related to the amplitude modulation of ∼2 Myr eccentricity cycle through non-linear process(es) of Earth system dynamics, suggesting possible impact of the chaotic behavior of Solar planets on climate change.