ArticlePDF Available

A New Geologic Time Scale, with special reference to Precambrian and Neogene

Authors:

Abstract and Figures

A Geologic Time Scale (GTS2004) is presented that inte- grates currently available stratigraphic and geochrono- logic information. Key features of the new scale are out- lined, how it was constructed, and how it can be further improved. The accompanying International Strati- graphic Chart, issued under auspices of the Interna- tional Commission on Stratigraphy (ICS), shows the cur- rent chronostratigraphic scale and ages with estimates of uncertainty for all stage boundaries. Special reference is made to the Precambrian part of the time scale, which is coming of age in terms of detail, and to the Neogene portion, which has attained an ultra-high-precision absolute-age calibration.
Content may be subject to copyright.
A Geologic Time Scale (GTS2004) is presented that inte-
grates currently available stratigraphic and geochrono-
logic information. Key features of the new scale are out-
lined, how it was constructed, and how it can be further
improved. The accompanying International Strati-
graphic Chart, issued under auspices of the Interna-
tional Commission on Stratigraphy (ICS), shows the cur-
rent chronostratigraphic scale and ages with estimates
of uncertainty for all stage boundaries. Special reference
is made to the Precambrian part of the time scale, which
is coming of age in terms of detail, and to the Neogene
portion, which has attained an ultra-high-precision
absolute-age calibration.
Introduction
The geologic time scale is the framework for deciphering the history
of the Earth and has three components:
(1) The international chronostratigraphic divisions and their cor-
relation in the global rock record,
(2) The means of measuring absolute (linear) time or elapsed
durations from the rock record, and
(3) The methods of effectively joining the two scales.
For convenience in international communication, the rock
record of Earth’s history is subdivided in a “chronostratigraphic”
scale of standardized global stratigraphic units, such as “Ordovi-
cian”, “Miocene”, “Harpoceras falciferum ammonite Zone” or
“polarity Chron C24r”. Unlike the continuous ticking clock of the
“chronometric” scale (measured in years before present), the
chronostratigraphic scale is based on relative time units, in which
global reference points at boundary stratotypes define the limits of
the main formalized units, such as “Devonian,” The chronostrati-
graphic scale is an agreed convention, whereas its calibration to
absolute (linear) time is a matter for discovery or estimation.
By contrast, Precambrian stratigraphy is formally classified
chronometrically, i.e. the base of each Precambrian eon, era and
period is assigned an arbitrary numerical age. This practice is now
being challenged (see below).
Continual improvements in data coverage, methodology, and
standardization of chronostratigraphic units imply that no geologic
time scale can be final. This brief overview of the status of the Geo-
logic Time Scale in 2004 (GTS2004), documented in detail by Grad-
stein et al. (2004), is the successor to GTS1989 (Harland et al.,
1990), which in turn was preceded by GTS1982 (Harland et al.,
1982). GTS2004 also replaces the International Stratigraphic Chart
2000 of the International Commission on Stratigraphy (ICS) and
UNESCO, issued four years ago (Remane, 2000).
There are several reasons why this new geologic time scale of
2004 was required, including:
Nearly 50 of 90+ Phanerozoic stage boundaries are now defined,
versus <15 in 1990;
Stable international stage subdivisions rendered invalid about
15% of the “stage” names of 1990;
Last 23 million years (Neogene) is orbitally tuned with 40 kyr
accuracy;
Orbital scaling has been successful in portions of the Paleocene,
lower Cretaceous, lower Jurassic, and upper Triassic;
Superior stratigraphic integration in Mesozoic has merged direct
dating, seafloor spreading (M-sequence), zonal scaling and
orbital tuning;
Superior stratigraphic scaling of Paleozoic was achieved using
high-resolution composite zonal standards;
A ‘natural’ geologic Precambrian time scale has been proposed
to replace the current artificial scale;
More accurate and precise age dating has provided over 200
Ar/Ar and U/Pb dates with external (systematic) error analysis,
of which only a few of these were available to GTS89;
Improved mathematical/statistical techniques can combine bios-
tratigraphic zones, polarity chrons, geologic stages and absolute
ages to calculate the linear time scale and estimate uncertainty.
A listing is provided at the end of this document of outstanding
issues that, once resolved, will pave the way for an updated version
of the standard Geologic Time Scale, scheduled under the auspices
of ICS for the year 2008.
Overview of construction of GTS2004
Since 1989, there have been major developments in time scale
research, including:
(1) Stratigraphic standardization through the work of the Inter-
national Commission on Stratigraphy (ICS) has greatly refined the
Episodes, Vol. 27, no. 2
Articles 83
by Felix M. Gradstein
1
, James G. Ogg
2
, Alan G. Smith
3
, Wouter Bleeker
4
, and
Lucas J. Lourens
5
A new Geologic Time Scale, with special
reference to Precambrian and Neogene
1. Geological Museum, University of Oslo, N-0318 Oslo, Norway. Email: felix.gradstein@nhm.uio.no
2. Department of Earth & Atmospheric Sciences, Purdue University, West Lafayette, Indiana 47907-1397, USA.
3. Department of Earth Sciences, Cambridge University, Cambridge CB2 3EQ, England.
4. Geological Survey of Canada, 601 Booth Str., Ottawa, Ontario K1A OE8, Canada.
5. Faculty of Earth Sciences, Utrecht University, 3508 TA Utrecht, The Netherlands.
Note: This article provides an excerpt of Geologic Time Scale 2004 (Cambridge University Press, ~500 pp.). The Time Scale Project is a joint undertaking of
F.M. Gradstein, J.G. Ogg, A.G. Smith, F.P. Agterberg, W. Bleeker, R.A. Cooper, V. Davydov, P. Gibbard, L.A. Hinnov, M.R. House (†), L.J. Lourens, H-P.
Luterbacher, J. McArthur, M.J. Melchin, L.J. Robb, J. Shergold, M. Villeneuve, B.R. Wardlaw, J. Ali, H. Brinkhuis, F.J. Hilgen, J. Hooker, R.J. Howarth, A.H.
Knoll, J. Laskar, S. Monechi, J. Powell, K.A. Plumb, I. Raffi, U. Röhl, A. Sanfilippo, B. Schmitz, N.J. Shackleton, G.A. Shields, H. Strauss, J. Van Dam, J. Veizer,
Th. van Kolfschoten, and D. Wilson, and is under auspices of the International Commission on Stratigraphy.
international chronostratigraphic scale. In some cases, such as in the
Ordovician or Permian periods, traditional European- or Asian-
based geological stages have been replaced with new subdivisions
that allow global correlation.
(2) New or enhanced methods of extracting linear time from the
rock record have enabled high-precision age assignments. Numerous
high-resolution radiometric dates have been generated that has led to
improved age assignments of key geologic stage boundaries. The use
of global geochemical variations, Milankovitch climate cycles, and
magnetic reversals have become important calibration tools.
(3) Statistical techniques of interpolating ages and associated
uncertainties to stratigraphic events have evolved to meet the chal-
lenge of more accurate age dates and more precise zonal assign-
ments. Fossil event databases with multiple stratigraphic sections
through the globe can be integrated into high-resolution composite
standards for internal scaling of geologic stages.
The compilation of GTS2004 involved a large number of spe-
cialists, listed above, including contributions by past and present
chairs of different subcommissions of ICS, geochemists working
with radiogenic and stable isotopes, stratigraphers using diverse
tools from traditional fossils to astronomical cycles to database pro-
gramming, and geomathematicians.
The methods used to construct Geologic Time Scale 2004
(GTS2004) integrate different techniques depending on the quality
of data available within different intervals (Figure 1). The set of
chronostratigraphic units (geologic stages, periods) and their com-
puted ages that constitute the main framework for the Geologic Time
Scale 2004 are summarized in the International Stratigraphic Chart
(Figure 2 and accompanying insert). Uncertainties on ages are
expressed at 2-sigma (95% confidence). Table 1 summarizes the sta-
tus of stratigraphic standardization, compiled by one of us (JGO), for
the entire geologic column. Steady progress is made with further
standardization of the stratigraphic scale.
The main steps involved in the GTS2004 time scale construc-
tion were:
Step 1. Construct an updated global chronostratigraphic scale
for the Earth’s rock record (Table 1).
Step 2. Identify key linear-age calibration levels for the
chronostratigraphic scale using radiometric age dates, and/or apply
astronomical tuning to cyclic sediment or stable isotope sequences
which had biostratigraphic or magnetostratigraphic correlations.
Step 3. Interpolate the combined chronostratigraphic and
chronometric scale where direct information is insufficient.
Step 4. Calculate or estimate error bars on the combined
chronostratigraphic and chronometric information to obtain a geo-
logic time scale with estimates of uncertainty on boundaries and on
unit durations.
Step 5. Peer review the geologic time scale through ICS.
The first step, integrating multiple types of stratigraphic infor-
mation in order to construct the chronostratigraphic scale, is the most
time-consuming. This relative geologic time scale summarizes and
synthesizes centuries of detailed geological research. The second
step, identifying which radiometric and cycle-stratigraphic studies
would be used as the primary constraints for assigning linear ages, is
the one that is evolved most rapidly during the past decade. Histori-
cally, Phanerozoic time scale building went from an exercise with
very few and relatively inaccurate radiometric dates, as used by
Holmes (1947, 1960), to one with many dates with greatly varying
analytical precision (like GTS89, or to some extent Gradstein et al.,
1994). Next came studies on relatively short stratigraphic intervals
that selected a few radiometric dates with high internal analytical
precision (e.g., Obradovich, 1993; Cande & Kent, 1992, 1995;
Cooper, 1999) or measured time relative to present using astronom-
ical cycles (e.g., Shackleton et al., 1999; Hilgen et al., 1995, 2000).
This later philosophy is adhered to in this scale.
In addition to selecting radiometric ages based upon their strati-
graphic control and analytical precision, we also applied the follow-
ing criteria or corrections:
(1) Stratigraphically constrained radiometric ages with the
U-Pb method on zircons were accepted from the isotope dilution
mass spectrometry (TIMS) method, but generally not from the high-
resolution ion microprobe (HR-SIMS, also known as “SHRIMP”)
that uses the Sri Lanka (SL)13 standard. An exception is the Car-
boniferous Period, where there is a dearth of TIMS dates, and more
uncertainty.
(2)
40
Ar-
39
Ar radiometric ages were re-computed to be in
accord with the revised ages for laboratory monitor standards: 523.1
±4.6 Ma for MMhb-1 (Montana hornblende), 28.34 ±0.28 Ma for
TCR (Taylor Creek sanidine) and 28.02 ±0.28 Ma for
FCT (Fish Canyon sanidine). Systematic (“external”) errors
and uncertainties in decay constants are partially incorpo-
rated. No glauconite dates are used.
The bases of the Paleozoic, Mesozoic and Cenozoic
eras are bracketed by analytically precise ages at their GSSP
(Global Standard Section and Point) or primary correlation
markers — 542 ±1.0 Ma, 251.0 ±0.4 Ma, and 65.5 ±0.3 Ma
— and there are direct age-dates on base-Carboniferous,
base-Permian, base-Jurassic, and base-Oligocene; but most
other period or stage boundaries prior to the Neogene lack
direct age control. Therefore, the third step, interpolation,
plays a key role for most of GTS2004. A set of detailed and
high-resolution interpolation processes incorporated several
techniques, depending upon the available information:
(1) A composite standard of graptolite zones spanning
the uppermost Cambrian, Ordovician and Silurian interval
was derived from 200+ sections in oceanic and slope envi-
ronment basins using the constrained optimization method.
With zone thickness taken as directly proportional to zone
duration, the detailed composite sequence was scaled using
selected, high precision zircon and sanidine age dates. For
the Carboniferous through Permian a composite standard of
conodont, fusulinid, and ammonoids events from many
classical sections was calibrated to a combination of U-Pb
and
40
Ar-
39
Ar dates with assigned external error estimates.
A composite standard of conodont zones was used for Early
Triassic. This procedure directly scaled all stage boundaries
and biostratigraphic horizons.
June 2004
84
Figure 1 Methods used to construct the Geologic Time Scale 2004
(GTS2004) integrate different techniques depending on the quality of
data available within different intervals.
85
Figure 2 The International Startigraphical Chart summarizes the set of choronostratigraphic units (geologic stages, periods) and their computed ages, which are the main framework for
Geologic Time Scale 2004. Uncertainties on ages expressed at 2-sigma (95% confidence).
June 2004
86
Table 1 Status of defining lower boundaries of geologic stages with GSSPs (as of May, 2004). Updates of this compilation can be
obtained from the website (www.stratigraphy.org) of the International Commission on Stratigraphy (ICS) under IUGS.
Episodes, Vol. 27, no. 2
87
(Continued)
June 2004
88
(Continued)
Episodes, Vol. 27, no. 2
89
(Continued)
June 2004
90
(Continued)
Episodes, Vol. 27, no. 2
91
(Continued)
June 2004
92
(Continued)
Episodes, Vol. 27, no. 2
93
(Continued)
(2) Detailed direct ages for Upper Cretaceous ammonite zones
of the Western Interior of the USA were obtained by a cubic spline
fit of the zonal events and 25
40
Ar-
39
Ar dates. The base-Turonian
age is directly bracketed by this
40
Ar-
39
Ar set, and ages of other
stage boundaries and stratigraphic events are estimated using cali-
brations to this primary scale.
(3) Seafloor spreading interpolations were done on a composite
marine magnetic lineation pattern for the Late Jurassic through Early
Cretaceous in the Western Pacific and for the late Cretaceous
through early Neogene in the South Atlantic Oceans. Ages of bios-
tratigraphic events were assigned according to their calibration to
these magnetic polarity time scales.
(4) Astronomical tuning of cyclic sediments was used for Neo-
gene and Upper Triassic, and for portions of the Lower and Middle
Jurassic, Lower Cretaceous, and Paleocene. The Neogene astronom-
ical scale is directly tied to the Present; the older astronomical scale
provides absolute-duration constraints on polarity chrons, biostrati-
graphic zones and entire stages.
(5) Proportional scaling relative to component biozones or sub-
zones. In intervals where none of the above information under Items
1 through 4 was available, it was necessary to return to the method-
ology employed by previous time scales. This procedure was neces-
sary in portions of the Middle Triassic, and Middle Jurassic. Devon-
ian stages were scaled from approximate equal duration of a set of
high-resolution subzones of ammonoids and conodonts, fitted to an
array of high-precision dates.
The geomathematics employed for data sets (Items 1, 2, 3 and
5) constructed for the Ordovician-Silurian, Devonian, Carbonifer-
ous-Permian, Late Cretaceous, and Paleogene intervals involved
cubic spline curve fitting to relate the observed ages to their strati-
graphic position. During this process, the ages were weighted
according to their variances based on the lengths of their error bars.
A chi-square test was used for identifying and reducing the weights
of relatively few outliers with error bars that are much narrower than
could be expected on the basis of most ages in the data set.
Stratigraphic uncertainty was incorporated in the weights
assigned to the observed ages during the spline-curve fitting. In the
final stage of analysis, Ripley’s algorithm for Maximum Likelihood
fitting of a Functional Relationship (MLFR) was used for error esti-
mation, resulting in 2-sigma (95% confidence) error bars for the
computed chronostratigraphic boundary ages and stage durations.
The uncertainties on older stage boundaries generally increase
owing to potential systematic errors in the different radiometric
methods, rather than to the analytical precision of the laboratory
measurements. In this connection, we mention that biostratigraphic
error is fossil event and fossil zone dependent, rather than dependent
on linear age.
In Mesozoic intervals that were scaled using the seafloor
spreading model or proportionally scaled using paleontological sub-
zones, the assigned uncertainties are conservative estimates based on
variability observed when applying different assumptions (see dis-
cussions in the Triassic, Jurassic and Cretaceous chapters of
GTS2004). Ages and durations of Neogene stages derived from
orbital tuning are considered to be accurate within a precession cycle
(~20 kyr), assuming that all cycles are correctly identified, and that
the theoretical astronomical-tuning for progressively older deposits
is precise.
Precambrian
From the time of initial accretion and differentiation (ca. 4560 Ma)
to the first appearance of abundant hard-bodied fossils (the onset of
the Cambrian Period at 542 Ma), the Precambrian spans 88 percent
of Earth history. Yet, there is no coherent view of a geological time
scale to help describe, analyze, calibrate, and communicate the evo-
lution of planet Earth.
The status quo is a geological time scale for the Precambrian
that is both incomplete and flawed (e.g., Cloud, 1987; Crook, 1989;
Nisbet, 1991; Bleeker, 2003a), and is defined in terms of arbitrary,
strictly chronometric, absolute age boundaries that are divorced
from the only primary, objective, record of planetary evolution: the
extant rock record.
At a recent conference in Canada on the geological time scale
and its calibration (NUNA, 2003), co-sponsored by the International
Committee on Stratigraphy (ICS), there was broad consensus on the
view that this arbitrary, chronometrically defined, Precambrian time
scale fails to convey the richness of the Precambrian rock record and
therefore impedes scientific understanding of geological processes
by diverting attention away from observable, first-order, strati-
graphic boundaries and transitions.
Specific criticisms of the present Precambrian time scale are
outlined in the chapter on Precambrian by Bleeker in Gradstein et al.
(2004), but one key point deserves elaboration here: the uncertainty
in decay constants of
238
U and
235
U. These uncertainties (e.g., Lud-
wig, 2000) conspire in such a way that most age dates for the Pre-
cambrian (predominantly upper intercept
207
Pb/
206
Pb zircon ages,
particularly prior to 1 Ga) have a non-trivial fundamental “fuzzi-
ness” (e.g., about ±6.5 million years at ca. 2500 Ma). This funda-
mental uncertainty increases to ±10 million years at 4000 Ma. Defi-
nition of boundaries in terms of arbitrary, round, absolute ages,
although superficially appealing, is therefore naïve. Absolute-age
correlation of such boundaries between distant sections, on the basis
of even our best geochronometer (U-Pb ages on single zircons), can
be no better than ±5–10 million years (in terms of linear ages), even
if all other sources of uncertainty (e.g., analytical scatter, Pb loss, or
cryptic inheritance) are negligible. In principle, this fundamental
uncertainty could be reduced by defining boundaries explicitly in
terms of
207
Pb/
206
Pb zircon ages or isotopic ratios, rather than linear
age, but this would make any time scale even less transparent. Fur-
thermore, it would not solve the problem of intercalibration between
different chronometers.
Clearly, there can only be one conclusion: the Precambrian time
scale should be (re)defined in terms of the only objective physical
standard we have, the extant rock record. Boundaries should be
placed at key events or transitions in the stratigraphic record, to high-
light important milestones in the evolution of our planet. This would
be analogous to the “golden spike” GSSP approach employed in the
Phanerozoic. Various geochronometers (U-Pb;
40
Ar-
39
Ar; Re-Os,
etc.), each with their own inherent but independent uncertainties,
should be employed to calibrate meaningful stratigraphic boundaries
in linear time. The ultimate result should be a calibrated “natural”
time scale for planet Earth that reflects first-order events and transi-
tions in its complex evolution.
To achieve this ‘natural’ time scale we propose that the
2004–2008 mandate of the International Subcommission on the Pre-
cambrian under ICS is a comprehensive and internally consistent, as
well as practical, “natural” time scale for planet Earth. This ‘natural’
time scale should be complete with agreed upon “golden spikes” and
type sections (i.e., GSSPs) for all Precambrian eon and era bound-
aries, and, where needed, for those of periods (systems).
Such an international effort would help focus significant atten-
tion on key stratigraphic boundaries and type sections, and, in turn,
will stimulate multidisciplinary science into the causes for specific
boundaries and transitions, the fundamental processes involved,
their rates, and their calibration in absolute time.
Building on efforts by the previous Subcommittee on Precam-
brian Stratigraphy (e.g., Plumb, 1991), such a “naturalizing” of the
Precambrian time scale could largely preserve existing nomencla-
ture, in so far as it has gained acceptance in the literature, while for-
malizing other eon and era names that are in widespread use today,
e.g. the Hadean. Thus, by 2008, we would have, for the first time, a
complete and natural time scale that reflects and communicates the
entire, protracted, and complex evolution of planet Earth.
Figures 3 and 4 highlight the key points of this discussion. Fig-
ure 3 shows the formal current subdivision of the Precambrian,
annotated with known key events in Earth’s evolution. The practical
Geon scale from Hofmann (1990, 1991) provides a quick chrono-
metric shorthand notation. The interval highlighted “early Earth” is
June 2004
94
an informal designation commonly used for Earth’s first giga-year
from the time of accretion to ~3.5 Ga. Exponentially decreasing
impact intensity (curve on right) is schematic and includes the “late
heavy bombardment” episode. Stars indicate Sudbury and Vredefort
impact craters with diameters >50 km.
In the proposed “natural” Precambrian time scale, Earth history
is divided into six eons, with boundaries defined by what can be con-
sidered first-order “watersheds” in the evolution of our planet (Fig-
ure 4). The six eons can be briefly characterized as follows:
(1) “Accretion & Differentiation” — planet formation, growth
and differentiation up to the Moon-forming giant impact event;
(2) Hadean (Cloud, 1972) — intense bombardment and its con-
sequences, but no preserved supracrustals;
Episodes, Vol. 27, no. 2
95
Figure 3 Formal subdivisions of the Precambrian annotated with key events in Earth’s evolution. Geon scale from Hofmann (1990, 1991)
provides a quick chronometric shorthand notation.
(3) Archean — increasing crustal record from the oldest
supracrustals of Isua greenstone belt to the onset of giant iron for-
mation deposition in the Hamersley basin, likely related to increas-
ing oxygenation of the atmosphere;
(4) “Transition” — starting with deposition of giant iron forma-
tions up to the first bona fide continental red beds;
(5) Proterozoic — a nearly modern plate-tectonic Earth but
without metazoan life, except at its very top; and
(6) The Phanerozoic— characterized by metazoan life forms of
increasing complexity and diversity.
Some of the boundaries are currently poorly calibrated in
absolute time, whereas the onset of the Archean should “float” with
the oldest preserved supracrustal rocks, a distinction currently held
by ~3820–3850 Ma rocks of the Isua greenstone belt. Comparison is
shown to the lunar time scale (e.g., Guest and Greeley, 1977; Murray
et al., 1981; Spudis, 1999).
Neogene
The most detailed segment of the modern geologic time scale in
terms of resolution and accuracy is that for the Neogene, 23 Ma to
Recent. The subdivision of the Neogene into its constituent stages is
presently well established and internationally accepted for the pre-
Pleistocene part (Table 1). New ICS task groups have been orga-
nized under the umbrella of the Subcommission on Quaternary
Stratigraphy to establish an international Pleistocene subdivision of
Lower, Middle and Upper, and to define the Holocene/Pleistocene
June 2004
96
Figure 4 Proposal for a “natural” Precambrian time scale. Earth history is divided into six eons, with boundaries defined by what can be
considered first-order key events in the evolution of our planet.
boundary. GSSPs have been formalized for the Aquitanian (defining
the Paleogene/Neogene boundary), Tortonian and Messinian stages
of the Miocene, and for the Zanclean, Piacenzian and Gelasian
stages of the Pliocene. In addition, the Pliocene-Pleistocene bound-
ary has been defined.
From the 1970’s until 1994, Neogene time scales were con-
structed using a limited number of radio-isotopic age calibration
points in geomagnetic polarity sequences that were primarily
derived from a seafloor anomaly profile in the south Atlantic, modi-
fied after Heirtzler et al. (1968). Biozonations and stage boundaries
were subsequently tied to the resulting geomagnetic polarity time
scale (GPTS), preferably via magneto-biostratigraphic calibrations
(Berggren et al., 1985). Alternatively, radio-isotopic age determina-
tions from both sides of stage boundaries were used to calculate a
best-fit radio-isotopic age estimate for these boundaries in a statisti-
cal way (chronogram method of Harland et al., 1982, 1990).
The “standard” method to construct time scales changed drasti-
cally with the advent of the astronomical dating method to the pre-
late Pleistocene. This method relies on the calibration, or tuning, of
sedimentary cycles or cyclic variations in climate proxy records to
target curves derived from astronomical solutions for the solar-plan-
etary and Earth-Moon systems. Quasi-periodic perturbations in the
shape of the Earth’s orbit and the tilt of the inclination axis are
caused by gravitational interactions of our planet with the Sun, the
Episodes, Vol. 27, no. 2
97
Figure 5a Neogene stratigraphic subdivisions, geomagnetic polarity scale, pelagic zonations and selected datums of planktonic
foraminifers and calcareous nannoplankton. Main trends in eustatic sea level are generalized. The “Quaternary”, shown schematically on
the right-hand side, is traditionally considered to be the interval of oscillating climatic extremes (glacial and interglacial episodes) that was
initiated at about 2.6 Ma, therefore encompassing the Holocene and Pleistocene epochs and Gelasian stage of late Pliocene. The
Quaternary composite epoch is not a formal unit in the chronostratigraphic hierarchy.
Moon and the other planets of our solar system. These interactions
give rise to cyclic changes in the eccentricity of the Earth's orbit,
with main periods of 100,000 and 413,000 years, and in the tilt
(obliquity) and precession of the Earth's axis with main periods of
41,000, and 21,000 years, respectively (Berger, 1977). These pertur-
bations in the Earth's orbit and rotation axis are climatically impor-
tant because they affect the global, seasonal and latitudinal distribu-
tion of the incoming solar insolation. Orbital forced climate oscilla-
tions are recorded in sedimentary archives through changes in sedi-
ment properties, fossil communities, chemical and isotopic charac-
teristics. While Earth scientists can read these archives to reconstruct
paleoclimate, astronomers have formulated models based on the
mechanics of the solar-planetary system and the Earth-Moon system
to compute the past variations in precession, obliquity and eccentric-
ity of the Earth’s orbit and rotation axis. As a logical next step, sed-
imentary archives can be dated by matching patterns of paleoclimate
variability with patterns of varying solar energy input computed
from the astronomical model solutions. This astronomical tuning of
the sedimentary record results in time scales based on measurable
physical parameters that are independent from those underlying
radio-isotopic dating and that are tied to the Recent through a direct
match with astronomical curves.
Astronomical tuning was first applied in the late Pleistocene in
order to build a common high-resolution time scale for the study of
orbital induced glacial cyclicity. Initial attempts to extend this time
scale back in time were unsuccessful due to lack of resolution or
incompleteness of the sedimentary succession. These problems were
overcome with the advent of the advanced piston corer (APC) tech-
nique in ocean drilling and the drilling of multiple offset holes per
site. Combined these innovations were used to construct spliced
June 2004
98
Figure 5b Neogene dinoflagellate cyst and radiolarian zonation with estimated correlation to magnetostratigraphy and planktonic
foraminifer zones.
composite sections in order to recover undisturbed and complete
successions marked by high sedimentation rates. Soon afterwards,
the astronomical time scale was extended to the base of the Pliocene
based on ODP sites (Shackleton et al., 1990) and land-based sections
in the Mediterranean (Hilgen, 1991a,b), the study of the latter pro-
viding another means to overcome the problem of incompleteness of
the stratigraphic record.
GTS2004 for the first time presents an Astronomically Tuned
Neogene Time Scale (ATNTS2004), based on cyclic sedimentary
successions from the western Equatorial Atlantic Ocean and
Mediterranean. The new time scale represents a continuation of a
development that led Berggren et al. (1995a) to incorporate the
Pliocene and Pleistocene astrochronology of Shackleton et al. (1990)
and Hilgen (1991a, b) in their Neogene time scale.
Construction of the new high-resolution Neogene time scale
was made possible through:
(1) Technological and procedural improvements in deep-sea
drilling of older Neogene strata,
(2) High-resolution studies of exposed marine sections in tec-
tonically active areas where ancient seafloor has been rapidly
uplifted, and
(3) Improvements in the accuracy of theoretical astronomical
solutions resulting in the La2003 numerical solution.
A seafloor anomaly profile from the Australia-Antarctic plate
pair was employed to complete the polarity time scale for the inter-
val between 13 and 23 Ma due to the lack of magnetostratigraphic
records for ODP Leg 154 sites. Biostratigraphic zonal schemes are
either directly tied to the new time scale via first-order calibrations,
such as the standard low-latitude calcareous plankton zonation, or
can be linked to it by recalibrating them to the associated polarity
time scale. Formally designated chronostratigraphic boundaries
(GSSPs of Neogene stages) are also directly tied to the new time
scale because they are defined in sections that have been used to
build the astronomically tuned integrated stratigraphic framework
that underlies the time scale. An overview of the tuned Neogene
stratigraphic framework is in Figures 5a and b.
The new time scale resulted in a significantly younger age of
23.03 Ma for the Oligocene-Miocene boundary than the 23.8 Ma
estimated in previous time scales; the latter age was based on radio-
metric age determinations that are not fully acceptable according to
current standards. The intercalibration of the independent astronom-
ical and radiogenic-isotopic dating methods is not yet solved, but
new results (Kuiper, 2003) point to an astronomical-derived age of
28.24 ±0.01 Ma for the Fish Canyon Tuff (FCT) sanidine and favor
the introduction of a directly astronomically dated standard in 40Ar-
39Ar dating.
The astronomically tuned Neogene time scale with an unprece-
dented accuracy (1–40 kyr) and resolution (<10 kyr), opens new per-
spectives for paleoclimatic and paleooceanographic studies of the
entire Neogene with a temporal resolution comparable to that of
Pleistocene research (i.e., Krijgsman et al., 1999; Zachos et al.,
2001).
GTS Quo Vadis?
The changing philosophy in time scale building has made it more
important to undertake high-resolution radiometric study of critical
stratigraphic boundaries, and extend the astronomical tuning into
progressively older sediments. Good examples are Bowring et al.
(1989) for basal-Triassic, Amthor et al. (2003) for basal-Cambrian
and Hilgen et al. (2000) for Messinian. The philosophy is that
obtaining high-precision age dating at a precisely defined strati-
graphic boundary avoids stratigraphic bias and its associated uncer-
tainty in rock and in time.
In this respect, it is of vital importance that ICS not only com-
pletes the definition of all stage boundaries, but also actively consid-
ers definition of standardized subdivisions within the many long
stages itself. Examples of long stages (spanning more than 10 myr)
that lack international standardization of internal divisions are the
Campanian, Albian, Aptian, Norian, Carnian, Sakmarian, Visean,
Tournaisian, Famennian and Tremadocian stages, and parts of the
Cambrian system. This consensus definition process should be com-
pleted in a timely manner. Regional and philosophical arguments
between stratigraphers should be actively resolved to reach consen-
sus conclusions which focus on the global correlation implications.
Stratigraphic standardization precedes linear time calibration.
Future challenges to time scale building, detailed in Gradstein
et al. (2004), may be summarized as follows:
(1) Formal definition of all Phanerozoic stage boundaries, and
interior definition of long stages.
(2) Orbital tuning of polarity chrons and biostratigraphic events
for the entire Cenozoic and Cretaceous (past 150 myr).
(3) A consensus Ar/Ar monitor age (? 28.24 ±0.01 Ma from
orbital tuning), and consensus values for decay constants in the K-Ar
isotope family.
(4) A detailed public database of high-resolution radiometric
ages that includes “best practice” procedures, full error propagation,
monitor ages and conversions.
(5) Resolving of zircon controversies across Devonian/Car-
boniferous, Permian/Triassic, and Anisian/Ladinian boundaries,
either through more sampling or re-evaluation of different laboratory
techniques.
(6) Detailed age dating of several ‘neglected’ intervals, includ-
ing Upper Jurassic–Lower Cretaceous (M-sequence spreading and
‘tuned’ stages), base Carboniferous (Kellwasser extinction event;
glaciation), and within Albian, Aptian, Norian, Carnian, Visean, and
intra Permian.
(7) More detailed composite standard zonal schemes for Upper
Paleozoic and Lower Mesozoic.
(8) On-line stratigraphic databases and tools (e.g., a rapid
expansion of the CHRONOS network).
The geochronological science community and ICS are focusing
on these challenging issues. The next version of the Geologic Time
Scale is planned for the 33rd IGC in 2008, concurrent with the
planned completion of boundary-stratotype (GSSP) definitions for
all international stages.
References
Allègre, C.J., Manhès, G., and Göpel, C., 1995, The age of the Earth;
Geochimica et Cosmochimica Acta, 59 (8), p. 1445-1456.
Amthor, J. E., Grotzinger, J.P., Schroder, S., Bowring, S.A., Ramezani, J.,
Martin, M.W., and Matter, A., 2003, Extinction of Cloudina and Namaca-
lathus at the Precambrian boundary in Oman. Geology, 31 (5), p. 431-
434.
Berger, A., 1977, Long term variations of the Earth's orbital elements: Celes-
tial Mechanics, 15, p. 53-74.
Berggren, W.A., Kent, D.V., and Van Couvering, J.A., 1985, The Neogene:
Part 2, Neogene geochronology and chronostratigraphy, in Snelling, N.J.,
ed., The Chronology of the Geological Record: Geological Society of
London Memoir 10, p. 211-250.
Berggren, W.A., Hilgen, F.J., Langereis, C.G., Kent, D.V., Obradovitch,
J.D., Raffi, I., Raymo, M., and Shackleton, N.J., 1995, Late Neogene
(Pliocene-Pleistocene) chronology: New perspectives in high-resolution
stratigraphy: Geological Society of America Bulletin, 107, p. 1272-1287.
Blake, T.S., and Groves, D.I., 1987,Continental rifting and the Archean-Pro-
terozoic transition. Geology, 15, p. 229-232.
Bleeker, W., 2003a, Problems with the Precambrian timescale: from accre-
tion to Paleoproterozoic plate break-up. See http://www.nunatime.ca.
Bleeker, W., 2003b, The late Archean record: a puzzle in ca. 35 pieces.
Lithos, 71 (2/4), p. 99-134.
Bowring, S.A., Ramezani, J., and Grotzinger, J.P., 2003, High-precision U-
Pb zircon geochronology and Cambrian-Precambrian boundary. See
http://www.nunatime.ca.
Cande, S.C., and Kent, D.V., 1992, A new geomagnetic polarity time scale
for the Late Cretaceous and Cenozoic. Journal of Geophysical Research,
97, p. 13917-13951.
Cande, S.C. and Kent, D.V., 1995, Revised calibration of the geomagnetic
polarity timescale for the Late Cretaceous and Cenozoic. Journal of Geo-
physical Research, 100, p. 6093-6095.
Episodes, Vol. 27, no. 2
99
Cloud, P., 1972, A working model of the primitive Earth. American Journal
of Science, 272, p. 537-548.
Cloud, P., 1987, Trends, transitions, and events in Cryptozoic history and
their calibration: apropos recommendations by the Subcommission on
Precambrian Stratigraphy. Precambrian Research, 37, p. 257-264.
Cooper, R.A., 1999, The Ordovician time scale - calibration of graptolite and
conodont zones: Acta Universitatis Carolinae Geologica, 43 (1/2), p. 1-4.
Crook, K.A.W., 1989, Why the Precambrian time-scale should be chronos-
tratigraphic: a response to recommendations by the Subcommittee on
Precambrian Stratigraphy. Precambrian Research, 43, p. 143-150.
Gradstein, F.M., Agterberg, F.P., Ogg, J.G., Hardenbol, J., van Veen, P.,
Thierry, T., and Huang, Z., 1994, A Mesozoic time scale. Journal of Geo-
physical Research, 99 (B12), p. 24051-24074.
Gradstein, F.M., Ogg, J.G., Smith, A.G., Agterberg, F.P., Bleeker, W.,
Cooper, R.A., Davydov, V., Gibbard, P., Hinnov, L.A., House, M.R. (†),
Lourens, L., Luterbacher, H-P., McArthur, J., Melchin, M.J., Robb, L.J.,
Shergold, J., Villeneuve, M., Wardlaw, B.R., Ali, J., Brinkhuis, H.,
Hilgen, F.J., Hooker, J., Howarth, R.J., Knoll, A.H., Laskar, J., Monechi,
S., Powell, J., Plumb, K.A., Raffi, I., Röhl, U., Sanfilippo, A., Schmitz,
B., Shackleton, N.J., Shields, G.A., Strauss, H., Van Dam, J., Veizer, J.,
van Kolfschoten, Th., and Wilson, D., 2004 (in press), A Geologic Time
Scale 2004. Cambridge University Press, ~500 pp
Guest, J.E., and Greeley, R., 1977. Geology on the Moon; The Wykeham Sci-
ence Series, Crane, Russak and Company, Inc., New York, 235 pp.
Harland, W.B., Cox, A.V., Llewellyn, P. G., Pickton, C.A.G., Smith, A.G.,
and Walters, R., 1982, A geologic time scale 1982, Cambridge University
Press, 131 pp.
Harland, W.B., Armstrong, R.L., Cox, A.V., Craig, L.E., Smith, A.G., and
Smith, D.G., 1990, A geologic time scale 1989. Cambridge University
Press, 263 pp.
Heirtzler, J.R., Dickson, G.O., Herron, E.M., Pitman, W.C., and Le Pichon,
X., 1968, Marine magnetic anomalies, geomagnetic field reversals, and
motions of the ocean floor and continents. Journal of Geophysical
Research, 73, p. 2119-2139.
Hilgen , F.J., 1991a, Extension of the astronomically calibrated (polarity)
time scale to the Miocene/Pliocene boundary. Earth and Planetary Sci-
ence Letters, 107, p. 349-368.
Hilgen, F.J., 1991b, Extension of the astronomically calibrated (polarity)
time scale to the Miocene-Pliocene boundary. Earth and Planetary Sci-
ence Letters, 107, p. 349-368.
Hilgen, F.J., Krijgsman, W., Langereis, C.G., Lourens, L.J., Santarelli, A.,
and Zachariasse, W.J., 1995, Extending the astronomical (polarity) time
scale into the Miocene. Earth and Planetary Science Letters, 136, p. 495-
510.
Hilgen, F.J., Bissoli, L., Iaccarino, S., Krijgsman, Meijer, R., Negri, A., and
Villa, 2000, Integrated stratigraphy and astrochronology of the Messinian
GSSG at Oued Akrech (Atlantic Morocco). Earth and Planetary Science
Letters, 182, p. 237-251.
Holmes, A., 1947, The construction of a geological time-scale. Transactions
Geological Society of Glasgow, 21, p. 117-152.
Holmes, A., 1960, A revised geological time-scale. Transactions of the Edin-
burgh Geological Society, 17, p. 183-216.
Krijgsman, W., Hilgen, F.J., Raffi, I., Sierro, F.J., and Wilson, D.S., 1999,
Chronology, causes and progression of the Messinian salinity crisis:
Nature, 400, p. 652-655.
Kuiper, K.F., 2003, Direct intercalibration of radio-isotopic and astronomical
time in the Mediterranean Neogene: Geologica Ultraiectina (Mededelin-
gen van de Faculteit Geowetenschappen, Universiteit Utrecht), 235, 223
pp.
Ludwig, K.R., 2000, Decay constant errors in U-Pb Concordia-intercept
ages. Chemical Geology, 166, p. 315-318.
Lumbers, S.B., and Card, K.D., 1991, Chronometric subdivision of the
Archean. Geology, 20, p. 56-57.
Murray, B., Malin, M.C., and Greeley, R., 1981. Earthlike planets; surfaces
of Mercury, Venus, Earth, Moon, Mars; W.H. Freeman and Company,
San Francisco, 387 pp.
Nisbet, E.G., 1991, Of clocks and rocks—The four aeons of Earth. Episodes,
14, p. 327-331.
NUNA, 2003, New Frontiers in the fourth dimension: generation, calibration
and application of geological timescales; NUNA Conference, Geological
Association of Canada; Mont Tremblant, Quebec, Canada, March 15-18,
2003. See http://www.nunatime.ca.
Obradovich, J.D., 1993, A Cretaceous time scale, in Caldwell, W.G.E., and
Kauffman, E.G., eds., Evolution of the Western Interior Basin, Geologi-
cal Association of Canada, Special Paper 39, p. 379-396.
Plumb, K.A., 1991, New Precambrian time scale. Episodes, 14, p. 139-140.
Plumb, K.A., and James, H.L., 1986, Subdivision of Precambrian time: Rec-
ommendations and suggestions by the commission on Precambrian
stratigraphy. Precambrian Research, 32, p. 65-92.
Remane, J., 2000, International Stratigraphic Chart, with Explanatory Note.
Sponsored by ICS, IUGS and UNESCO. (distributed at the 31st Interna-
tional Geological Congress, Rio de Janeiro 2000), 16 pp.
Shackleton, N.J., Berger, A., and Peltier, W.R., 1990, An alternative astro-
nomical calibration of the lower Pleistocene timescale based on ODP site
677. Transactions of the Royal Society of Edinburgh, 81, p. 251-261.
Shackleton, N.J., Crowhurst, S.J., Weedon, G.P., and Laskar, J., 1999, Astro-
nomical calibration of Oligocene-Miocene time. Philosophical Transac-
tions of the Royal Society of London, A, (357), p. 1907-1929.
Spudis, P.D., 1999. The Moon. In: The New Solar System, edited by J.K.
Beatty, C. Collins Petersen and A. Chaikin, Cambridge University Press,
Cambridge, p. 125-140.
Trendall, A.F., 1991, The “geological unit” (g.u.)—A suggested new mea-
sure of geologic time. Geology, 19, p. 195.
Windley, B.F., 1984, The Archaean-Proterozoic boundary. Tectonophysics,
105, 43-53.
Zachos, J., Pagani, M., Sloan, L., Thomas, E., and Billups, K., 2001, Trends,
rhythms, and aberrations in global climate 65 Ma to present. Science,
292, p. 686-693.
June 2004
100
Felix M. Gradstein is chair of the
International Commission on Stratig-
raphy. Following retirement from the
Geological Survey of Canada and
Saga Petroleum Norway, he joined
the Natural History Museum, Univer-
sity of Oslo as stratigraphy/micropa-
leontology professor, where he is
developing relational stratigraphic
databases for offshore Norway. His
activities have included quantitative
stratigraphy (he chaired previous
IGCP and IUGS programs), Ocean
Drilling Program legs in the Atlantic
and Indian oceans, and coordinating
compilation of Mesozoic and
Phanerozoic geologic time scales.
He is an avid skier and offshore
sailor.
Jim Ogg, a professor of stratigraphy
at Purdue University in Indiana
USA, has been serving as Secretary-
General of the International Com-
mission on Stratigraphy of IUGS
since 2000. His research concen-
trates on the Mesozoic and Paleo-
gene, especially paleoceanography
(including ten DSDP-ODP drilling
cruises), time scales of cyclic sedi-
mentation and magnetic polarity
chrons, and integrated Earth his-
tory. Gabi Ogg, his wife and fellow
stratigrapher, was responsible for
most of the graphics on the ICS web-
site and in the GTS2004 book.
... Comparison of age spectra on samples from the ASRR shear system. The numerical and stage time scales are those of Gradstein et al. (2004). Numbers on data points refer to the following sources: 1 L. S. Zhang and Schärer (1999) Leloup et al. (2001), p. 18 Leloup and Kienast (1993), p. 19 X. Y. Chen et al. (2015), p. 20 F. L. Liu et al. (2013), p. 21 Harrison et al. (1992, p. 22 Harrison et al. (1996), p. 23 P. L. Wang et al. (1998), p. 24 P. L. Wang et al. (2000), p. 25 Maluski et al. (2001), p. 26 Cao, Neubauer, et al. (2011), p. 27 Guo et al. (2016, p. 28 Wan et al. (1997) Tables 1 and 2 ( Figure 5). ...
... Comparison of age spectra on samples from the ASRR shear system. The numerical and stage time scales are those ofGradstein et al. (2004). Numbers on data points refer to the following sources: 1 L. S.Zhang and Schärer (1999), p. 2 Cao, Liu, et al. (2011), p. 3 Qi et al. (2014), p. 4 Zhao et al. (2014), p. 5 Lin et al. (2012), p. 6 Schärer et al. (1994), p. 7 Tang, Liu, et al. (2013, p. 8 J. L. Liu et al. (2015), p. 9 Sassier et al. (2009), p. 10 B. L. Li et al. (2014), p. 11 Chung et al. (1997), p. 12 Palin et al. (2013), p. 13 Anczkiewicz et al. (2012), p. 14 Anczkiewicz and Viola (2003), p. 15 F. L. Liu et al. (2015), p. 16 B. Zhang et al. (2014), p. 17 Leloup et al. (2001), p. 18 Leloup and Kienast (1993), p. 19 X. Y. Chen et al. (2015), p. 20 F. L. Liu et al. (2013), p. 21 Harrison et al. (1992), p. 22 Harrison et al. (1996), p. 23 P. L. Wang et al. (1998), p. 24 P. L. Wang et al. (2000), p. 25 Maluski et al. (2001), p. 26 Cao, Neubauer, et al. (2011), p. 27 Guo et al. (2016), p. 28 Wan et al. (1997), p. 29 Q. ...
Article
Full-text available
The response mechanism of the southeastern Tibet Plateau to the Indian-Eurasian collision is still controversial. The widely distributed shear systems in the Sanjiang area especially Ailaoshan-Red River shear system contribute a window to understand the tectonic evolution. Here we compile comprehensive geochronological data of the high-grade metamorphic rocks from the Ailaoshan-Red River shear system. The data reveal that the Ailaoshan metamorphic belt should be metamorphic complex with Middle-Late Triassic magmatic rocks and Cenozoic granites besides the Precambrian basement. The Cenozoic U-Pb age results of granitic rocks by accessory mineral within the Ailaoshan-Red River shear system range from ~41 Ma to ~20 Ma. The results of Ar-Ar dating of various minerals are in the range of ~40–5 Ma, concentrated in ~35–20 Ma. The U-Pb age results coincide basically with Ar-Ar data, indicating a rapid cooling process and reflecting a possible tectonic transition at ~20 Ma. The metamorphism and magmatism since the Cenozoic were not only unique to the Ailaoshan metamorphic belt, but also existed in the Chongshan and Gaoligongshan metamorphic belts. We infer that the Sanjiang area has a similar lateral extrusion deformation background and corresponds to the response of the India-Eurasia convergence process.
... The keywords "name-of-continent, fossil pollen" were also used to identify suitable articles. The bounds of the geological stages were accepted as given in Gradstein et al. (2004) and midpoints used for plotting purposes. The extant affiliations of pollen (palynomorphs) and their hypothetical ancestors were obtained from the Supplementary section of Sauquet et al. (2009) and other specialist papers and their expected time of origin obtained from our chronogram. ...
... Dalziel et al. (2013) considered that the westward drift of South America during the mid-Cretaceous resulted in extension of the Andes Cordillera, obduction of the Rocas Verdes Basin and closure of Drake's Passage. Although no precise dates are given by them, the mid-Cretaceous lies at 99.6 Ma (Gradstein et al., 2004). Leppe et al. (2012) stated that closure was in the Turonian (93.6-88.6 ...
... A lone Orbulina universa marker at 2880 ft marks this zone at the top while the base is defined by the top of the Globigerinoides trilobus immaturus marker at 2970 ft The Zone correlates with the N9 planktonic foraminiferal Zone of Caron (1985), Berggren et al. (1998), Hardenbol et al. (1998 and Gradstein et al. (2004), as well as the N9a zone of Boudagher-fadel (2015). The age is Early to Middle-Miocene. ...
Article
Full-text available
Late Cretaceous-Miocene foraminiferas were recovered from the BG-1 well in the offshore Eastern Dahomey Basin. Four (4) lithostratigraphic units comprising Ogwashi-Asaba, Upper Araromi, Lower Araromi and Afowo Formations have been delineated and assigned based on the textural characteristics of the sediments. Forty-nine (49) foraminiferal species were identified, with a total count of 47 calcareous species (96%) comprising both planktonic and benthic forms and two arenaceous forms (3%). Seven (7) foraminiferal zones were recognised and dated from the Upper Cenomanian to Late Miocene age. The Globotruncana aegyptica zone, marked by the disappearance of the Maastrichtian forms at 3060 ft, coincided with the appearance of the Early Paleocene benthic forms such at the Anomalinoides umboniferus- Anomalinoides midwayensis zone. The Cretaceous-Paleogene (K-Pg) boundary was recognised at 3060 ft. Seven MFSs and SBs dated from the Upper Cenomanian and Upper Miocene ages were identified in the sequence stratigraphic analysis. The stacking patterns of the lowstand and highstand systems tracts reveal the interplay of progradational and aggradational parasequence signatures of siltstones and sandstone lithologies. The paleodepositional sedimentary packages of the BG-1 well are recognised from the Inner-Neritic to Bathyal environment. The sequence stratigraphic integration of lithofacies and foraminifera assemblages in this study has created a model of the distribution of the elements in the hydrocarbon system of the offshore Eastern Dahomey basin. Therefore, this study will underscore the critical role of sequence biostratigraphy in enhancing the accuracy and efficiency of exploration and development efforts in the hydrocarbon industry.
... Botryococcus has been extensively documented within the fossil record, notably in Precambrian oil shales (Gradstein et al. 2004), where it has been considered the primary contributor to crude oil. In addition, Botryococcus fossils have also been found in the Carboniferous sequence of France (Renault 1894) and in the Upper Carboniferous-Permian deposits of Rotliegend in central Europe (Bertrand and Renault 1892a b). ...
... Botryococcus has been extensively documented within the fossil record, notably in Precambrian oil shales (Gradstein et al. 2004), where it has been considered the primary contributor to crude oil. In addition, Botryococcus fossils have also been found in the Carboniferous sequence of France (Renault 1894) and in the Upper Carboniferous-Permian deposits of Rotliegend in central Europe (Bertrand and Renault 1892a b). ...
Article
Full-text available
The influence of marine components in the study area is attributed to seafloor spreading and opening in the Indian Ocean during the Aptian, coupled with eustatic sea-level rise throughout the Aptian–Albian period. The present investigation delineates a marine incursion in the Chintalapudi sub-basin of the Godavari Valley Coalfield based on the occurrence of dinocysts, acritarchs and prasinophytes. Noteworthy stratigraphic marker taxa are recognized as key components in the age assignment of the sedimentary rocks. Sum total of four zones have been recognised in the studied part of the sequence. The oldest zone in the deepest set of sedimentation is identified as Cicatricosisporites australiensis Zone. This zone is assigned based on the first occurrence (FO) of Cicatricosisporites australiensis and Biretisporites eneabbaensis. The next zone, further above in the sequence is Foraminisporis wonthaggiensis Zone. The FO of Foraminisporis wonthaggiensis attested to the initiation of this zone. The third zone is Cyclosporites hughesii Zone recognised by the FO of Foraminisporis asymmetricus. The accessory palynotaxa comprised of Retimonocolpites sp., Tricolpites confessus, Tricolpites sp., Tricolporites sp. and Triporites sp. in the recognition of this zone. The youngest zone of this sequence is recognised at the top of the sequence named Crybelosporites striatus zone based on the FO of Crybelosporites striatus. Palynological analysis places the studied part of succession between the Berriasian to early Albian. The primary objective of this investigation is to explicate the characteristics of sedimentation and paleoenvironmental contexts, with a specific emphasis on utilizing palynofacies and Botryococcus morphotypes as distinct benchmarks. Based on the palynofacies investigation three different depositional regimes have been assigned, which are distinguished as follows: (a) continental (384.30–203.75 m) dysoxic–anoxic basin; (b) shallowest nearshore–inner shelf, marginal marine/brackish water settings; and (c) heterolithic proximal shelf deposition with significant sea transgression (94.50–36.50 m). Unpredictable morphotaxonomical responses across diverse habitats with varying pH levels are projected to enhance comprehension of the palaeoecological and palaeoenvironmental significance of fossil Botryococcus colonies. The uppermost marine part of the sequence denotes the least abundance, reduced size, and ill-defined amorphous forms of fossil Botryococcus colonies. The continental part displays the less abundant but well-defined and fully grown sporadic distribution of fossil Botryococcus colonies; the middle part records its presence with a well-diversified, sturdy structural framework and cellular cup-like structures. The investigation of the Cretaceous strata reveals a dynamic geological history influenced by tectonic activities, including plate drifting and rifting. Present investigation enhances our comprehension of palynostratigraphy, and paleoenvironmental history and facilitates the formulation of models elucidating environmental processes during the Early Cretaceous. In addition, the lipid-rich composition of Botryococcus renders it promising for biofuel production, attracting attention in biotechnology and renewable energy research endeavors.
... The South Atlantic was divided into four segments from south to north-the Falkland, austral, central, and equatorial segments ( Fig. 1a) Biari et al., 2021). The austral segment opened between the anomalies M13 and M4 (Rabinowitz and LaBrecque, 1979;Austin and Uchupi, 1982;Curie, 1984;Moulin et al., 2010), corresponding to 139.5 and 130 Ma, respectively (Gradstein et al., 2004). The seafloor spreading of the central segment is generally regarded to begin after 18 Myr, when the evaporite sedimentation ended (Upper Aptian-Lower Albian: around 112 Ma) (Davison, 1999;Mohriak, 2001;Torsvik et al., 2009;Moulin et al., 2005Moulin et al., , 2010 although there is an ongoing dispute for the opening time. ...
Article
Gravity inversion is a highly effective method for investigating regional geological structures, and this paper proposes an optimization scheme for constrained three-dimensional (3D) gravity inversion to obtain a 3D density model, utilizing prior geological and geophysics information. Specifically, the proposed method enhances deep structural imaging resolution and minimizes false structures by progressively inverting deep and shallow-density structures using long and short-wavelength signals of gravity anomaly with prior information. The scheme is applied in the southeast passive continental margin of Brazil, and the results show that the density model is consistent with the previous reflection and refraction seismic data. Moreover, the 3D density model reveals several insights: (i) The Abimael Ridge (AR) and the S˜ao Paulo Plateau (SPP) exhibit thin crustal thickness (~5–7 km) indicative of proto-oceanic crust. The SPP and AR area crustal thinning may be related to an aborted opposing rift propagator pair. (ii) The rifting modes of Santos and Campos Basins differ significantly. Campos Basin exhibits a depth-dependent lithospheric stretching model with a relatively intact upper crust. In contrast, Santos Basin shows a highly brittle upper crust that is partially thinned and, in some regions, even absent under the far-field effect of spreading failed rifts, while the lower crust remains relatively intact. Moreover, the upper crustal stretching factor is about five to ten higher than the lower crustal. Thus, the constrained 3D gravity inversion scheme provides a new avenue for continental rifted margin geological structure studies.
Preprint
Full-text available
RESUMO. – Chimpanzés e seres humanos derivaram de um mesmo ancestral comum. Evidências fósseis e moleculares indicam que o ramo ancestral se dividiu em dois há 5-8 milhões de anos. De um lado, prosperou a linhagem que daria origem aos chimpanzés; de outro, a que daria origem aos humanos. As duas linhagens se diferenciaram e se afastaram, dando origem a múltiplas ramificações e a diferentes estilos de vida. Os ancestrais dos chimpanzés continuaram a viver em florestas fechadas, enquanto os nossos ancestrais desceram das árvores e passaram a viver em hábitats mais abertos, dominados por gramíneas e outras plantas herbáceas. SUMMARY. – (Coming down from the trees: Climate, habitat, niche and the origins of humanity.) Chimpanzees and humans derived from the same common ancestor. Fossil and molecular evidence indicates that the ancestral branch split in two 5-8 million years ago. On the one hand, the lineage that would give rise to chimpanzees prospered; on the other, the lineage that would give rise to humans. The two lineages differentiated and moved apart, giving rise to multiple branches and different lifestyles. Chimpanzee ancestors continued to live in closed forests, while our ancestors descended from the trees and began to live in more open habitats dominated by grasses and other herbaceous plants.
Article
Full-text available
Sulawesi Island, known for its geological complexity, offers a unique opportunity for researchers to investigate further into sedimentary basins particularly in South Sulawesi area. The Cretaceous Balangbaru Shale is one of the rocks that has an extensive distribution in South Sulawesi especially in the western part. Due to limitation of geological information related to this formation we performed the paleodepositional investigation to elucidate the paleoweathering, geochemistry characterization, and detrital provenance of the Balangbaru Shale. Indeed, previous studies have demonstrated that fine-grained sedimentary rocks can serve as valuable archives of geochemical information, enabling the reconstruction of paleoclimate evolution during the deposition period. Employing XRF quantitative analysis, we examine the dispersion of major elements within eight distinct layers of the Shale. Additionally, comprehensive whole-rock analyses enable us to ascertain the abundance of major elements and subsequently deduce detrital SiO 2 and biogenic SiO 2 content, thereby comprehending the inorganic geochemical composition of the Balangbaru Shale. The present study seeks to address the limited understanding of the geochemical properties of the Balangbaru Shale, specifically pertaining to paleoweathering intensity and detrital provenance. Acknowledging the scarcity of recent data available on the subject matter, this research aims to bridge the limitation data gap by conducting an in-depth investigation into the geochemical characteristics of the Balangbaru Shale by employing major oxides concentrations (SiO 2 , Al 2 O 3 , Fe 2 O 3 , K 2 O, P 2 O 5 , MgO, CaO, Na 2 O, MnO, TiO 2 , Cr 2 O 3 ) and trace elements (Sr, Cu, Ba, Th, Ni, V, Th, U) into the equations such as C-value, chemical index alteration (CIA), plagioclase index of alteration (PIA), weathering index of parker (WIP) and ratio of Rb/Sr, U/Th, Sr/Cu, Sr/Ba. Furthermore, the major oxides ratio of Al 2 O 3 /TiO 2 and TiO2/Zr were performed to specify the sediment provenance and the source-area rock compositions of the Balangbaru Shale. The obtained data reveals a moderate to high degree of weathering intensity in the detrital source environment of the Balangbaru Shale, suggesting warm to humid climate conditions in the source region. Additionally, the analysis of the detrital fraction indicates the possible origin of the Balangbaru Shale from felsic rocks. Overall, these findings contribute significantly to the understanding of paleoenvironmental conditions associated with the Balangbaru Shale, shedding light on its geological history.
Article
Full-text available
The Messinian salinity crisis is widely regarded as one of the most dramatic episodes of oceanic change of the past 20 or so million years (refs 1-3). Earliest explanations were that extremely thick evaporites were deposited in a deep and desiccated Mediterranean basin that had been repeatedly isolated from the Atlantic Ocean, but elucidation of the causes of the isolation - whether driven largely by glacio-eustatic or tectonic processes - have been hampered by the absence of an accurate time frame. Here we present an astronomically calibrated chronology for the Mediterranean Messinian age based on an integrated high-resolution stratigraphy and 'tuning' of sedimentary cycle patterns to variations in the Earth's orbital parameters. We show that the onset of the Messinian salinity crisis is synchronous over the entire Mediterranean basin, dated at 5.96 ± 0.02 million years ago. Isolation from the Atlantic Ocean was established between 5.59 and 5.33 million years ago, causing a large fall in Mediterranean water level followed by erosion (5.59-5.50 million years ago) and deposition (5.50- 5.33 million years ago) of non-marine sediments in a large 'Lago Mare' (Lake Sea) basin. Cyclic evaporite deposition is almost entirely related to circum- Mediterranean climate changes driven by changes in the Earth's precession, and not to obliquity-induced glacio-eustatic sea-level changes. We argue in favour of a dominantly tectonic origin for the Messinian salinity crisis, although its exact timing may well have been controlled by the ~400-kyr component of the Earth's eccentricity cycle.
Article
A new data base now has been created using the latest developments in geochronometry, the 40Ar/39Ar laser fusion approach. Re-evaluation of many of the fossil collections from the Western Interior and from Europe has revealed the presence of common faunal elements, which now permit precise correlation of many of the Western Interior biozones with those of the type sections of Europe. Coupling the new geochronometry with this improved biozonal framework based on ammonites has allowed the time scale to be compiled. It is particularly applicable to the Late Cretaceous. -from Author
Article
The time-scale constructed in 1947 was based on certain assumptions that have recently been shown to be wrong. Appalachian pegmatites dated at 350 million years (m.y.) and thought to be Taconic (Ordovician) are now found to be Acadian (late Devonian), while others, dated at 255 m.y. and thought to be Acadian can now be referred to the Permian. Other recent evidence consistently leads to an extension of the 1947 scale that carries the beginning of the Cambrian back to about 600 m.y. ago. The scale now constructed from the data available up to October, 1959, is as follows (in m.y.): One of the more significant consequences of this revision is that many dated rocks from Africa and the other “Gondwanaland” continents which were formerly ascribed to the late Precambrian now become Cambrian. Important orogenic and plutonic phases of a major geological cycle, implying by analogy an extensive system of geosynclines, occurred at about the close of the Precambrian and early in the Ordovician.
Article
Lithological cyclicity was observed aboard the JOIDES RESOLUTION in sediment sequences recovered from the Ceara Rise durin ODP Leg 154. Shipboard work led to the conclusion that the Oligocene was probably characterized by ca. 41 ka cycles. Weedon and others were able to confirm this, and created a provisional time–scale for the Oligocene by assumin that the cyclicity is a response to orbital obliquity variation, and by using spectral analysis to estimate the mean wavelengt and hence the sedimentation rate of successive intervals of core. We have extended this work by intercorrelating almost al the 9.5 m sediment cores from each of the four sites that recovered Oligocene sediment. We have successfully correlated al the material covering a time–interval of ca. 10 Ma from 18 Ma to 28 Ma, as well as most of the sediment from the 14 to 18 Ma and 28 to 34 Ma intervals. Although variabilit is dominated by the 41 ka cycle there is sufficient variability at the precession period (amplitude–modulated by eccentricity to permit an absolute placement of this section with reference to the calculated orbital history. Further work is needed t establish precisely the implications of this calibration for the geological time–scale but it appears that the true ages o events close to the Oligocene–Miocene boundary are ca. 0.9 Ma younger than they appear on recently published time–scales. The sedimentary record preserves information concernin the amplitude modulation of the obliquity signal that is of astronomical as well as geological significance.
Article
A Geologic Time Scale (GTS2004) is presented that integrates currently available stratigraphic and geochronologic information. The construction of Geologic Time Scale 2004 (GTS2004) incorporated different techniques depending on the data available within each interval. Construction involved a large number of specialists, including contributions by past and present subcommissions officers of the International Commission on Stratigraphy (ICS), geochemists working with radiogenic and stable isotopes, stratigraphers using diverse tools from traditional fossils to astronomical cycles to database programming, and geomathematicians. Anticipated advances during the next four years include formalization of all Phanerozoic stage boundaries, orbital tuning extended into the Cretaceous, standardization of radiometric dating methods and resolving poorly dated intervals, detailed integrated stratigraphy for all periods, and on-line stratigraphic databases and tools. The geochronological science community and the International Commission on Stratigraphy are focusing on these issues. The next version of the Geologic Time Scale is planned for 2008, concurrent with the planned completion of boundary-stratotype (GSSP) definitions for all international stages.
Article
We present a revised Neogene geochronology based upon a best fit to selected high temperature radiometric dates on a number of identified magnetic polarity chrons (within the late Cretaceous, Paleogene, and Neogene) which minimizes apparent accelerations in sea-floor spreading. An assessment of first order correlations of calcareous plankton biostratigraphic datum events to magnetic polarity stratigraphy yields the following estimated magnetobiochronology of major chron- ostratigraphic boundaries: Oligocene/Miocene (Chron C6CN): 23.7 Ma; Miocene/Pliocene (slightly younger than Gilbert/Chron 5 boundary): 5.3 Ma; Pliocene/Pleistocene (slightly younger than Olduvai Subchron): 1.6 Ma. Changes to the marine time-scale are relatively minor in terms of recent and current usage except in the interval of the middle Miocene where new DSDP data reveal that previous correlations of magnetic anomalies 5 and 5A to magnetic polarity Chrons 9 and 11, respectively, are incorrect. Our revized magnetobiostratigraphic correlations result in a 1.5-2 m.y. shift towards younger magneto- biochronologic age estimate in the middle Miocene. Radiometric dates correlated to bio- and magnet- ostratigraphy in continental section generally support the revized marine magnetobiochronology presented here. Major changes, however, are made in marine-non-marine correlations in the Miocene in Eurasia which indicate African-Eurasian migrations through the Persian Gulf as early as 20 Ma. The 12.5 Ma estimate of the Hipparion datum is supported by recent taxonomic revisions of the hipparions and magnetobiostratigraphic correlations which show that primitive hipparions first arrived in Eurasia and North Africa at c.12.5 Ma and a second wave in the tropics (i.e. Indian and central Africa) at c. 10 Ma.